ISSN (Online) 1878-9080 RESEARCH ARTICLE www.ingentaconnect.com/content/nhn/pimj https://doi.org/10.3767/persoonia.2019.42.06 A phylogenetic and taxonomic revision of sequestrate Russulaceae in Mediterranean and temperate Europe J.M. Vidal1*, P. Alvarado2*, M. Loizides3, G. Konstantinidis4, P. Chachuła5, P. Mleczko6, G. Moreno7, A. Vizzini8, M. Krakhmalnyi9, A. Paz10, J. Cabero11, V. Kaounas12, M. Slavova13, B. Moreno-Arroyo14, J. Llistosella15 Key words Abstract A comprehensive morphological and genetic study of type material and new collections of sequestrate Russulales species formerly belonging to the genera Arcangeliella, Elasmomyces, Gymnomyces, Hydnangium, angiocarpic Hymenogaster, Macowanites, Martellia, Secotium and Zelleromyces is here undertaken, for the purpose of providing Arcangeliella a complete taxonomical revision of sequestrate Russulaceae species in the Mediterranean and temperate regions Elasmomyces of Europe. As a result, seven distinct taxa in the genus Lactarius and 18 in the genus Russula are identified. Six Europe of them are new species: L. populicola, L. subgiennensis, R. bavarica, R. candidissima, R. hobartiae and R. medi-Gymnomyces terraneensis, and seven represent new combinations: L. josserandii (≡ Zelleromyces josserandii), L. soehneri Lactarius (≡ Hydnangium soehneri), R. candida (≡ Hydnangium candidum), R. cerea (≡ Hydnangium cereum), R. messapica Macowanites var. messapicoides (≡Macowanites messapicoides), R. meridionalis (≡Zelleromyces meridionalis) and R. neuhoffii Martellia (≡ Hydnangium neuhoffii). Twenty-two of the 25 taxa are illustrated, while descriptions, microscopy images, as well pseudoangiocarpic as extensive information on the ecology, chorology and phylogeny for all taxa are provided. A key is further included taxonomy to facilitate their identification. Zelleromyces Article info Received: 2 July 2018; Accepted: 4 December 2018; Published: 5 April 2019. INTRODUCTION Sequestrate Russulaceae produce pseudoangiocarpic (‘secotioid’) basidiomata, where the stipe is external and more or less reduced, to angiocarpic (‘gasteroid’) basidiomata, where the stipe is internal, reduced to a columella, or altogether absent. Sequestrate Russulaceae seem to be widespread throughout the globe, especially in arid and semi-arid regions of Australia and New Zealand (Bougher 1997, Lebel 1998, 2001, 2002, 2003a, b, Lebel & Trappe 2000, Bougher & Lebel 2001, Lebel & Castellano 2002, Trappe & Claridge 2003, Lebel & Tonkin 2007) and North America (Zeller & Dodge 1919, 1937, Singer 1 C/ Massaballs 10, E-17118 Sant Sadurní de l’Heura, Girona, Spain. 2 ALVALAB, La Rochela 47, E-39012 Santander, Spain; corresponding author e-mail: pablo.alvarado@gmail.com. 3 P.O. Box 58499, 3734 Limassol, Cyprus. 4 Agiou Kosma 25, 51100 Grevena, Greece. 5 Pieniny National Park, Jagiellońska 107b, PL-34-450 Krościenko nad Dunajcem, Poland. 6 Institute of Botany, Jagiellonian University in Krak, Gronostajowa 3, PL-30-387 Krak, Poland. 7 Dept. de Ciencias de la Vida (Botánica), Facultad de Biología, Universidad de Alcalá, E-28871 Alcalá de Henares, Spain. 8 Dipt. de Scienze della Vita e Biologia dei Sistemi, Università di Torino, Viale P.A. Mattioli 25, I-10125 Torino, Italy. 9 Institute of Evolution and Department of Evolutionary and Environmental Biology, University of Haifa, 199 Abba Khoushy Ave, Haifa, 3498838, Israel. 10 C/ Vall Ter 791, UrbanitzaciLlac del Cigne, E-17455 Caldes de Malavella, Girona, Spain. 11 C/ El Sol 6, E-49800 Toro, Zamora, Spain. 12 Sokratous 40, TK-19016 Artemis Attika, Greece. 13 Krivolak 4, 1164 Sofia, Bulgaria. 14 Dept. de Biología Vegetal, Facultad de Biología, Universidad de Cdoba, Colonia San José 4, Campus de Rabanales, E-14014 Cdoba, Spain. 15 Dept. de Biologia Evolutiva, Ecologia i Ciències Ambientals, Universitat de Barcelona, Av. Diagonal 643, E-08028 Barcelona, Spain. * These authors contributed equally. & Smith 1960, Smith 1963, Miller & Lebel 1999, Fogel & States 2001, Desjardin 2003, Smith et al. 2006). However, several sequestrate species have also been documented in tropical forests in Africa (Dring & Pegler 1978, Eberhardt & Verbeken 2004, Verbeken & Walleyn 2010, Beenken et al. 2016) and Asia (Corner & Hawker 1953, Heim 1959, Zhang & Yu 1990, Tao et al. 1993, Verbeken et al. 2014a, b), as well as in temperate Nothofagus forests in Patagonia (Trierveiler-Pereira et al. 2015). The generic organization of sequestrate taxa within the Rus-sulaceae has long been a subject of debate (Singer & Smith 1960, Pegler & Young 1979, Beaton et al. 1984, Zhang & Yu 1990). The subtle differences between Gymnomyces, Martellia, Cystangium, Elasmomyces and Macowanites, as well as the latex-bleeding Arcangeliella and Zelleromyces, led to different taxonomic arrangements to accommodate an increasing number of emerging species (Lebel & Trappe 2000, Trappe et al. 2002, Vidal 2004a). Early authors hypothesized a single evolutionary lineage from the gymnocarpic to the angiocarpic genera in the Russulales (Malençon 1931, Heim 1938, Singer & Smith 1960, Oberwinkler 1977), but genetic studies revealed that sequestrate russuloid lineages are polyphyletic and nest within the genera Russula and Lactarius (Calonge & Martín 2000, Miller et al. 2001, Larsson & Larsson 2003, Eberhardt & Verbeken 2004, Nuytinck et al. 2003, Shimono et al. 2004, Smith et al. 2006, Lebel & Tonkin 2007). The split of Lactarius due to genetic evidence (Buyck et al. 2008, 2010) led to the restoration of the old genus Lactifluus and the erection of Multifurca, although no sequestrate species are yet known to occur in either of these two lineages. The genus Gymnomyces, typified by G. pallidus (Massee 1898), a species found in myrtaceous forests whose current name is R. paneeroides (Lebel 2017), was characterized by the absence of a stipe-columella and the presence of sphaerocytes in the hymenophoral trama. The European species G. xanthosporus © 2019 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute You are free to share - to copy, distribute and transmit the work, under the following conditions: Attribution: You must attribute the work in the manner specified by the author or licensor (but not in any way that suggests that they endorse you or your use of the work). Non-commercial: You may not use this work for commercial purposes. No derivative works: You may not alter, transform, or build upon this work. For any reuse or distribution, you must make clear to others the license terms of this work, which can be found at http://creativecommons.org/licenses/by-nc-nd/3.0/legalcode. Any of the above conditions can be waived if you get permission from the copyright holder. Nothing in this license impairs or restricts the author’s moral rights. was originally described by Hawker (1952) as a variety of Hydnangium carneum, to accommodate some specimens found in the UK, and subsequently re-combined into Gymnomyces by Smith (1962), because of its amyloid spores and sphaerocytes. Several species were described later from the Mediterranean basin or re-combined into this genus, such as G. ilicis (Llistosel-la & Vidal 1995), G. dominguezii (Moreno-Arroyo et al. 1999, and subsequently validated in Calonge 2000), G. sublevisporus (Moreno-Arroyo et al. 2002) and G. meridionalis (Vidal 2004a), the latter originally treated in the genus Zelleromyces by Moreno-Arroyo et al. (1998b, and subsequently validated in Calonge 2000). The genus Martellia, typified by M. mistiformis (Mattirolo 1900), was first described from several collections found in Sardinia (Italy) under Quercus suber. It was characterized by the absence of a stipe-columella and lack of sphaerocytes in the hymenophoral trama (Singer & Smith 1960). In the late 20th century, new species and combinations were proposed: M. mediterranea (Moreno et al. 1991), M. pila (Vidal 1991b), M. monospora (Astier & Pacioni 1998), the last two originally described from France in the genus Hydnangium by Patouillard (1910) and Boudier & Patouillard (1888) respectively, and M. aurantiaca (Astier & Pacioni 1998), originally described from Spain in the genus Hydnangium by Heim et al. (1934) and later re-combined into the genus Stephanospora by Vidal (2004c). After comparing the types of Gymnomyces and Martellia, Lebel & Trappe (2000) considered that the presence or absence of sphaerocytes in the hymenophoral trama was of no taxonomic relevance in keeping the two genera separate, and subsumed Martellia into Gymnomyces (Trappe et al. 2002). The genus Elasmomyces, typified by E. mattiroloanus (Cavara 1897), was proposed to accommodate some stipitate collections obtained under Abies alba in Italy, that lacked sphaerocytes in the hymenophoral trama. Saccardo & Saccardo (1905) proposed the combinations E. krjukowensis and E. michailowskianus for two species of the genus Secotium, originally found in Russia by Bucholtz (1901, 1903). The genus Macowanites, typified by M. agaricinus (Kalchbren-ner 1882), which was originally placed in the genus Macowania (Kalchbrenner 1876), was first proposed for a stipitate fungus found in Eastern Cape (South Africa), exhibiting sphaerocytes in the hymenophoral trama. Much later, some Mediterranean species were added or re-combined into this genus, namely M. galileensis from Mount Carmel, Israel (Moser et al. 1977) and M. messapicoides (Llistosella & Vidal 1995), M. ammophlilus (Vidal et al. 2002), originally described as Gymnomyces by Calonge & Vidal (1999), and M. vinaceodorus (Calonge & Vidal 2001), all from Spain. Recently, M. candidus (Vidal 2004b), a species originally described from Northern France in the genus Hydnangium (Tulasne & Tulasne 1843), was also placed into this genus. Same as for the angiocarpic genera Gymnomyces and Martellia, the pseudoangiocarpic taxa Elasmomyces and Macowanites were also considered synonyms by Lebel & Trappe (2000), following comparison of their type collections. The genus Cystangium, typified by C. sessile (Singer & Smith 1960), is another russuloid sequestrate genus, but appears to be confined to Australia and South America, with no known representatives in Europe. It is characterized by an epithelial pileipellis, in contrast to Gymnomyces, Martellia, Elasmomyces and Macowanites, all of which have a trichodermal or undifferentiated pileipellis. The stipe in this genus can be external, reduced to a columella, or completely absent (Lebel & Trappe 2000). The latex-bleeding genus Arcangeliella was typified by A. bor-ziana (Cavara 1900), and proposed for collections found under Abies alba in central Italy. Arcangeliella borziana is a sessile species that sometimes presents a poorly developed or residual stipe and a percurrent columella, especially in young basidiomata. Despite these features, Arcangeliella was chosen by Singer & Smith (1960) to describe several non-European sequestrate species with a well-developed stipe-columella, whereas Vidal (2004a) maintained this genus only for angiocarpic, sessile spe-cies, keeping Gastrolactarius for pseudoangiocarpic, stipitate species. American mycologists (Dodge 1931, Zeller & Dodge 1935) combined several European taxa into Arcangeliella, e.g., A. asterosperma var. asterosperma, A. asterosperma var. depauperata, A. asterosperma var. hololeuca, A. laevis or A. stephensii; the first four, however, belonging to the order Boletales, have since been transferred to the genus Octaviania (Paz et al. 2016). Finally, Mader & Mader (1992) proposed A. volemoides for a collection made under Picea abies in the Austrian Alps, but this species was considered a synonym of A. borziana by Nuytinck et al. (2003) and Vidal (2004a). The latex-bleeding genus Zelleromyces, typified by Z. cinnabarinus (Singer & Smith 1960), was proposed for angiocarpic species devoid of a stipe-columella. Several Mediterranean and Central European species were later described or re-com bined into this genus: Z. josserandii from Morocco and France (Malençon 1975), Z. hispanicus (Calonge & Pegler 1998) and Z. giennensis (Moreno-Arroyo et al. 1998a) from Spain, and Z. soehneri (Trappe et al. 2002), originally described from Germany in the genus Hydnangium by Zeller & Dodge (1935). Recently, several taxa previously placed in the sequestrate ge-nera Cystangium, Elasmomyces, Gymnomyces, Macowanites, and Martellia, were re-combined into Russula by Lebel (2017) and Elliott & Trappe (2018). About 30 sequestrate species of Russulaceae have been described from Mediterranean and temperate regions of Europe, but the generic placement for many of these still remains unclear. In the present work, all of them are reviewed and revised in the light of extensive genetic data. New sequences were produced from herbarium speci-mens and new collections of pan-European origins, in order to evaluate the phylogenetic basis of characters traditionally employed to discriminate among sequestrate taxa, and propose the most appropriate supraspecific affiliation of these species within Lactarius and Russula. MATERIAL AND METHODS Fungal collections Morphological studies include type material and old representative collections of selected species placed in the genera Arcangeliella, Elasmomyces, Gymnomyces, Hydnangium, Hymenogaster, Macowanites, Martellia, Octaviania, Secotium and Zelleromyces kept in the public herbaria of BPI (Beltsville, USA), FH (Cambridge, USA), IB (Innsbruck, Austria), K (Kew, UK), M (Munich, Germany), MA (Madrid, Spain), MPU (Montpellier, France), NICE (Nice, France), NY (New York, USA), PC (Paris, France), PRM (Praha, Czech Republic), UPS (Uppsala, Sweden) and WU (Wien, Austria), as well as recent collections from different localities in Europe and the Middle East (Bulgaria, Cyprus, France, Germany, Greece, Hungary, Israel, Italy, Poland, Portugal and Spain) kept in the public herbaria of AH (Alcalá de Henares, Spain), BCN (Barcelona, Spain), HAI (Haifa, Israel), KRA (Krak, Poland), MCVE (Venezia, Italy), or in private herbaria of the Asociaci Vallisoletana de Micología (AVM), B. Moreno Arroyo (BM), G. Konstantinidis (GK), A. Paz Conde (IC), J. Cabero (JC), J.M. Vidal (JMV), M. Loizides (ML), M. Slavova (MSL), V. Kaounas (VK) and other contributors to this paper (see acknowledgements). Newly collected specimens have been deposited in the Herbarium of the Universidad de Alcalá (AH), in the Herbarium of J.M. Vidal et al.: Revision of sequestrate Russulaceae the CeDocBiV of the Universitat de Barcelona (BCN-myc), in the Herbarium of the Jagiellonian University in Krak (KRA), and in the Mycological Collection of the Bulgarian Academy of Sciences (SOMF). Morphological study Most species are fully described, with the exception of the re-cently described species, for which only a summary description is provided. Detailed macromorphological notes and images were taken from the majority of collections, and the presence/ absence of latex and columella, as well as ecological data (altitude, associated vegetation, putative ectomycorrhizal partner, substrate) were annotated. Colour of basidiomata, hymenophore, and spore mass observed in locules was de-scribed according to Kornerup & Wanscher (1978). Methuen code colour of fresh specimens was obtained from digital images. For microscopical features, a Carl Zeiss Jena Jena-val microscope with a DIC device, a Leica BM E binocular, and a Leica DMRB with condenser P 1.40 OIL S1, provided with planapochromatic optics 25×, 40×, 50×, 63× and 100×, were employed, coupled with Reflex Pentax K-20 (14.6 Mpx) cameras with remote control. Amyloid reaction was checked from mounts in Melzer’s reagent, while ammoniacal (10 %) or potassium hydroxide (2–5 %) Congo red was employed to enhance contrast in microscopical observations. The terminology of Heilmann-Clausen et al. (1998) is adopted for naming the pileipellis structure. For the different types of secre-tory hyphae, the terminology of Clémençon (2004) is adopted, using the terms: 1) ‘laticifera’ for non-septate, latex-containing hyphae of granular content (crystalloid in exsiccata), staining violet in contact with sulfovainillin, typical of Lactarius (Fig. 12c, f); 2) ‘gloeoplera’ for the septate hyphae not bleeding latex, also of granular or crystalloid content, not staining violet with sulfovainillin, which are found in the trama and context of Russula and generate both fertile (basidia) and sterile elements (macrocystidia and dermatocystidia) (Fig. 15f); and 3) ‘thromboplera’ for the tortuose hyphae of homogeneous, non-granular, yellow to brownish content, found in both Lactarius and Russula (Fig. 10b–c, 12c). We use the term ‘cystidiole’ (Romagnesi 1944, Josserand 1983) for sterile hymenial elements similar to dermatocystidia firstdeveloping in external locules atthe same level as basidioles, and sometimes remaining in the mature hymenium as in Russula mistiformis (Fig. 25j–l). We apply the term ‘paraphysoid cells’ (Hawksworth et al. 1995) to cylindrical or slightly clavate and sometimes septate sterile elements with-out granular content, present in the hymenium of many of the sequestrate species studied, e.g., sterile cells similar to septate basidioles found in the hymenium of Lactarius josserandii (Fig. 8g–h), called ‘paracystidia’ by Malençon (1975), or those in the hymenium of L. borzianus and L. stephensii (Fig. 11b–c), called ‘hymenial hairs’ or ‘paracystidia’ in Vidal (2004a). The measurements of spores and other elements were obtained with the aid of ‘Mycometre VA’ software (G. Fannechère). Spores were measured in side view, always excluding ornamentation and hilar appendix, and at least 20 spores from each specimen were studied, indicating the minimum and maximum average and the Q-value. Scanning electron microscopes (SEM) Hitachi S-4100 and Zeiss DSM-960A (Unitat de Microscia dels Serveis Tècnics de Recerca de la Universitat de Girona, UdG), Hitachi S-4700 (Laboratory of Scanning Microscopy, Institute of Geological Sciences, Jagiellonian University in Krak, UJ) and Zeiss DSM-950 (Servicio de Microscopía Electrica de la Universidad de Alcalá de Henares, AH), were also employed for imaging spores. Light microscopy images of spores were stacked digitally with the aid of ‘Helicon Focus’ (Helicon Soft Ltd.). Molecular analysis DNA extraction, amplification and sequencing DNA extraction and PCR amplification were performed as described by Alvarado et al. (2012). Primers ITS1F and ITS4 (White et al. 1990, Gardes & Bruns 1993) were used for the ITS region; primers LR0R and LR5 (Vilgalys & Hester 1990) were used for the 28S rDNA ribosomal region (28S rDNA); bRPB2-6F2 (reverse from bRPB2-6R2) and bRPB2-7.1R2 (modified from bRPB2-7.1R) for the DNA-directed second largest subunit two of RNA polymerase II (rpb2) gene (Matheny 2005, Matheny et al. 2007, Gelardi et al. 2015); and in some samples, EF1-983F and EF1-1567R (Rehner & Buckley 2005) for the translation elongation factor 1α (tef1) gene for both, PCR and sequencing. Sequences were edited for errors in MEGA v. 5 (Tamura et al. 2011). New sequences are listed in Table 1 highlighted in bold characters, and newly sequenced speci-mens are marked with an asterisk symbol (*) in the paragraph of material studied. Attempts to obtain genetic data from some old specimens were conducted without success. Phylogenetic analyses Newly generated sequences were aligned with the most similar sequences in the International Nucleotide Sequence Database Collaboration (INSDC) identified through BLASTn searches (Altschul et al. 1997). A multigenic ITS – 28S rDNA – rpb2 alignment of Russulaceae was built including all lineages identified in the most recent phylogenetic works (Kong et al. 2015, Looney et al. 2016), as well as all sequences available originated from sequestrate species. Reference publications for the sequences employed are provided in Table 1. Stereum hirsutum, Auriscalpium vulgare, Amylostereum laevigatum and Echinodontium tinctorium were employed as outgroups for the general alignment because these are known to be the closest lineages outside the Russulaceae clade (Buyck et al. 2008). Sequences were first aligned in MEGAv. 5 (Tamura et al. 2011) with Clustal W (Higgins et al. 1994) and the resulting alignments edited manually. First, a full query for the best-scoring maximum likelihood tree was performed on the general alignment in RAxML (Stamatakis 2006) using the standard search algo-rithm (ITS, 28S rDNA, rpb2 data partitioned, 2000 bootstrap replications). Four independent alignments were taken from the general one: 1) Lactarius; 2) Russula sect. Ingratae; 3) Russula sect. Rigidae; and 4) Russula s.str., using Lactifluus piperatus as outgroup for all of them. GBlocks v. 0.91b (Castresana 2000) was employed with the most permissive conditions to remove ambiguously aligned positions from ITS rDNA data. Bayesian analyses were performed in MrBayes v. 3.2.6 (Ronquist & Huelsenbeck 2003) employing optimal models determined for ITS, 28S rDNA and rpb2 partitions in MrModeltest v. 2.3 (Nylander 2004) package loaded in PAUP v. 4.0b10, two simultaneous runs, six chains, temperature set to 0.2, and sampling every 100th generation until convergence parameters (< 0.01) were met after 9.35 M generations in Lactarius, 10.4 M in Russula sect. Ingratae, 0.74 M in Russu-la sect. Rigidae and 8.64 M generation in Russula s.str. The first 25 % trees were discarded as burn-in. Fig. 1 represents the phylogenetic tree resulting from the ML analysis of all sequences produced in the present work and retrieved from public databases with the major clades condensed. Fig. 2–5 were obtained from the previous one, collapsing all nodes except a single major clade: Lactarius (Fig. 2), Russula sect. Ingratae (Fig. 3), Russula sect. Rigidae (Fig. 4) and Russula s.str. (Fig. 5). Nodes were annotated with support values from Maximum Likelihood and Bayesian analyses if at least one of these was consideredsignificant, or annotated inbrackets when at least one of these was considered subsignificant. Support values were considered significant when ML bootstrap (BP) values were above 70 % or posterior probability (PP) values Table 1 Specimens used in molecular phylogenetic studies and their GenBank accession numbers. Species (*) sequestrate1 Original identification (*) type Origin Voucher Id (Herbarium)2 Reference GenBank accession codesITS rDNA 28S rDNA RPB2 Eurussuloid clade/amylostereaceae Amylostereum laevigatum Amylostereum laevigatum France olrim409, 623.84 (CBS) Wang et al. (2004) AY781246 AF287843 AY218469 /auriscalpiaceae Auriscalpium vulgare Auriscalpium vulgare USA PBM944 (WTU) Lutzoni et al. (2004) DQ911613 DQ911614 – Echinodontium tinctorium Echinodontium tinctorium USA 16666 (DAOM) Binder & Hibbett (2002) AY854088 AF393056 AY218482 /russulaceae/lactarius Lactarius sp. uncultured Kohout (unpubl.) UDB027052 – – Lactarius sp. * Arcangeliella sp. Tasmania BH2148F, T991 (HO) Horton (2011) JF960610 – – Zelleromyces sp. Tasmania BH1714P (HO) Horton (2011) JF960854 – – Zelleromyces sp. Tasmania BH3528R (HO) Horton (2011) JF960852 – – Zelleromyces sp. Tasmania BH2174F (HO) Horton (2011) JF960853 – – Lactarius acris Lactarius acris Germany EU014 (UPS) Buyck et al. (2008) DQ421988 DQ421988 DQ421922 Lactarius akahatsu Lactarius akahatsu Thailand AV2004-141 (GENT) Verbeken et al. (2014b) KF133269 KF133301 KF133333 Lactarius angiocarpus* Lactarius angiocarpus* Zambia DA00-448 (GENT) Eberhardt & Verbeken (2004), Stubbe et al. (2010) AY606942 DQ421981 GU258310 Lactarius atroviridis Lactarius atroviridis USA AV05-306 (GENT) Verbeken et al. (2014b) KF133270 KF133302 KF133334 Lactarius auriolla Lactarius auriolla Sweden RW1601 (GENT) Verbeken et al. (2014b) KF133257 KF133288 KF133321 Lactarius azonites Lactarius azonites Belgium DS08-520 (GENT) Stubbe & Verbeken (2012) JQ446096 JQ446168 JQ446231 Lactarius baliophaeus Lactarius baliophaeus Malawi AV05-155 (GENT) Verbeken et al. (2014b) GU258277 GU265576 GU258312 Lactarius bisporus* Lactarius bisporus* Thailand FH12-160 (GENT) Verbeken et al. (2014a) KF856293 – – Lactarius borzianus* Arcangeliella borziana France JMV800279b (BCN) This work MK105610 MK105696 MK102740 Arcangeliella borziana Italy 16944 (MCVE) Osmundson et al. (2013) JF908775 – – Arcangeliella borziana Switzerland FA96-05-3344 (WSL) Peter et al. (2001) AF286204 AF286203 – Lactarius camphoratus Lactarius camphoratus Sweden UE04.09.2004 (UPS) Verbeken et al. (2014b) DQ422009 DQ422009 DQ421933 Lactarius chrysorrheus Lactarius chrysorrheus Italy UE04.10.2002-8 (UPS) Verbeken et al. (2014b) KF133261 KF133293 KF133325 Lactarius citriolens Lactarius citriolens Sweden UE20.09.2004-03 (UPS) Buyck et al. (2008) DQ422003 DQ422003 DQ421931 Lactarius crassiusculus Lactarius crassiusculus Thailand HTL369 (GENT) Verbeken et al. (2014b) EF560684 KF133303 KF133335 Lactarius crassus* Arcangeliella crassa USA T17996 Gordon et al. (unpubl.) KT968581 KT968636 – Arcangeliella crassa USA 41826 (OSC) Gordon et al. (unpubl.) KT968563 KT968615 – Lactarius cyanescens Lactarius cyanescens Malaysia DS06-058 (GENT) Stubbe et al. (2010) GU258278 JN388999 JN375602 Lactarius cyathuliformis Lactarius cyathuliformis Sweden UE04.09.2004-2 (UPS) Verbeken et al. (2014b) KF133266 KF133298 KF133330 Lactarius deliciosus Lactarius deliciosus Slovakia JN2001-046 (GENT) Verbeken et al. (2014b) KF133272 KF133305 KF133337 Lactarius echinellus* Lactarius echinellus Sri Lanka AV07-169 (GENT) Verbeken et al. (2014b) KF133287 KF133320 KF133352 Lactarius echinus* Lactarius echinus* Sri Lanka AV07-168 (GENT) Verbeken et al. (2014b) KF133273 KF133306 KF133338 Lactarius evosmus Lactarius evosmus Sweden 536 (UP) Nygren et al. (2007) DQ658882 – – Lactarius falcatus* Lactarius falcatus* Thailand KVP08-038 (GENT) Verbeken et al. (2014b) KF133274 KF133307 KF133339 Lactarius flavopalustris Lactarius flavopalustris Finland JV23334 (TURA) Barge et al. (2016) KR090904 – KR090982 Lactarius flexuosus Lactarius flexuosus Sweden UE06.09.2002-1 (UPS) Buyck et al. (2008) DQ421992 DQ421992 DQ421925 Lactarius fuliginosus Lactarius fuliginosus Sweden MTB97-24 (GENT) De Crop et al. (2017) JQ446111 JQ446180 JQ446240 Lactarius fulvissimus Lactarius fulvissimus Germany JKLAC10082002 (GENT) Wisitrassameewong et al. (2014) KF432970 – KR025662 Lactarius fulvissimus Germany JN2012-025 (GENT) Wisitrassameewong et al. (2014) KR025576 – KR025661 Lactarius fumosibrunneus Lactarius fumosibrunneus Mexico EG10 (XAL) Garay-Serrano et al. (2012) JN797633 – – Lactarius fumosus Lactarius fumosus Canada HRL0894 Bérubé et al. (unpubl.) KJ705224 – – Lactarius gardneri* Zelleromyces gardneri USA 1537 (SOC) Southworth (2016) JN022500 – – Zelleromyces gardneri USA 513 (SOC) Frank et al. (2006) DQ453696 – – Lactarius giennensis* Zelleromyces giennensis* Spain Fungi 38674 (MA) Calonge & Martín (2000) AF230900 – – Zelleromyces giennensis Spain AVM 1615, JMV800629 (BCN) This work MK105611 MK105697 – Zelleromyces giennensis Spain JC20061020 (pers. herb.) This work MK105612 MK105698 MK102741 Lactarius glyciosmus Lactarius glyciosmus USA 20923 (DGB) Barge et al. (2016) KR090908 – KR090985 Lactarius haugiae Lactarius haugiae Mexico LM4994 (XAL) Bandala et al. 2016 KT583642 KT583649 KT736507 Lactarius helvus Lactarius helvus Sweden UE08.09.2004-1 (UPS) Verbeken et al. (2014b) KF133263 KF133295 KF133327 Lactarius hispidulus Lactarius hispidulus Guinea AB152 (GENT) Verbeken et al. (2014b) KF133258 KF133289 KF133322 J.M. Vidal et al.: Revision of sequestrate Russulaceae Table 1 (cont.) Species (*) sequestrate1 Original identification (*) type Origin Voucher Id (Herbarium)2 Reference GenBank accession codes ITS rDNA 28S rDNA RPB2 Lactarius horakii Lactarius horakii* Indonesia EH8336 (ZT) Nuytinck et al. (2007) EF685069 – – Lactarius josserandii* Lactarius josserandii Spain JMV800621 (BCN) This work MK105613 MK105699 MK102742 Zelleromyces hispanicus Spain Fungi 37497 (MA) Calonge & Martín (2003) AF231911 – – Zelleromyces hispanicus* Spain Fungi 37498 (MA) Calonge & Martín (2003) AF231912 – – Zelleromyces hispanicus Spain Fungi 53339 (MA) Calonge & Martín (2003) AJ555567 – – Lactarius kabansus Lactarius kabansus Zimbabwe AV99-205 (GENT) Verbeken et al. (2014b) KF133259 KF133291 KF133323 Lactarius lanceolatus Lactarius lanceolatus* USA F 4239 (VPI) Barge et al. (2016) KR090915 – KR090989 Lactarius lignyotus Lactarius lignyotus USA PBM2424 (CUW) Matheny et al. (2007) DQ221107 AY631898 DQ408128 Lactarius lignyotus Sweden UE06.09.2003-5 (UPS) Buyck et al. (2008) DQ421993 DQ421993 DQ421926 Lactarius lilacinus Lactarius lilacinus Belgium RW3774 (GENT) Verbeken et al. (2014b) KF133275 KF133308 KF133340 Lactarius luculentus var. laetus Lactarius luculentus var. laetus USA F 024643 (DBG) Barge et al. (2016) KR090922 – KR090994 Lactarius luridus Lactarius luridus Belgium OB11-011 (GENT) Verbeken et al. (2014b) KF241547 – – Lactarius mammosus Lactarius mammosus Sweden UE09.09.2004-5 (UPS) Verbeken et al. (2014b) KF133265 KF133297 KF133329 Lactarius montoyae Lactarius montoyae* India KD1065 (BSD) Stubbe et al. (2010) EF560673 GU265641 GU258380 Lactarius nanus Lactarius nanus USA EB106-13 (MONT) Barge et al. (2016) KR090928 – KR091000 Lactarius necator Lactarius necator France AV04-231 (GENT) Verbeken et al. (2014b) KF133276 KF133309 KF133341 Lactarius olympianus Lactarius olympianus USA ED08-018 (GENT) De Crop et al. (2017) KR364089 KR364220 KR364320 Lactarius pallidomarginatus Lactarius pallidomarginatus USA CLC1470 (MONT) Barge et al. (2016) KR090939 KR090939 KR091009 Lactarius peckii Lactarius peckii USA JN2004-020 (GENT) Verbeken et al. (2014b) KF133277 KF133310 KF133342 Lactarius pomiolens Lactarius pomiolens* Sri Lanka AV07-159 (GENT) Verbeken et al. (2014b) KF133282 KF133315 KF133347 Lactarius populicola* Lactarius stephensii Belgium RW2930 (GENT) Nuytinck et al. (2003) AY331012 – – Lactarius populicola* Greece GK4831, JMV800648 (BCN) This work MK105614 MK105700 MK102743 Lactarius psammicola Lactarius psammicola USA BPL869 (TENN) Looney et al. (2016) KY848507 – – Lactarius pterosporus Lactarius pterosporus France PAM06100705 (GENT) Stubbe & Verbeken (2012) JQ446153 – JQ446275 Lactarius quieticolor Lactarius quieticolor Sweden UE10.09.2004-1 (UPS) Buyck et al. (2008) DQ422002 DQ422002 DQ421930 Lactarius quietus Lactarius quietus Sweden UE16.09.2004 (UPS) Verbeken et al. (2014b) KF133264 KF133296 KF133328 Lactarius repraesentaneus Lactarius repraesentaneus USA CLC2318 (MONT) Barge et al. (2016) KR090948 – KR091020 Lactarius romagnesii Lactarius romagnesii France UE29.09.2002-6 (UPS) Buyck et al. (2008) DQ421989 DQ421989 DQ421923 Lactarius ruginosus Lactarius ruginosus Italy DS09-613 (GENT) Stubbe & Verbeken (2012) JQ446150 JQ446216 JQ446272 Lactarius salicis-reticulatae Lactarius salicis-reticulatae Finland CLC2885 (MONT) Barge et al. (2016) KR090959 – KR091030 Lactarius salmonicolor Lactarius salmonicolor Germany UE158/FO46879 (TUB) Eberhardt et al. (2000) AF140265 AF325284 – Lactarius saturnisporus* Lactarius saturnisporus* Sri Lanka AV07-170 (GENT) Verbeken et al. (2014b) KF133283 KF133316 KF133348 Lactarius scrobiculatus Lactarius scrobiculatus Italy 732 (MCVE) Osmundson et al. (2013) JF908281 – – Lactarius shoreae* Lactarius shoreae* Sri Lanka AV07-164 (GENT) Verbeken et al. (2014b) KF133278 KF133311 KF133343 Lactarius silviae* Arcangeliella camphorata USA 64481 (OSC) Gordon (unpubl.) EU834192 – – Arcangeliella camphorata USA 74231 (OSC) Gordon (unpubl.) EU644700 EU652366 – Lactarius soehneri* Lactarius soehneri Spain 39272 (AH) This work MK105615 MK105701 – Lactarius soehneri Spain 46013 (AH) This work MK105616 MK105702 MK102744 Lactarius sphagneti Lactarius sphagneti UK PL2805 (pers. herb.) Verbeken et al. (2014b) KF133268 KF133300 KF133332 Lactarius spinosporus* Lactarius spinosporus* China FAN445 (BJTC) Sang et al. (2016) KY270490 KY270494 – Lactarius spinulosus Lactarius spinulosus Sweden AT2003068 (UPS) Verbeken et al. (2014b) KF133262 KF133294 KF133326 Lactarius stephensii* Lactarius stephensii Poland F-2014-147 (KRA) This work MK105617 –– Arcangeliella stephensii Spain JMV20010609-1 (BCN) This work MK105618 MK105703 MK102745 Zelleromyces stephensii UK Kew(M)64067 (K) Brock et al. (2009) EU784439 – – Lactarius subdulcis Lactarius subdulcis Belgium JV2006-024 (GENT) Verbeken et al. (2014b) KF133279 KF133312 KF133344 Lactarius subgiennensis* Lactarius subgiennensis* Cyprus ML211152E, JMV800627 (BCN) This work MK105619 MK105704 MK102746 Lactarius subgiennensis Cyprus ML61132Z (pers. herb.) This work MK105620 –– Lactarius subgiennensis Cyprus ML411172Z (pers. herb.) This work MK105621 MK105705 MK102747 Lactarius subsericatus Lactarius subsericatus Sweden UE11.10.2004-8 (UPS) Buyck et al. (2008) DQ422011 DQ422011 DQ421934 Lactarius tenellus Lactarius tenellus Benin ADK3598 (BR) Verbeken et al. (2014b) KF133280 KF133313 KF133345 Lactarius thyinos Lactarius thyinos Canada AVoitk23-08-2004 (GENT) Verbeken et al. (2014b) KF133271 KF133304 KF133336 Lactarius torminosus Lactarius torminosus Czech Republic RW3183 (GENT) Verbeken et al. (2014b) KF133281 KF133314 KF133346 Table 1 (cont.) Species (*) sequestrate1 Original identification (*) type Origin Voucher Id (Herbarium)2 Reference GenBank accession codesITS rDNA 28S rDNA RPB2 Lactarius trivialis Lactarius trivialis Sweden UE27.08.2002-17a (UPS) Buyck et al. (2008) DQ421991 DQ421991 DQ421924 Lactarius uvidus Lactarius uvidus Estonia 185041 (TAA) Tedersoo et al. (2003) AJ534936 AJ534936 – Lactarius vietus Lactarius vietus Sweden UE11.19.2004-1 (UPS) Verbeken et al. (2014b) KF133267 KF133299 KF133331 Lactarius zonarius Lactarius zonarius France UE27.09.2002-4 (UPS) Buyck et al. (2008) EU278678 EU278678 EU278679 Lactarius zonarius var. riparius Lactarius zonarius var. riparius USA CLC2933 (MONT) Barge et al. (2016) KX394302 – – Zelleromyces daucinus* Zelleromyces daucinus Australia T18311 (OSC) Miller et al. (2001) – AF265543 – Zelleromyces striatus* Zelleromyces striatus Australia T18858 (OSC) Miller et al. (2001) – AF265544 – /lactifluus Lactifluus edulis Lactarius edulis Zimbabwe AV99-041 (GENT) Buyck et al. (2007) AY606973 DQ421977 DQ421916 Lactifluus emergens Lactifluus emergens Zimbabwe AV99-012 (GENT) De Crop et al. (2017) KR364021 KR364148 KR364276 Lactifluus longisporus Lactarius longisporus* Zimbabwe AV94-557 (GENT), BB00.1519 (PC) Buyck et al. (2007), De Crop et al. (2017) KR364118 KR364244 DQ421910 Lactifluus nodosicystidiosus Lactarius nodosicystidiosus* Madagascar BB97-072 (PC) Buyck et al. (2007) AY606975 DQ421976 DQ421915 Lactifluus phlebophyllus Lactarius phlebophyllus Madagascar BB00-1388 (PC) Buyck et al. (2007) AY606974 DQ421979 DQ421918 Lactifluus piperatus Lactarius piperatus Sweden UE09.08.2004-6 (UPS) Buyck et al. (2008) DQ422035 DQ422035 DQ421937 Lactifluus vellereus Lactarius vellereus Sweden UE20.09.2004-22 (UPS) Buyck et al. (2007) DQ422034 DQ422034 DQ421936 Lactifluus velutissimus Lactarius velutissimus Zimbabwe AV99-185 (GENT) Buyck et al. (2007) AY606982 DQ421973 DQ421912 Lactifluus volemus Lactarius volemus Sweden UE09.08.2004-5 (UPS) Buyck et al. (2008) DQ422008 DQ422008 DQ421932 /multifurca Multifurca furcata Lactarius furcatus Costa Rica RH7804 (NY) Buyck et al. (2008) DQ421994 DQ421994 DQ421927 Multifurca ochricompacta Russula ochricompacta USA BB02.107 (PC) Buyck et al. (2008) DQ421984 DQ421984 DQ421940 Multifurca stenophylla Multifurca stenophylla Australia CWD584 (MEL) Lebel et al. (2013) JX266628 JX266633 – Multifurca zonaria Russula zonaria* Thailand DED7442 (PC) Buyck et al. (2008) DQ421990 DQ421990 DQ421942 /russula/basal clades Russula acrifolia Russula acrifolia USA r-05065 Davis (unpubl.) JF834363 – JF834460 Russula albonigra Russula albonigra Sweden AT2002064 (UPS) Buyck et al. (2008) DQ422029 DQ422029 DQ421966 Russula archaea Russula aff. camarophylla USA BB2004-255 (PC) Buyck et al. (2017) EU598165 EU598165 – Russula cf. brevipes Russula cf. brevipes Canada F30230 (UBC) Bazzicalupo et al. (2017) KX812841 KX812863 KX813647 Russula camarophylla Russula camarophylla France PAM01081108 (PC) Buyck et al. (2008) DQ421982 DQ421982 DQ421938 Russula cascadensis Russula cascadensis Canada F23910 (UBC) Bazzicalupo et al. (2017) KJ146726 KJ146726 – Russula aff. chloroides Russula aff. chloroides USA r-01002 Davis (unpubl.) JF834332 JF834478 JF834427 Russula compacta Russula compacta USA BPL242 (TENN) Looney et al. (2016) KT933960 KT933819 KT933890 Russula crassotunicata Russula crassotunicata Canada F30159 (UBC) Bazzicalupo et al. (2017) KX812837 KX812861 KX813645 Russula delica Russula delica Belgium FH12-272 (GENT) Wisitrassameewong et al. (2014), De Crop et al. (2017) KF432955 KR364224 KR364340 Russula earlei Russula earlei USA WCR00-412 (PC) Buyck et al. (2008) DQ422025 DQ422025 DQ421963 Russula farinipes Russula farinipes France UE28.09.2002-4 (UPS) Buyck et al. (2008) DQ421983 DQ421983 DQ421939 Russula littoralis Russula littoralis Europe 1222IS87, PAM93071601 (pers. herb.) Miller & Buyck (2002) AY061702 – – Russula metachromatica Russula metachromatica French Guiana MCA1856 (pers. herb.) Smith et al. (2011) JN168745 JN168745 – Russula nigricans Russula nigricans Sweden UE20.09.2004-07 (UPS) Buyck et al. (2008) DQ422010 DQ422010 DQ421952 Russula pallescens Russula pallescens Norway 146/2002 (TUR) Buyck et al. (2008) DQ421987 DQ421987 DQ421941 Russula pallidospora Russula pallidospora Denmark JV02-218 (C) Eberhardt (unpubl.) DQ422032 DQ422032 – Russula polyphylla Russula polyphylla USA BB07.134 (PC) Liu et al. (2015) KP033486 KP033497 KP033508 Russula pumicoidea* Russula pumicoidea* Australia T14771 (MEL, OSC) Lebel & Tonkin (2007) EU019931 EU019931 – Russula sinuata* Russula sinuata* Australia H4755 (HO) Lebel & Tonkin (2007) EU019943 – – /ingratae/rigidae/russula s.str. Russula sp. uncultured Russula Iran Bahram et al. (2012) FR852097 – – Russula sp. Estonia B181 Bahram et al. (2011) FN669244 FN669244 – Russula sp. China EMF33 Ding et al. (unpubl.) JF273535 – – Russula sp. China ZWG2011 Ge et al. (2012) JN129407 – – Russula sp. China CFXY10066 Tian (unpubl.), Li et al. (2018) KR082870 – – Russula sp. Pakistan MAM0077 (LAH) Adamčik et al. (2016) KU886598 – – Russula sp. USA r-03045 Davis (unpubl.) JF834346 JF834493 JF834442 Table 1 (cont.) Species (*) sequestrate1 Original identification (*) type Origin Voucher Id (Herbarium)2 Reference GenBank accession codesITS rDNA 28S rDNA RPB2 Russula sp. * Gymnomyces sp. USA T13003 (OSC) Whitbeck (2003) AY239337 – – Gymnomyces sp. USA T13006 (OSC) Whitbeck (2003) AY239340 – – Gymnomyces sp. USA 1059 (SOC) Gladish et al. (2010) FJ789600 – – Gymnomyces sp. USA T35827 (OSC) Frank (unpubl.) KF386763 – – Gymnomyces alveolatus USA TK1719, SNF35 Izzo et al. (2005) AY558757 – – Martellia sp. USA TK165 Izzo et al. (2005) AY558788 – – Russula absphaerocellaris* Russula absphaerocellaris* China FAN492 (BJTC) Sang et al. (2016) KY270486 KY270493 – Russula absphaerocellaris China FAN448 (BJTC) Sang et al. (2016) KY270489 – – Russula aeruginea Russula aeruginea Sweden AT2003017 (UPS) Buyck et al. (2008) DQ421999 DQ421999 DQ421946 Russula albobrunnea Russula albobrunnea Australia TL2136 (MEL) Lebel & Tonkin (2007) EU019933 EU019933 – Russula alutacea Russula alutacea Italy 6472 (MCVE) Osmundson et al. (2013) JF908676 – – Russula amethystina Russula amethystina Germany hue215 (TUB) Eberhardt (2002) AF418640 AY606971 – Russula ammophila* Macowanites ammophilus Portugal 42956 (AH) This work MK105622 –– Gymnomyces ammophilus Portugal Fungi 40132 (MA) Calonge & Martín (2000) AF230890 – – Gymnomyces ammophilus Portugal Fungi 40137 (MA) Calonge & Martín (2000) AF230891 – – Macowanites ammophilus Portugal Fungi 51167 (MA) Vidal et al. (2002) AJ438036 – – Macowanites ammophilus Spain Fungi 51165 (MA) Vidal et al. (2002) AJ438038 – – Macowanites ammophilus Spain Fungi 51166 (MA) Vidal et al. (2002) AJ438037 – – Macowanites ammophilus Spain 43950 (AH) This work MK105623 –– Macowanites ammophilus Spain 46370 (AH) This work MK105624 MK105706 MK102748 Russula ammophila Spain IC09010703 (BCN) This work MK112566 MK108033 – Russula amoenicolor Russula amoenicolor Europe 311IX76, BB99.823 (PC) Miller & Buyck (2002) AY061655 – – Russula cf. amoenolens Russula cf. amoenolens USA CDW122 Avis (2012) JQ622327 – – Russula cf. amoenolens USA PRL4149 Avis (2012) JQ622333 – – Russula cf. amoenolens USA 12794 (MICH) Bazzicalupo et al. (2017) KF245512 – – Russula cf. amoenolens USA BPL232 (TENN) Looney et al. (2016) KT933954 KT933813 KT933884 Russula amoenolens Russula amoenolens France 12838 (MICH) Bazzicalupo & Berbee (unpubl.) KF245510 – – Russula amoenolens Spain 46371 (AH) This work MK105625 – MK102749 Russula amoenolens Spain 46372 (AH) This work MK105626 MK105707 MK102750 Russula amoenolens New Zealand 77763 (PDD) Johnston & Park (unpubl.) GU222264 – – Russula andaluciana* Gymnomyces sublevisporus* Spain BM360 (pers. herb.) This work MK105627 MK105708 MK102751 Gymnomyces sublevisporus Spain 39198 (AH) This work MK105628 –– Gymnomyces sublevisporus Spain 39239 (AH) This work MK105629 MK105709 – Russula aquosa Russula aquosa Estonia 101708 (TU) Bazzicalupo et al. (2017) UDB011290 KX812873 KX813654 Russula aromatica* Gymnomyces fragrans USA PNW 5607 (OSC) Whitbeck (2003) AY239331 – – Russula atrorubens Russula atrorubens Estonia 101718 (TU) Bazzicalupo et al. (2017) KX579812 – – Russula atroglauca Russula atroglauca Poland ID PAN 248 Trocha & Rudy (unpubl.) KM085418 – – Russula atrovirens Russula atrovirens New Zealand 77744 (PDD) Johnston & Park (unpubl.) GU222260 – – Russula aurantioflammans Russula aurantioflammans Europe r3245 Adamčik et al. (2016) KU928167 – – Russula aurea Russula aurea Estonia 101733 (TU) Bazzicalupo et al. (2017) UDB011363 KX812878 KX813659 Russula azurea Russula azurea Italy BB537/08.668 (PC) Schoch et al. (2012) JN944002 JN940591 JN993614 Russula aff. betularum Russula aff. betularum USA r-09003 Davis (unpubl.) JF834375 JF834520 JF834469 Russula brevipileocystidiata* Russula brevipileocystidiata* China FAN455 (BJTC) Sang et al. (2016) KY270487 KY270492 – Russula brunneola Russula brunneola USA r-03034 Davis (unpubl.) JF834341 JF834489 JF834438 Russula brunneonigra* Russula brunneonigra* Australia H5813 (DAR) Lebel & Tonkin (2007) EU019945 – – Russula caerulea Russula caerulea Germany hue146, 504IS77 (TUB) Eberhardt (2002) AF418633 AF325297 – Russula californica* Martellia californica USA T16027 (OSC) Whitbeck (2003) AY239308 – – Russula candida* Elasmomyces mattiroloanus Germany GG171 (M), JMV799894 (BCN) This work MK105630 –– Russula candida Spain JC171001N (pers. herb.) This work MK105631 MK105710 MK102752 Russula candida Spain JMV20100724b (BCN) This work MK105632 MK105711 – Russula candidissima* Macowanites candidus Italy ELG930704-1, JMV800185 (BCN) This work MK105633 MK105712 – Russula candidissima Poland F-2009-57 (KRA) This work MK105634 –– J.M. Vidal et al.: Revision of sequestrate Russulaceae Table 1 (cont.) Species (*) sequestrate1 Original identification (*) type Origin Voucher Id (Herbarium)2 Reference GenBank accession codes ITS rDNA 28S rDNA RPB2 Russula candidissima* (cont.) Russula candidissima Poland F-2017-2 (KRA) This work MK105635 –– Russula candidissima* Spain JMV20110906-6a (BCN) This work MK105636 MK105713 MK102753 Russula aff. cerea* Russula cerea Spain JMV20000817-1 (BCN) This work MK105637 MK105714 – Russula cerea* Russula cerea Germany JMV800660 (BCN) This work MK105638 MK105715 – Russula cerea Poland F-2012-28 (KRA) This work MK105639 – MK102754 Gymnomyces xanthosporus Spain JMV961003-8 (BCN) This work MK105640 –– Russula cerea Spain JMV20160705 (BCN) This work MK105641 –– Russula cerea Spain JMV800656 (BCN) This work MK105642 –– Gymnomyces sp. U.K. LM1492 Suz et al. (2014) KM576488 – – Russula cerolens Russula cerolens China HE2720 Sun et al. (unpubl.) KC505578 – – Russula cerolens China C3 Chai (unpubl.) JX434671 – – Russula cessans Russula cessans Slovakia BB525/07.219 (PC) Schoch et al. (2012) JN944011 JN940595 JN993601 Russula cheelii Russula cheelii Australia JET1002 (MEL) Lebel et al. (2013) JX266623 JX266638 – Russula chiui Russula chiui* China 250410 (HMAS) Li et al. (2015) KF225491 – – Russula chlorineolens* Macowanites chlorinosmus USA 36339 (OSC) Gordon (unpubl.) KT968579 KT968633 – Macowanites chlorinosmus USA 51028 (OSC) Gordon (unpubl.) KT968578 KT968632 – Russula clariana Russula clariana Europe 492RUS26, BB2733 (PC) Miller & Buyck (2002) AY061664 – – Russula claroflava Russula claroflava Germany FH12-212 (GENT) Looney et al. (2016) KT933997 KT933858 KT933929 Russula consobrina Russula consobrina Estonia 101714 (TU) Bazzicalupo et al. (2017) UDB011295 KX812874 KX813655 Russula corallina Russula corallina USA BB229/06.324 (PC) Schoch et al. (2012) JN944006 JN940605 JN993609 Russula cremeolilacina Russula cremeolilacina Guadeloupe (FR) BA02.11.04 Sene et al. (2015) FR682087 – – Russula crustosa Russula crustosa USA BPL251 (TENN) Looney et al. (2016) KT933963 KT933822 KT933894 Russula cuprea Russula cuprea Germany FH12-250 (GENT) Looney et al. (2016) KT934010 KT933871 KT933942 Russula curtipes Russula curtipes Germany FH12-206 (GENT) Looney et al. (2016) KT933995 KT933856 KT933927 Russula cf. cyanoxantha Russula cf. cyanoxantha USA BPL280 (TENN) Looney et al. (2016) KT933976 KT933837 KT933908 Russula cyanoxantha Russula cyanoxantha Germany FH12-201 (GENT) De Crop et al. (2017) KR364093 KR364225 KR364341 Russula decolorans Russula decolorans Germany FH12-196 (GENT) Looney et al. (2016) KT933992 KT933853 KT933924 Russula dryadicola Russula dryadicola France 16243 (MCVE) Osmundson et al. (2013) JF908710 – – Russula dryadicola Finland 151632 (TURA) Adamčik et al. (2016) KU928146 – KY616724 Russula ellipsospora* Gymnomyces ellipsosporus USA T61608 (OSC) Whitbeck (2003) AY239304 – – Gymnomyces ellipsosporus USA 58973 (OSC) Whitbeck (2003) AY239306 – – Russula emetica Russula emetica Sweden UE05.10.2003-11 (UPS) Buyck et al. (2008) DQ421997 DQ421997 DQ421943 Russula exalbicans Russula exalbicans Germany n179/93 (TUB) Eberhardt (2002) AF418622 AF325306 – Russula faustiana Russula faustiana China XJ2013100408 Qiao et al. (unpubl.) KX655858 – – Russula fellea Russula fellea Germany FH12-185 (GENT) Looney et al. (2016) KT933989 KT933850 KT933921 Russula firmula Russula firmula Sweden AT2004142 (UPS) Buyck et al. (2008) DQ422017 DQ422017 DQ421958 Russula cf. flavisiccans Russula cf. flavisiccans USA BB2004-254 (PC) Kong et al. (2015) EU598162 EU598162 – Russula foetens Russula foetens Germany hue124 (TUB) Eberhardt (2002) AF418613 AF325299 – Russula fontqueri Russula fontqueri Germany FH12-223 (GENT) Looney et al. (2016) KT934003 KT933864 KT933935 Russula galbana* Russula galbana Australia H4667 (MEL) Lebel & Tonkin (2007) EU019936 EU019936 – Russula galileensis* Macowanites galileensis Israel G-83 (HAI) This work MK105643 MK105716 – Macowanites galileensis Israel G-166 (HAI) This work MK105644 MK105717 MK102755 Macowanites galileensis Israel G-201 (HAI) This work MK105645 –– Macowanites galileensis Israel G-202 (HAI) This work MK105646 –– Macowanites galileensis Israel G-209 (HAI) This work MK105647 –– Russula gamundiae* Cystangium sp. Chile T26316 (OSC) Trierveiler-Pereira et al. (2015) KF819810 – – Russula gilkeyae* Gymnomyces monosporus USA 117360 (OSC) Gordon & Zych (unpubl.), Trendel et al. (2017) EU669222 EU669273 – Gymnomyces gilkeyae USA T2572 (OSC) Whitbeck (2003) AY239346 – – Russula globispora Russula globispora Germany FH2007-BT111 (GENT) Adamčik et al. (2016) KU928144 – KY616671 Russula globispora Germany FH2007-BT121 (GENT) Adamčik et al. (2016) KU886594 – – Russula gracillima Russula gracillima Sweden UE23.08.2004-14 (PC) Buyck et al. (2008) DQ422004 DQ422004 DQ421949 Russula granulata Russula granulata USA BPL272 (TENN) Looney et al. (2016) KT933971 KT933832 KT933903 Table 1 (cont.) Species (*) sequestrate1 Original identification (*) type Origin Voucher Id (Herbarium)2 Reference GenBank accession codesITS rDNA 28S rDNA RPB2 Russula grisea Russula grisea Sweden UE2005.08.16-01 (UPS) Buyck et al. (2008) DQ422030 DQ422030 DQ421968 Russula heterophylla Russula heterophylla Sweden UE20.08.2004-2 (UPS) Buyck et al. (2008) DQ422006 DQ422006 DQ421951 Russula hiemisilvae Russula hiemisilvae Cameroon 1203IS83, SLM (pers. herb.) Kleine et al. (2013) JQ902080 – – Russula hobartiae* Russula hobartiae Cyprus ML4193GY (pers. herb.) This work MK105648 MK105718 MK102756 Russula hobartiae Cyprus ML411161GY (pers. herb.) This work MK105649 –– Russula hobartiae * Cyprus ML110131GY, JMV800628 (BCN) This work MK105650 MK105719 – Russula hobartiae Cyprus GK5889, JMV800647 (BCN) This work MK105651 MK105720 MK102757 Russula ilicis Russula ilicis Europe 563IC52, MS10/18/99 (pers. herb.) Miller & Buyck (2002) AY061682 – – Russula illota Russula illota Sweden UE26.07.2002-3 (UPS) Buyck et al. (2008) DQ422024 DQ422024 DQ421967 Russula inamoena Russula inamoena Italy 107 Melera et al. (2017) KJ834597 – – Russula inflata Russula inflata Madagascar SLM00/344 Kleine et al. (2013) JQ902062 – – Russula insignis Russula insignis Europe 1223IS85, BB00.2149 (PC) Miller & Buyck (2002) AY061700 – – Russula integra Russula integra Germany FH12-172 (GENT) Looney et al. (2016) KT933984 KT933845 KT933916 Russula intervenosa Russula intervenosa* India AM273 (CUH), 1272 (CAL) Crous et al. (2016) KT824241 KU928135 – Russula kanadii Russula kanadii India AM086 (CUH) Dutta et al. (2015) KJ866936 – – Russula kermesina* Macowanites carmineus New Zealand 92049 (PDD) Johnston & Park (unpubl.) GU222323 – – Russula laricina Russula laricina Italy BB575/08.681 (PC) Schoch et al. (2012) JN944008 JN940593 KU237846 Russula laricinoaffinis Russula laricinoaffinis Italy 662 (MCVE) Osmundson et al. (2013) JF908646 – – Russula laurae* Cystangium sp. Chile T26311 (OSC) Trierveiler-Pereira et al. (2015) KF819811 – – Russula laurocerasi Russula laurocerasi Germany FH12-178 (GENT) Looney et al. (2016) KT933988 KT933849 KT933920 Russula lepida Russula lepida Belgium HJB9990 (UPS) Buyck et al. (2008) DQ422013 DQ422013 DQ421954 Russula lilacea Russula lilacea Slovakia BB435/07.213 (PC) Schoch et al. (2012) JN944005 JN940592 JN993610 Russula livescens Russula livescens China F0177 Xie et al. (2010) GU371295 – – Russula longisterigmata* Cystangium sp. Chile T26265 (OSC) Trierveiler-Pereira et al. (2015) KF819808 – – Russula maculata Russula maculata Germany FH2010-BT184 (GENT) Adamčik et al. (2016) KU928154 – – Russula mairei Russula mairei Germany FH12-262 (GENT) Looney et al. (2016) KT934013 KT933874 KT933945 Russula mariae Russula mariae Korea 19111 (HCCN) Park et al. (2013) KF361762 KF361812 KF361712 Russula mariae Korea 20120922-08 (SFC) Park et al. (2013) KF361778 KF361828 KF361728 Russula mattiroloana* Russula mattiroloana Greece GK3901, JMV800638 (BCN) This work MK105652 MK105721 – Russula mattiroloana Greece GK8136, JMV800644 (BCN) This work MK105653 MK105722 MK102758 Russula mattiroloana Poland F-2012-153 (KRA) This work MK105654 –– Russula mattiroloana Poland F-2017-1 (KRA) This work MK105655 –– Russula mattiroloana Poland F-2018-1 (KRA) This work MK105656 MK105723 – Russula mattiroloana Poland F-2018-2 (KRA) This work MK105657 MK105724 MK102759 Russula mattsmithii* Gymnomyces compactus USA T13171 (OSC) Whitbeck (2003) AY239303 – – Gymnomyces compactus USA T13565 (OSC) Whitbeck (2003) AY239342 – – Russula mediterraneensis* Russula mediterraneensis Greece GK3930, JMV800639 (BCN) This work MK105658 MK105725 – Russula mediterraneensis Greece GK7286, JMV800642 (BCN) This work MK105659 –– Russula mediterraneensis* Greece GK6710, JMV800641 (BCN) This work MK105660 MK105726 – Russula mediterraneensis Italy MG630, 29085 (MCVE) This work MK105661 – MK102760 Russula mediterraneensis Italy MG636, 29086 (MCVE) This work MK105662 – MK102761 Russula megapseudocystidiata Russula megapseudocystidiata* China FAN454 (BJTC) Sang et al. (2016) KY270488 KY270491 – Russula megaspora* Gymnomyces megasporus Australia T18717 (OSC) Miller et al. (2001) – AF265535 – Russula meridionalis* Zelleromyces meridionalis* Spain BM410 (pers. herb.) This work MK105663 –– Russula meridionalis Spain IC20051417 (BCN) This work MK105664 MK105727 – Russula meridionalis Spain IC24051506 (BCN) This work MK105665 MK105728 MK102762 Russula meridionalis Spain JC180617NR (pers. herb.) This work MK105667 MK105729 – Russula messapica Russula messapica Italy 562IC52, MS6/11/89 (pers. herb.) Miller & Buyck (2002) AY061692 – – var. messapica Russula messapica Italy 46373 (AH) This work MK105668 – MK102763 Russula messapica Spain JL201111182 (BCN) This work MK105669 MK105730 MK102764 Russula messapica Macowanites messapicoides Spain IC28110613 (BCN) This work MK105666 –– aff. var. messapicoides* J.M. Vidal et al.: Revision of sequestrate Russulaceae Table 1 (cont.) Species (*) sequestrate1 Original identification (*) type Origin Voucher Id (Herbarium)2 Reference GenBank accession codesITS rDNA 28S rDNA RPB2 Russula messapica Russula messapica var. messapicoides* var. messapicoides Greece VK2998 (pers. herb.) This work MK105670 MK105731 MK102765 Russula messapica var. messapicoides Greece VK3368 (pers. herb.) This work MK105671 MK105732 MK102766 Russula messapica var. messapicoides Greece VK3411, JMV800682 (BCN) This work MK105672 –– Russula messapica var. messapicoides Greece GK9341, JMV800645 (BCN) This work MK105673 –– Macowanites messapicoides* Spain JL1493 (BCN) This work MK105674 –– Russula mistiformis* Russula mistiformis Greece JMV800652 (BCN) This work MK105675 –– Gymnomyces mistiformis Italy AM1653 (pers. herb.) Whitbeck (2003) AY472079 – – AY472080 – – Russula mistiformis Italy JMV800125 (BCN) This work MK105676 –– Russula mistiformis Spain JC170305NR, JMV800661 (BCN) This work MK105677 –– Russula mistiformis Spain JC131117NR, JMV800662 (BCN) This work MK105678 MK105733 MK102767 Russula mistiformis Spain JC101206BT, JMV800663 (BCN) This work MK105679 –– Russula mistiformis Spain AMC H-69, JMV971127 (BCN) This work MK105680 –– Russula mistiformis Spain JC110102NR (pers. herb.) This work MK105681 –– Martellia pila Spain JMV920626-1 (BCN) This work MK105682 –– Russula mistiformis Spain JMV20140607-3 (BCN) This work MK105683 –– Russula mistiformis Spain JMV20160524-1 (BCN) This work MK105684 –– Martellia mediterranea* Spain GM-RG11057 (AH) This work MK105685 MK105734 – Russula monospora* Russula monospora Spain 46459 (AH) This work MK105686 MK105735 – Russula monospora Spain JMV800671 (BCN) This work MK105687 –– Russula monospora Spain JMV800672 (BCN) This work MK105688 –– Russula monospora Spain JMV800673 (BCN) This work MK105689 –– Russula monticola* Gymnomyces monticola USA 56167 (OSC) Whitbeck (2003) AY239313 – – Russula murrillii Russula murrillii USA r-04075 Davis (unpubl.) JF834352 JF834500 JF834449 Russula mustelina Russula mustelina Germany FH-12-226 (GENT) Looney et al. (2016) KT934005 KT933866 KT933937 Russula nana Russula nana 101701 (TU) Wood et al. (unpubl.) KX579809 – – Russula nauseosa Russula nauseosa Germany FH12-173 (GENT) Looney et al. (2016) KT933985 KT933846 KT933917 Russula neerimea Russula neerimea Australia 2101871 (MEL) Lebel & Tonkin (2007) EU019915 EU019915 – Russula nitida Russula nitida USA BB2004-272 (PC) Hughes & Buyck (unpubl.) EU598164 EU598164 – Russula nitida Germany FH12-218 (GENT) Looney et al. (2016) KT934001 KT933862 KT933933 Russula nondistincta* Gymnomyces nondistincta USA 62139 (OSC) Gordon (unpubl.) KP859276 KP859294 – Russula nothofagi* Cystangium sp. Chile T26350 (OSC) Trierveiler-Pereira et al. (2015) KF819809 – – Russula nympharum Russula nympharum France HM-R-9702 (M) Adamčik et al. (2016) KU928159 – – Russula ochroleuca Russula ochroleuca Germany FH12-211 (GENT) Looney et al. (2016) KT933996 KT933857 KT933928 Russula cf. ochrophylla Russula cf. ochrophylla USA BPL231 (TENN) Looney et al. (2016) KT933953 KT933812 KT933883 Russula ochrospora Russula ochrospora Italy GD20.07.2004 (UPS) Buyck et al. (2008) DQ422012 DQ422012 DQ421953 Russula odorata Russula odorata Slovakia BB526/07.186 (PC) Schoch et al. (2012) JN944010 JN940607 JN993603 Russula osphranticarpa* Gymnomyces redolens New Zealand SLM41I62 (pers. herb.) Smith et al. (2006) DQ403803 AF265536 – Russula paludosa Russula paludosa Germany FH12-216 (GENT) Looney et al. (2016) KT934000 KT933861 KT933932 Russula parazurea Russula parazurea Sweden BW06.09.2002-16/MF01.10.2003 (UPS) Buyck et al. (2008) DQ422007 DQ422007 DQ421945 Russula parksii* Gymnomyces parksii USA T14997 (OSC) Whitbeck (2003) AY239335 – – Russula pauriensis Russula pauriensis India 1395 (CAL) Das et al. (2017) MF535185 – – Russula peckii Russula peckii USA BPL270 (TENN) Looney et al. (2016) KT933970 KT933830 KT933901 Russula pectinatoides Russula pectinatoides USA BPL276 (TENN) Looney et al. (2016) KT933975 KT933836 KT933907 Russula persicina Russula persicina Sweden UE21.09.2003-01 (UPS) Buyck et al. (2008) DQ422019 DQ422019 DQ421960 Russula pila* Russula pila Spain JMV970816-8 (BCN) This work MK105690 –– Russula pila Spain JMV800654 (BCN) This work MK105691 MK105736 MK102768 Russula aff. pilosella* Russula aff. pilosella Australia H4784 (MEL) Lebel & Tonkin (2007) EU019932 EU019932 – Russula pilosella* Russula pilosella Australia H5974 (BRI) Lebel & Tonkin (2007) EU019941 – – Table 1 (cont.) Species (*) sequestrate1 Original identification (*) type Origin Voucher Id (Herbarium)2 Reference GenBank accession codesITS rDNA 28S rDNA RPB2 Russula praetervisa Russula praetervisa* Italy 1997/0812 (IB) Sarnari & Eberhardt (unpubl.) UDB019331 – – Russula pseudocarmesina Russula pseudocarmesina Burundi BB5401 (PC) Kleine et al. (2013) JQ902068 – – Russula pseudopectinatoides Russula pseudopectinatoides* China 251523 (HMAS) Li et al. (2015) KM269077 – – Russula puellaris Russula puellaris Estonia 101839 (TU) Bazzicalupo et al. (2017) UDB015995 KX812887 KX813668 Russula puellula Russula puellula Europe 2-1130IS76, RW27065 (E) Miller & Buyck (2002) AY061710 – – Russula puellula Slovakia F-3107 (SAV) Cabon et al. (unpubl.) KY582704 – KY616697 Russula puiggarii Russula puiggarii French Guiana G3130 Roy et al. (2016) KJ786689 KJ786592 – Russula pulverulenta Russula pulverulenta USA 4-1144IS79, RF566 (pers. herb.) Miller & Buyck (2002) AY061736 – – Russula punicea Russula punicea USA r-03043 Davis (unpubl.) JF834344 JF834492 JF834441 Russula purpureoflava Russula purpureoflava Australia JET1128 (MEL) Lebel et al. (2013) JX266626 JX266641 – Russula putida Russula putida* Italy 1997/0791 (IB) Borovicka (unpubl.) HG798527 HG798526 – Russula risigallina Russula risigallina Sweden UE03.07.2003-08 (UPS) Buyck et al. (2008) DQ422022 DQ422022 DQ421961 Russula romellii Russula romellii Germany FH12-177 (GENT) Looney et al. (2016) KT933987 KT933848 KT933919 Russula rosea Russula rosea France BB430/07.780 (PC) Schoch et al. (2012) JN944003 JN940602 JN993613 Russula roseipes Russula roseipes Estonia 101806 (TU) Bazzicalupo et al. (2017) UDB015972 KX812884 KX813665 Russula rostraticystidia* Russula rostraticystidia* Australia H6165 (BRI) Lebel & Tonkin (2007) EU019938 – – Russula rubellipes Russula rubellipes USA BPL240 (TENN) Looney et al. (2016) KT933958 KT933817 KT933888 Russula rubrolutea* Macowanites rubroluteus* Australia T12610 (PDD, OSC) Lebel & Tonkin (2007) EU019940 EU019940 – Russula rubropunctatissima Russula rubropunctatissima Brazil JDuque55 Duque et al. (unpubl.) KY087554 KY087583 – Russula rugulosa Russula rugulosa USA BPL654 (TENN) Looney et al. (2016) KY848516 – KY701373 Russula sanguinea Russula sanguinea Germany FH12-240 (GENT) Looney et al. (2016) KT934008 KT933869 KT933940 Russula sarnarii Russula sarnarii* India 1395 (CAL) Ghosh et al. (2017) KY284154 – – Russula seminuda* Cystangium seminudum Australia H5346 (MEL, OSC) Lebel & Tonkin (2007) EU019947 – – Russula sessilis* Cystangium sessile Australia H5038 (OSC, MEL) Miller et al. (2001), Lebel & Tonkin (2007) EU019948 AF265533 – Russula setigera* Gymnomyces setigerus USA 29622 (OSC) Whitbeck (2003) AY239317 – – Russula shafferi* Gymnomyces brunnescens USA 51046 (OSC) Whitbeck (2003) AY239327 – – Russula shingbaensis Russula shingbaensis* India KD11-094 (BSHC) Das et al. (2014) KM386692 – – Russula sichuanensis* Russula sichuanensis Pakistan MSM0025 (LAH) Saba & Khalid (2015) KM596865 – – Russula similaris* Gymnomyces fallax Mexico GO2009-239 (MEXU) Kong et al. (2015) KC152107 – – Russula cf. sororia Russula sp. Korea ASIS22640 (HCCN) Lee et al. (2017) KX574703 – – Russula sororia Korea 2208ASI (HCCN) Lee et al. (2017) KX574701 – – Russula stewartii* Gymnomyces monosporus USA 61382 (OSC) Gordon (unpubl.) KP859259 KP859279 – Gymnomyces monosporus USA 61637 (OSC) Gordon (unpubl.) KP859261 KT968617 – Russula cf. subfoetens Russula cf. subfoetens Spain JMV941022-7 (BCN) This work MK105692 MK105737 – Russula subfulva* Gymnomyces subfulvus USA T14998 (OSC) Whitbeck (2003) AY239321 – – Russula subloculata* Macowanites americanus USA SLM480MC (pers. herb.) Miller et al. (2001) AF265540 – – Russula subochracea* Gymnomyces subochraceus USA T13386 (OSC) Whitbeck (2003) AY239325 – – Russula suecica Russula suecica Norway JR7994 (KUO) Vauras et al. (2016) KX988014 – – Russula tapawera* Macowanites tapawera* New Zealand T12607 (OSC), 83696 (PDD) Lebel & Tonkin (2007) EU019935 EU019935 – Russula tawai Russula tawai New Zealand 77760 (PDD) Johnston & Park (unpubl.) GU222263 – – Russula theodoroui* Cystangium theodoroui Australia SLM43I84 (pers. herb.) Smith et al. (2006) DQ403804 – – Russula tricholomopsis Russula tricholomopsis New Zealand 77749 (PDD) Johnston & Park (unpubl.) GU222261 – – Russula turci Russula turci Finland 101874 (TU) Bazzicalupo et al. (2017) UDB016082 KX812891 KX813672 Russula unicalifornica* Gymnomyces cinnamomeus USA T19930 (OSC) Whitbeck (2003) AY239338 – – Russula variispora* Russula variispora* Australia H5855 (DAR) Lebel & Tonkin (2007) EU019934 EU019934 – Russula velenovskyi Russula velenovskyi Slovakia F-2921 (SAV) Cabon et al. (unpubl.) KY582701 – KY616695 Russula venezueliana Russula venezueliana French Guiana TH7874 (pers. herb.) Smith et al. (2017) KT339269 – – Russula versicolor Russula versicolor USA r-04039 Davis (unpubl.) JF834350 JF834498 JF834447 Russula versicolor Slovakia BB589/07.288 (PC) Schoch et al. (2012) JN944009 JN940594 KU237859 Russula vesca Russula vesca Sweden AT2002091 (UPS) Buyck et al. (2008) DQ422018 DQ422018 DQ421959 Russula vidalii* Gymnomyces ilicis Spain JC100508BT01, JMV800688 (BCN) This work MK105693 MK105738 – Russula vidalii Spain JMV20160517-1 (BCN) This work MK105694 MK105739 MK102769 Russula vinaceodora* Macowanites vinaceodorus Spain Fungi 46524 (MA) Vidal et al. (2002) AJ438035 – – J.M. Vidal et al.: Revision of sequestrate Russulaceae Table 1 (cont.) Species (*) sequestrate1 Original identification (*) type Origin Voucher Id (Herbarium)2 Reference GenBank accession codesITS rDNA 28S rDNA RPB2 Russula vinaceodora* (cont.) Macowanites vinaceodorus* Spain Fungi 47416 (MA) Vidal et al. (2002) AJ438034 – – Macowanites vinaceodorus Spain 46374 (AH) This work MK105695 MK105740 MK102770 Russula vinosa Russula vinosa Sweden F124791 (UPS) Looney et al. (2016) KX812857 KX812900 KX813680 Russula violeipes Russula violeipes Korea 20121010-06 (SFC) Park et al. (2013) KF361808 KF361858 KF361758 Russula violeipes Korea 16735 (HCCN) Park et al. (2013) KF361783 KF361833 KF361733 Russula virescens Russula virescens Belgium HJB9989 (UPS) Buyck et al. (2008) DQ422014 DQ422014 DQ421955 Russula viscida Russula viscida Canada F 16576 (UBC) Berbee et al. (unpubl.) FJ627039 – – Russula werneri Russula werneri* Italy 1997/0786 (IB) Eberhardt (unpubl.) DQ422021 DQ422021 – Russula wollumbina Russula wollumbina Australia 2238232 (MEL) Lebel & Tonkin (2007) EU019921 EU019921 – Russula xerampelina Russula xerampelina Canada OUC97303 (DAVFP) Durrall et al. (2006) DQ367916 DQ367916 – Russula xerophila* Gymnomyces xerophilus* USA 82218, SRC672 (OSC) Smith et al. (2006) DQ028473 – – Gymnomyces xerophilus USA 82219, SRC648 (OSC) Smith et al. (2006) DQ028476 – – uncultured UDB027052 – – /stereaceae/stereum Stereum hirsutum Stereum hirsutum FPL 8805 Binder & Hibbett (2002) AY854063 AF393078 AY218520 1 New species, new combinations and sequences produced from this study are in bold. Sequestrate taxa are marked with an asterisk (*) symbol. 2 Collection sources: AB = Amadou Bâ (GENT); ADK = André De Kesel (BR); AM = Amer Montecchi (pers. herb.); AT = Andrew Taylor (UPS); AV = Annemieke Verbeken (GENT); AVoitk = Andrus Voitk (GENT); BB = Bart Buyck (PC); BH = Bryony M. Horton (HO); BM = Baldomero Moreno-Arroyo(pers. herb.); BPL = Brian P. Looney (TENN); BW = Birgitta Wasstorp (UPS); CLC = Cathy L. Cripps (MONT); CWD = Christopher W. Dunk (MEL); DA = David Arora (GENT); DED = Dennis E. Desjardin (PC, SFSU); DS = Dirk Stubbe (GENT); EB = Edward G. Barge (MONT); EG = Edith Garay-Serrano (XAL); EH = Egon Horak (ZT); EU = Elisabeth Uhlmann (UPS); FA = François Ayer (WSL); FAN = Li Fan (BJTC); FH = Felix Hampe (GENT); FO = Franz Oberwinkler (TUB); GD = Giuseppe Donelli (UPS); GG = Gerhardt Gross (M); GK = Georges Konstantinidis (pers. herb.); GM-RG = Gabriel Moreno and Ricardo Galán (AH); GO = Roberto Garibay Orijel (MEXU); HJB = Henri J. Beker (UPS); HM-R = Hermann Marxmler (M); HRL = Renée Lebeuf (pers.herb.); HTL = Huyen Than Le (CMU, GENT); IC = Aurelia Paz Conde (BCN); JC = Julio Cabero (pers. herb.); JET = Jennifer E. Tonkin (MEL); JL = Jaume Llistosella (BCN); JMV = Josep M. Vidal (BCN); JN = Jorinde Nuytinck (GENT); JR = Juhani Ruotsalainen (KUO); JV (TURA) = Jukka Vauras (TURA); JV (C, GENT) = Jan Verterholt (C, GENT); KD = Kanad Das (BSD); KVP = Kobeke Van de Putte (GENT); LM = Leticia Montoya (XAL); MAM = Malka Saba (LAH); MCA = M. Catherine Aime (pers. herb.); MF = Marco Floriani (UPS); ML = Michael Loizides (pers. herb.); MS = Mauro Sarnari (pers. herb.); MTB = Maria Teresa Basso (GENT); PAM = Pierre-Arthur Moreau (GENT, PC); PBM = P. Brandon Matheny (CUW, WTU); PL = Patrick Leonard (pers. herb.); RF = Raymond M. Fatto (pers. herb.); RH = Roy E. Halling (NY); RW (GENT) = Ruben Walleyn (GENT); RW (E) = Roy Watling (E); SLM = Steven L. Miller (pers. herb.); T = James M. Trappe (MEL, OSC, PDD); TH = Terry W. Henkel (pers. herb.); TL = Teresa Lebel (MEL); UE = Ursula Eberhardt (TUB, UPS); VK = Vasileios Kaounas (pers. herb.); WCR = William C. Roody (PC); ZWG = Zai-Wei Ge. were above 0.95, and subsignificant if BP values were above 60 % or PP values were above 0.90. RESULTS Maximum-likelihood analysis of the general ITS – 28S rDNA – rpb2 dataset including all new sequences and those recovered from public databases (Fig. 1) produced a phylogenetic tree consisting of four significantly supported clades, those of genera Lactarius, Lactifluus, Multifurca and Russula, in accordance with previous reconstructions of the Russulaceae lineage (Buyck et al. 2008, Lebel et al. 2013, Verbeken et al. 2014b, Kong et al. 2015, Looney et al. 2016, Song et al. 2016, De Crop et al. 2017). Lactarius (Fig. 2) was composed of five main clades, namely Lactarius subg. Plinthogalus, /L. hispidulus, /L. bisporus, /L. ka-bansus, and a lineage including Lactarius subg. Lactarius (= Lactarius subg. Piperites) and Lactarius subg. Russularia, all of them receiving significantorsubsignificantsupport,in agreementwith previous analyses of Lactarius (Eberhardt & Verbeken 2004, Buyck et al. 2008, Verbeken et al. 2014a, b, Barge et al. 2016), except for subg. Lactarius (PP 0.88, BP 58). All European taxa of sequestrate Lactarius nested within the monophyletic clade formed by L. subg. Lactarius and L. subg. Russularia. Seven significantly supported sequestrate species were identified, two of which belong to L. subg. Russularia and five fall outside and are therefore considered part of L. subg. Lactarius. Lactarius subg. Russularia received subsignificant support (PP 0.94, BP 47), AFTOL-ID 492 Stereum hirsutum AFTOL-ID 1897 Auriscalpium vulgare 100 olrim409 Amylostereum laevigatum AFTOL-ID 455 Echinodontium tinctorium 91 87 LACTARIUS 100 90 LACTIFLUUS 99 MULTIFURCA BPL242 Russula compacta PALLIDOSPORINAE, /R. cascadensis, /R. metachromatica 73 100 LACTARIOIDEAE 98 NIGRICANTINAE ARCHAEINAE 74 100 FARINIPEDES 100 82 RUSSULA 99 INGRATAE 99 RIGIDAE (2) RIGIDAE (1) (62) (68) RUSSULA S. STR. 98 0.05 Fig. 1 Consensus phylogram of the family Russulaceae obtained in RAxML after 2000 bootstrap iterations of a combined alignment of ITS rDNA, 28S rDNA and rpb2. Major clades were collapsed and rooting branch shortened for publication. Nodes were annotated if supported by > 70 % ML BP, but non- significant support values are exceptionally represented inside parentheses. J.M. Vidal et al.: Revision of sequestrate Russulaceae 0.05 Fig. 2 Extended phylogram of genus Lactarius obtained from Fig. 1. Rooting branch was shortened for publication. Nodes were annotated if supported by > 0.95 bayesian PPor > 70 % MLBP, but non-significant support values are exceptionally represented inside parentheses. Major lineages are highlighted with coloured boxes. Sequestrate species are marked with black circle (●) symbols. Specimens sequenced in the present work are highlighted in bold. BPL272 (TENN) Russula granulata /R. granulata 1.00/100 HE2720 Russula cerolens 1.00/100 0.99/90 C3 Russula cerolens 12794 (MICH) Russula cf. amoenolens 1.00/99 BPL232 (TENN) Russula cf. amoenolens PRL4149 Russula cf. amoenolens 1.00/100 CDW122 Russula cf. amoenolens 2208ASI (HCCN) Russula cf. sororia 1.00/97 ASIS22640 (HCCN) Russula cf. sororia H 5974 (BRI) Russula pilosella AUSTRALIA ● F0177 Russula livescens BB00.2149 (PC) Russula insignis 0.90/100 RF566 Russula pulverulenta 1.00/100 T13003 (OSC) Gymnomyces sp. USA ● T13006 (OSC) Gymnomyces sp. USA ● T19930 (OSC) Russula unicalifornica USA ● 0.99/89 TK165 Martellia sp. USA ● (0.92/63) TK1719 ‘Gymnomyces alveolatus’ USA ● BPL276 (TENN) Russula pectinatoides 1.00/100 1.00/90 1997/0812 (IB) Russula praetervisa Type 251523 (HMAS) Russula pseudopectinatoides Type hue124 (TUB) Russula foetens T13386 (OSC) Russula subochracea USA ● 0.93/90 H4784 (MEL) Russula aff. pilosella AUSTRALIA ● 0.99/97 2101871 (MEL) Russula neerimea AUSTRALIA ● H5813 (DAR) Russula brunneonigra Type AUSTRALIA ● (0.84/29) H4667 (MEL) Russula galbana AUSTRALIA ● FH12-178 (GENT) Russula laurocerasi UE26.07.2002-3 (UPS) Russula illota 107 Russula inamoena 1.00/100 1997/0791 (IB) Russula putida Type JMV941022-7 (BCN) Russula cf. subfoetens 51046 (OSC) Russula shafferi USA ● GO2009-239 (MEXU) Russula similaris MEXICO ● PNW 5607 (OSC) Russula aromatica USA ● 1.00/73 T14998 (OSC) Russula subfulva USA ● 1.00/93 T14997 (OSC) Russula parksii USA ● 62139 (OSC) Russula nondistincta USA ● JMV800660 (BCN) Russula cerea GERMANY ● F-2012-28 (KRA) Russula cerea POLAND ● 0.92/85 LM1492 Russula cerea U.K. ● JMV800656 (BCN) Russula cerea SPAIN ● (0.88/69) JMV20160705 (BCN) Russula cerea SPAIN ● 0.95/66 0.98/73 1.00/74 1.00/100 IC09010703 (BCN) Russula ammophila 43950 (AH) Russula ammophila SPAIN ● Fungi 51165 (MA) Russula ammophila SPAIN ● 46371 (AH) Russula amoenolens SPAIN 46372 (AH) Russula amoenolens SPAIN 12838 (MICH) Russula amoenolens FRANCE Fungi 51167 (MA) Russula ammophila PORTUGAL ● 46370 (AH) Russula ammophila SPAIN ● 77763 (PDD) Russula amoenolens NEW ZEALAND Fungi 51166 (MA) Russula ammophila SPAIN ● Fungi 40137 (MA) Russula ammophila PORTUGAL ● Fungi 40132 (MA) Russula ammophila PORTUGAL ● 42956 (AH) Russula ammophila PORTUGAL ● JMV800661 (BCN) Russula mistiformis SPAIN ● 0.05 JMV961003-8 (BCN) Russula cerea SPAIN ● JMV20000817-1 (BCN) Russula aff. cerea SPAIN ● JMV970816-8 (BCN) Russula pila SPAIN ● JMV800654 (BCN) Russula pila SPAIN ● JMV800662 (BCN) Russula mistiformis SPAIN ● AM1653 Russula mistiformis ITALY ● JMV800125 (BCN) Russula mistiformis ITALY ● JMV800663 (BCN) Russula mistiformis SPAIN ● JMV800652 (BCN) Russula mistiformis GREECE ● JMV20140607-3 (BCN) Russula mistiformis SPAIN ● JC110102NR Russula mistiformis SPAIN ● JMV20160524-1 (BCN) Russula mistiformis SPAIN ● GM-RG11057 (AH) Martellia mediterranea Type SPAIN ● JMV920626-1 (BCN) Russula mistiformis SPAIN ● JMV971127 (BCN) Russula mistiformis SPAIN ● /R. amoenolens /R. pilosella Subvelatae Pectinatinae Foetentinae Fig. 3 Extended phylogram of Russula sect. Ingratae obtained from Fig. 1. Rooting branch was shortened for publication. Nodes were annotated if supported by > 0.95 bayesian PPor > 70 % MLBP, but non-significant support values are exceptionally represented inside parentheses. Major lineages are highlighted with coloured boxes. Sequestrate species are marked with black circle (●) symbols. Specimens sequenced in the present work are highlighted in bold. J.M. Vidal et al.: Revision of sequestrate Russulaceae 1.00/99 INGRATAE JET1002 (MEL) Russula cheelii /R. cheelii KD11-094 (BSHC) Russula shingbaensis Type Pluviales Cyanoxanthinae Ilicinae BA 02.11.04 Russula cremeolilacina Griseinae (1) Griseinae (2) XJ20131004 Russula faustiana BPL251 (TENN) Russula crustosa FH-12-226 (GENT) Russula mustelina Virescentinae HJB9989 (UPS) Russula virescens AM086 (CUH) Russula kanadii BB5401 Russula pseudocarmesina 1203IS83 Russula hiemisilvae Pseudoepitheliosinae M00/344 Russula inflata G3130 Russula puiggarii r-03034 Russula brunneola AT2002091 (UPS) Russula vesca Heterophyllae UE20.08.2004-2 (UPS) Russula heterophylla 1395 (CAL) Russula pauriensis 39198 (AH) Russula andaluciana SPAIN ● BM360 Gymnomyces sublevisporus Type SPAIN ● 39239 (AH) Russula andaluciana SPAIN ● 1.00/99 1.00/98 16735 (HCCN) Russula violeipes 20121010-06 (SFC) Russula violeipes H6165 (BRI) Russula rostraticystidia Type AUSTRALIA ● AM273 (CUH) Russula intervenosa Type 0.99/93 H5855 (DAR) Russula variispora Type AUSTRALIA ● BB99.823 (PC) Russula amoenicolor 19111 (HCCN) Russula mariae 20120922-08 (SFC) Russula mariae 1.00/100 1.00/96 Amoeninae BB2004-272 (PC) Russula nitida 0.05 Fig. 4 Extended phylogram of Russula sect. Rigidae obtained from Fig. 1. Rooting branch was shortened for publication. Nodes were annotated if supported by > 0.95 bayesian PPor > 70 % MLBP, but non-significant support values are exceptionally represented inside parentheses. Major lineages are highlighted with coloured boxes. Sequestrate species are marked with black circle (●) symbols. Specimens sequenced in the present work are highlighted in bold. a result similar to that of Verbeken et al. (2014b), probably due Russula was divided in eight significantly supported monophy-to the insufficient phylogenetic signal in the species analyzed, letic clades (Fig. 1, 3–5): gaps in the phylogenetic diversity because of an incomplete sam-1) an isolated branch for /R. compacta; pling, or too much phylogenetic noise in the dataset employed. 2) a clade formed by Russula subsect. Pallidosporinae (Bon However, a significantly supported clade formed by L. borzianus, 1988), as well as the lineages of /R. cascadensis and L. josserandii (= Z. hispanicus) and some gymnocarpic species, /R. metachromatica; such as L. fulvissimus and L. subsericatus, in accordance with 3) Russula subsect. Lactarioideae (including R. delica); results of Verbeken et al. (2014b), Liu et al. (2015), or Barge et al. 4) Russula sect. Nigricantinae; (2016). The remaining European sequestrate Russulaceae 5) Russula sect. Archaeinae; species included in L. subg. Lactarius were identified as the 6) Russula subsect. Farinipedes (including R. crassotunicata existing species L. giennensis, L. stephensii and Z. soehneri, or and R. pallescens); otherwise accommodated into the new taxa L. populicola and 7) a monophyletic lineage formed by Russula sect. Ingratae L. subgiennensis. (≡ subg. Ingratula) and Russula sect. Rigidae (≡ subg. Heterophyllidia); and 8) a large Russula s.str. clade including the remaining spe-cies. Persoonia – Volume 42, 2019 Fig. 5 Extended phylogram of Russula s.str. obtained from Fig. 1. Rooting branch was shortened for publication. Nodes were annotated if supported by J.M. Vidal et al.: Revision of sequestrate Russulaceae These clades were also significantly supported by Kong et al. (2015), Looney et al. (2016) and Bazzicalupo et al. (2017). All European sequestrate Russulaceae taxa nested within the Ingratae-Rigidae clade or Russula s.str. Russula sect. Ingratae (Fig. 3) received significant support (PP 1.00, BP 99). It was composed of R. subsect. Subvelatae, Pectinatinae, and Foetentinae, as well as the lineages of /R. granulata, /R. amoenolens and /R. pilosella, all of them significantly supported except Foetentinae (PP 0.84, BP 29), probably because of incomplete data from many American sequestrate taxa. These results agree with those obtained by other researchers (Kong et al. 2015, Li et al. 2015, Lee et al. 2017, Melera et al. 2017). Sequestrate Russulaceae taxa are present in most of these lineages, but all currently known European sequestrate species were related to subsect. Foetentinae or the /R. amoenolens lineage. Russula ammophila was not significantly different from R. amoenolens employing ITS, 28S rDNA and rpb2 data, but significant differences were found in tef1 gene-data (data not shown). Russula sect. Rigidae (Fig. 4) was split into two clades. The first one was composed of R. subsect. Cyanoxanthinae and Pluviales (Singer 1986, Buyck & Ovrebo 2002, Cheype & Campo 2012), as well as the /R. cheelii lineage (PP 0.99, BP 55). The second clade was formed by the remaining species of Rigidae (PP 0.98, BP 62). This clade was composed of subsect. Ilicinae (BP 100, PP 1.00) on the one side, and the remaining subsections of Rigidae (PP 0.87, BP 77) on the other, including subsections Griseinae (divided in two clades, those of /R. ochrospora and /R. grisea), Virescentinae, Pseudoepitheliosinae, Heterophyllae and Amoeninae. Most of these clades were also significantly supported by the phylogenetic reconstructions of Rigidae conducted by Dutta et al. (2015) and Zhao et al. (2015). The only European sequestrate species related to sect. Rigidae was R. andaluciana, a replacement name for Gymnomyces sublevisporus, whose isotype was significantly related with subsect. Amoeninae. Russula s.str. (Fig. 5) was composed of a basal clade which did not receive significant support (PP 0.73, BP 50), including Russula subsect. Russula, as well as subsect. Sardoninae, subsect. Viscidinae, and the lineages of R. consobrina (subsect. Consobrinae) and R. fellea (subsect. Felleinae). A second large clade with significantsupportclustered all otherspecies ofRussula, and had itself two basal clades: subsections Lepidinae and Firmiores, the latter including gymnocarpic, pseudoangiocarpic and angiocarpic species. The remaining species of Russula grouped into several distinct clades with significant support: subsections Cupreinae, Laricinae, Lilaceinae, Maculatinae, Paludosinae, Puellarinae, Roseinae, a monophyletic clade con-taining subsections Amethystinae and Chamaeleontinae, and finally the so-called ‘crown’ clade (Looney et al. 2016), con-taining species from subsections Decolorantinae, Insidiosinae, Integrae, Olivaceinae, Vinosinae and Xerampelinae, as well as a lineage including the type species of genus Cystangium (C. sessile ≡ Russula sessilis), and many sequestrate and gymnocarpic species from Australia (Lebel & Tonkin 2007) and South America (Trierveiler-Pereira et al. 2015). Most European sequestrate taxa were related to subsections Firmiores, Laricinae, Maculatinae and Puellarinae, but there were also two independent lineages: one of them, R. monospora, was subsignificantly related with R. consobrina, the type species of subsect. Consobrinae (PP 0.94, BP 53), while the other, R. hobartiae, represented a new species from Cyprus signifi cantly related with a putative specimen of Russula ochrophylla. Lineages matching known species of sequestrate Russulaceae are listed below or re-described, and new names are proposed with complete descriptions for those representing new taxa. TAXONOMY Lactarius borzianus (Cavara) Verbeken & Nuytinck, Belg. J. Bot. 136, 2: 151. 2003 ― Fig. 6 Basionym. Arcangeliella borziana Cavara, Nuovo Giorn. Bot. Ital., Nuov. Ser. 7, 2: 126. 1900. Synonyms. Octaviania borziana (Cavara) Svrček in Pilát, Flora ČSR B1, Gasteromycetes: 740. 1958. Arcangeliella stephensii var. borziana (Cavara) Krieglst., Z. Mykol. 57, 1: 19. 1991. Arcangeliella volemoides K. Mader & A. Mader, Österr. Z. Pilzk. 1: 5. 1992. Basidiomata 1–4 cm wide, angiocarpic, subglobose to tuberi-form, often bilobed, sessile or subsessile, with a residual stipe; radially alveolate in the base. Pileus dry, first smooth, then areo- Fig. 6 Lactarius borzianus. a. Reproduction of Cavara’s plate of Arcangeliella borziana (Cavara 1900, tav. VII). ― b. MSL1962F0412. Basidiomata. ― c–e. AM1973. c. Spores in Melzer; d–e. SEM images of spores. ― Scale bars: b = 1 cm, c = 10 µm; d–e = 5 µm. ― Photos: b. M. Slavova; c. J.M. Vidal; d–e. UdG. late, basally closed, or open exposing the hymenophore; initially yellowish white (2A2) to pale yellow (3A3–4A5), then brownish orange (7C7) with reddish brown to dark brown (8E8–F8) maculae. Hymenophore loculate, labyrinthoid, yellowish white (2A2), pale yellow (4A4) to pale brown (7D7); locules irregularly arranged, 0.6–1.8 × 0.15–3 mm (1–2 per mm), elongated, sinuous, empty; fresh spore mass in locules yellowish white (4A1–A2). Columella percurrent or branched, apparent only in young specimens. Latex scant, more abundant in columella and pileal context, colourless to white, changing to yellow. Odour of wild bug (Nazara viridula) or musky; taste unpleasant, bitterish after a while. Spores 9.5–13(–15) × 8–10(–11) µm, Q = 1.16–1.28, hetero-tropic, subglobose to broadly ellipsoid, echinate; warts 1–1.5 µm high, isolated, amyloid; suprahilar plage absent; hilar appendix 1–1.5µm long, sometimes withashort sterigmalfragment 1–3 µm long. Basidia 3–4-spored, 30–65 × 10–15 µm, cylindri-cal to subclavate; sterigmata conical, curved, 5–8 µm long. Macrocystidia absent. Cystidioles infrequent, 20–50 × 4–6 µm, septate, cylindrico-rostrate to ventricose-rostrate, acuminate, more abundant in sterile locules. Paraphysoid cells abundant, 15–40 × 5–12 µm, cylindrical to slightly clavate, sometimes subglobose, often 1–2-septate. Pseudocystidia not observed. Subhymenium cellular,formed byprismaticcells8–25 µmdiam. Hymenophoral trama homiomerous, formed by septate hyphae 2–6.5 µm diam, with subglobose thickenings up to 12 µm wide in tramal anastomoses. Pileipellis and context 100–250 µm thick; suprapellis trichodermal to oedotrichodermal, formed by cylindrical, sinuous, septate hairs 10–40 × 3–6 µm, and lanceo-late dermatocystidia similar to those of the sterile locules, that soon collapses in a brown mass; subpellis made of prismatic hyphae up to 15 µm diam. Pileal context a cutis with the same composition of the hymenophoral trama, plenty of laticifera. Laticifera 2.5–10 µm diam, scarce in the hymenophoral trama and more abundant in the pileal context. Thromboplera 2–8 µm diam, abundant in young specimens. Some specimens present isolated sphaerocytes in the hymenophoral trama and pileal context, 15–30 µm diam, being more abundant and grouped in rosettes or chains in the columella. Habitat, Distribution & Season ― Gregarious, hypogeous or semi-hypogeous under needles, in subalpine conifer forests of Abies and Picea, on siliceous soil. Summer and autumn. So far known from the Alps, Apennine (Italy), and Rhodope mountains (Greece), at 1000–1800 m altitude. Material studied. AUSTRIA, Lower Austria, Gfl, near Dorf Brunn, under Picea abies, 12 Sept. 1970, A. & K. Mader (WU 10875, holotype of A. volemoides). – BULGARIA, Smolyan, Pamporovo, Southern Rhodope mountains, 1650–1750 m, under Picea abies, on siliceous soil, 15 Aug. 2017, M. Slavova (MSL1962F0412, duplicate BCN JMV800685). – FRANCE, Rhe-Alpes, Savoie, Héry-sur-Ugine, Réserve Naturelle du Nant Pareu-Merdassier, 1450 m, under Picea abies in the vicinity of Lactarius aurantiofulvus, on siliceous soil, 13 Aug. 2002, P.-A. Moreau as A. borziana (PAM02081306, duplicate BCN JMV800279b)*. – ITALY, Emilia-Romagna, Reggio Emilia, Civago, Abetina Reale forest, Tuscan-Emilian Apennines, under Abies alba, on sili-ceous soil, 5 Aug. 1999, A. Montecchi as A. borziana (AM1973, duplicate BCN JMV800239); Tuscany, Firenze, Reggello, Vallombrosa Forest, Tuscan-Emilian Apennines, ‘In silvis abiegnis Vallisumbrosae (Etruria), aestate, F. Cavara’ (NY, herb. S.M. Zeller, labelled ‘Elasmomyces mattiroloanus, Italy, Vallombrosa, coll. F. Cavara, semiepigeous in fallen needles under Abies pectinata, autumn 1896, sent to Zeller by Cavara, type’; isotype of A. borziana). Notes ― Arcangeliella borziana was collected in the summer of 1898 in Tuscany (Italy) by Cavara (1900), growing hypogeously under Abies alba. According to Zeller & Dodge (1937), collections of A. borziana preserved in their personal herbaria are duplicates of type material preserved at the Herbarium of the Università degli Studi di Napoli (Naples, Italy), but these were apparently not studied in detail by Zeller & Dodge, who just included in their works a brief description of A. borziana taken from Cavara (Zeller & Dodge 1919). Later descriptions by Singer & Smith (1960), Pegler & Young (1979), Lebel & Trappe (2000), probably based on the isotypes of Zeller & Dodge (C.W. Dodge 2087 at FH, and S.M. Zeller 1671 at NY), highlighted the same group of features: presence of a percurrent stipecolumella, hymenophoral trama and pileal context containing nests of sphaerocytes and laticifera, hymenium with abundant macrocystidia, and large ochraceous globose spores measuring 9–15 × 10–13 µm. Vidal (2004a), conducted a detailed study of these specimens, and observed that these microscopical features do not match the protologue of this species, but are more compatible with that of Elasmomyces mattiroloanus, suggesting that they do not represent the original concept of A. borziana sensu Cavara, and probably both species were erroneously labelled. This is all the more likely if laticifera had been mistaken for gloeoplera. The original concept of Lactarius borzianus is of a sessile or subsessile species that may have a rudimentary stipe-columella, a hymenophoral trama and pileal context rarely containing sphaerocytes (sometimes just represented by scattered globose elements, but never forming nests), a fertile hymenium deprived of macrocystidia (with only cystidioles present in the sterile hymenium), and whose spores are hyaline and broadly ellipsoid. It is a silicicolous species growing in wet, subalpine conifer forests of Abies and Picea, receiving large amounts of precipitation (about 2000 mm). Genetic studies by Peter et al. (2001) conducted on specimens collected in the Swiss Alps (F. Ayer 96-05-3344 WSL) show that this species is significantly related to members of L. subg. Russularia. Nuytinck et al. (2003) compared sequences of Z. stephensii with those of A. borziana obtained by Peter et al. (2001), and found both species to be genetically distinct, recombining them into Lactarius. Neither Nuytinck et al. (2003) nor Vidal (2004a) were able to observe any significant macro- or microscopical differences between the type collection of A. volemoides and A. borziana, and both taxa were therefore considered to be synonyms by these authors. Another specimen (MCVE 16944) collected at Abetina Reale (Reggio Emilia, Italy) by Amer Montecchi was sequenced by Osmundson et al. (2013) and found to be genetically identical with the Swiss collection already mentioned. Species of Lactarius subg. Russularia are characterized by reddish brown or orange pilei sometimes producing remarkable smells, bleeding white or transparent latex not changing in colour upon exposure or slowly changing to yellow, globose to ellipsoid spores ornamented with a more or less complete reticulum and an inamyloid or distally amyloid plage, as well as dermatocystidia (Hesler & Smith 1979, Wisitrassameewong et al. 2014). Other sequestrate species in Lactarius subg. Russularia include the European L. josserandii (= Z. hispanicus), the North American L. silviae (≡ Elasmomyces camphoratus) and the Australian Zelleromyces striatus and Z. daucinus, all of them matching the characteristic features of Lactarius subg. Russularia (Singer & Smith 1960, Beaton et al. 1984, Miller et al. 2001, Verbeken et al. 2014b). The North American species L. paulus (≡ Arcangeliella parva) and L. variegatus (≡ A. variegata) could also be related to subgenus Russularia, but their phylogenetic affiliation is still dubious because of the scarce genetic data available (Miller et al. 2001). Lactarius giennensis (Mor.-Arr. et al.) Pierotti, Index Fungorum 254: 1. 2015 ― Fig. 7 Basionym. Zelleromyces giennensis Mor.-Arr. et al., Cryptog. Mycol. 19, 1–2: 108. 1998. Synonym. Arcangeliella giennensis (Mor.-Arr. et al.) J.M. Vidal, Rev. Catalana Micol. 26: 74. 2004. Basidiomata 1–2.5 cm wide, angiocarpic, subglobose to tuberi-form, somewhat lobate, sessile. Pileus whitish to pale yellowish, J.M. Vidal et al.: Revision of sequestrate Russulaceae Fig. 7 Lactarius giennensis. a. JC20061020. Basidiomata. ― b–c. JMV20010204-1. b. Pileipellis and context; c. cystidioles of external immature locule. ― d–f. JMV931123-1. d. Basidiomata; e. basidia; f. spores in Melzer. ― g–i. JMV20010204-1. SEM images of spores. ― Scale bars: a, d = 1 cm; b = 20 µm; c, e–h = 10 µm; i = 5 µm. ― Photos: a. J. Cabero; b–f. J.M. Vidal; g–i. UdG. membranous, smooth, partially evanescent, indistinctly scrobiculate or with some scattered minute openings. Hymenophore loculate, whitish to pale cream or pinkish. Columella branched, inconspicuous. Latex not observed. Odour mild. Spores 10–13 × 7–9 µm, Q = 1.15–1.35, ovoid to ellipsoid, orthotropic; reticulum 1–2 μm high, complete, amyloid. Basidia 1-spored, 30–45 × 6–10 µm, cylindrical, clavate or sublageni-form. Macrocystidia absent. Cystidioles 30–42 × 4–8 µm, la-geniform-urticiform. Subhymenium ramose to cellular. Hymenophoral trama homoiomerous, made of septate hyphae 3–6 µm diam. Pileipellis and context 150–200 µm thick; suprapellis a trichoderm of septate hyphae. Pileal context a prosenchyma of intricate hyphae 3–6 µm diam. Laticifera 5–8 µm diam, present in the hymenophoral trama and pileal context. Habitat, Distribution & Season ― Gregarious, hypogeous, associated with Cistus and Halimium, often accompanied by Pinus, on siliceous soil. Autumn to spring. Xerophytic. So far restricted to the western Mediterranean region, in Central Spain, between 600–1000 m altitude. Material studied. SPAIN, Andalusia, Jaén, La Aliseda, under Pinus halepensis, 28 Feb. 1994, J. Gmez & B. Moreno-Arroyo (MA-Fungi 38674, holotype of Z. giennensis); Castilla and Leon, Le, Lugan, under Cistus sp. and Eucalyptus sp., 25 Sept. 1993, T. Pérez-Jarauta (BCN JMV930925); Sala-manca, Casarito, under Halimium alyssoides, on siliceous soil, 6 Dec. 1992, T. Pérez-Jarauta (BCN JMV921206); Salamanca, Miranda del Castar, under Cistus ladanifer, 18 Apr. 2003, A. García-Blanco, M. Sanz-Carazo & J.B. Del Val as Z. giennensis (AVM 1615, duplicate BCN JMV800629)*; Zamora, Tábara, 825 m, under Cistus ladanifer and Pinus pinaster, on siliceous soil, 4 Feb. 2001, P. Juste & F. García (BCN JMV20010204-1); ibid., 20 Oct. 2006, J. Cabero as Z. giennensis (JC20061020)*; Castilla-La Mancha, Ciudad Real, El Viso del Marqués, under Halimium ocymoides, on siliceous soil, 23 Nov. 1993, T. Pérez-Jarauta (BCN JMV931123-1). membranous pileus with perforations opening to the hymenophoral locules, not bleeding latex when cut. Spores are ovoid or ellipsoid, ornamented with a closed reticulum composed of ridges up to 2 µm high. Abundant laticifera and thromboplera were observed in the hymenophoral trama, as well as lageni-form-urticiform cystidioles in the hymenium of immature external locules. Genetic data from the holotype specimen of L. giennensis (MA-Fungi 38674, GenBank AF230900, incorrectly published as AF230800 by Calonge & Martín 2000), evidenced that this species is related with other lactarioid taxa. In the present work this species is subsignificantly related with the lineages of subg. Lactarius and Russularia, so a hypothetical affiliation with subg. Lactarius is suggested. The clade containing L. giennensis has an unusually long branch, suggesting it could represent a highly evolved or ancestral lineage, or else evidence an in-complete sampling. The holotype of L. giennensis was found at Jaén (Andalusia, Spain) growing subhypogeously under Pinus halepensis, but the collections analyzed in the present work suggest a broader distribution, being present in central Spain, associated with Cistaceae hosts in acidic soils. Lactarius josserandii (Malençon) J.M. Vidal & P. Alvarado, comb. nov. ― MycoBank MB828496; Fig. 8 Basionym. Zelleromyces josserandii Malençon, Rev. Mycol. (Paris) 39: 303. 1975. Synonyms. Arcangeliella josserandii (Malençon) J.M. Vidal, Rev. Catalana Micol. 26: 75. 2004. Zelleromyces hispanicus Calonge & Pegler, Cryptog. Mycol. 19, 1–2: 100. 1998. Notes ― Lactarius giennensis is morphologically charac-Basidiomata 2–5 cm wide, angiocarpic, irregular, obpyriform to terized by its whitish to pinkish basidiomata, with a rather thin, subglobose, tuberiform to lobulate, sessile, with a depressed or conical base immersed into the soil. Pileus dry, smooth, continuous, with minute circular depressions in the surface, but not exposing the locules; initially pale orange (5A5), then orange-red (8B8), brownish red (8C8) or reddish brown (9D8), like in Lactarius fulvissimus, remaining pale orange in unexposed areas. Hymenophore loculate, labyrinthoid, pale, initially yellowish white (4A2), then orange-white (5A2); locules 0.5–1.5 × 0.15–0.3 mm (1–2 per mm), elongated, sinuous, empty; fresh spore mass in locules whitish; yellowish white (4A2) in exsiccata. Columella absent or rudimentary, with thin whitish veins. Latex scant, white, unchanging, more abundant in the pileal context. Odour fruity, taste sweetish, later astringent. Spores 8.5–12.5(–13.5) × 7–9.5(–10.5) μm, Q = 1.1–1.3, sub- globose to ellipsoid, heterotropic, initially hyaline then yellow, reticulated; reticulum amyloid, complete or with some inter- ruptions, formed by 0.5–1 μm high ridges and warts; hilar appendix 0.7–1.4 × 0.7–1.1 μm, short, cylindrical to conical; suprahilar plage not seen. Basidia typically 4-spored, but also 1–3-spored, 40–50 × 7.5–10 μm, cylindrical, sinuous, with granular content; sterigmata 5–8 μm long, conical. Basidioles 20–30 × 7–10 µm, cylindrical. Macrocystidia absent. Cystidioles 42–44 × 7–8 µm, lageniform, scarce, present only in external immature locules. Paraphysoid cells abundant, 10–30 × 4–11 µm, entire or 1–2-septate. Pseudocystidia not observed. Sub-hymenium ramose to cellular, formed by chains of prismatic cells 8–10 μm diam. Hymenophoral trama 80–125 μm wide, Fig. 8 Lactarius josserandii. a–b. Reproduction of Malençon’s plates of Zelleromyces josserandii corresponding to the sample n° 3242 conserved in the herbarium of Montpellier (MPU C03565, ©Université de Montpellier-Herbier). ― c. MPU 0310529 (holotype of Z. josserandii). Spores in Melzer. ― d–m. JMV800621. d–e. Basidiomata; f. pileipellis and context; g–h. basidia, basidioles, paraphysoid cells and subhymenium; i. cystidiole of an external immature locule; j. spores in Melzer; k. spores in ammonia; l–m. SEM images of spores. ― Scale bars: c, i–k = 10 µm; d–e = 1 cm; f–h = 20 µm; l–m = 5 µm. ― Photos: c, f–k. J.M. Vidal; d–e. F. García; l–m. UdG. J.M. Vidal et al.: Revision of sequestrate Russulaceae homoiomerous, formed by septate, tortuous hyphae 2–7 μm diam, some of them inflated, but lacking sphaerocytes, with abundantlaticifera 2–8 μmdiam,and some thromboplera 2–6 µm diam. Pileipellis and context 175–300 µm thick; pileipellis 55–65 μm thick, arranged in an oedotrichoderm resembling that of L. fulvissimus, formed by: 1) a suprapellis of cylindrical and sinuous septate hairs, soon collapsing in a brown mass; and 2) a subpellis composed of prismatic hyphae 8–10 μm diam. Pileal context a prosenchyma, with the same composition of the hymenophoral trama, with plenty of laticifera and some thromboplera. Habitat, Distribution & Season ― Gregarious, hypogeous to semi-hypogeous, growing under needles in montane conifer forests of Pinus and Cedrus, on siliceous soil, frequently in the vicinity of Lactarius aurantiacus. Autumn. Occurring in submediterranean regions, in Morocco, France and Spain, between 1250–1900 m altitude. Material studied. MOROCCO, Taza-Al Hoceima-Taounate, Ketama, Llano Amarillo, ‘Rif cristallin, 1600 m, parmi les jeunes cèdres mêlés de Cistus laurifolius’, 16 Nov. 1957, G. Malençon (MPU 0310528, herb. G. Malençon 3242, paratype of Z. josserandii); ibid., ‘a peu de distance de la 1er (env. 300 m)’, 18 Nov. 1957, G. Malençon (MPU 0310529, herb. G. Malençon 3242 Bis, holotype of Z. josserandii); ibid., ‘sous les cèdres’, 29-X-1972, R. Bertault (MPU 0310527, herb. G. Malençon 7252, paratype of Z. josserandii). – SPAIN, Castilla and Leon, Segovia, Navafría, Puerto de Navafría, El Nevero, Sierra de Guadarrama, 1800–1900 m, under Pinus sylvestris, on siliceous soil, 6 Oct. 2006, F. García (BCN JMV800621)*; Community of Madrid, Cercedilla, Dehesas, under Pinus sylvestris, on siliceous soil, 27 Nov. 1996, F.D. Calonge & C. García-Ruz (MA-Fungi 37498, holotype of Z. hispanicus); ibid., 23 Nov. 1996, J. Daniel & J.M. Santos (MA-Fungi 37497, paratype of Z. hispanicus); ibid., 16 Nov. 1997, R. Cifuentes (MA-Fungi 38311, paratype of Z. hispanicus). Notes ― Malençon (1975) proposed thenameZelleromyces josserandii for an orange hypogeous fungus, bleeding white latex when cut, and displaying reticulate spores, found in 1957 and 1972 in Morocco and France (five collections made by G. Malençon in the Rif mountains, one found in the same area by R. Bertault, and two more collections found by R. Ruffier- Lanche and J. Mornand in the Aitone forest in Corsica under Pinus nigra subsp. laricio at 1250 m altitude). Later on, Calonge & Pegler (1998) studied several collections found under Pinus sylvestris at Sierra de Guadarrama (Madrid, Spain), with mor-phological features closely resembling those of Z. josserandii, which they named Z. hispanicus. Vidal (2004a) proposed a synonymy between Z. josserandii and Z. hispanicus after observing identical macro- and microscopical features in the type specimens, but all attempts to sequence the type material of Z. josserandii have been unsuccessful. Calonge & Pegler (1998) first suggested a close relationship between Z. hispanicus and Lactarius aurantiacus, and later Calonge & Martín (2000, 2003) highlighted the slight genetic differences (37/634 bp) between ITS sequences of these spe-cies, both belonging to Lactarius subg. Russularia. However, phylogenetic analyses conducted by Verbeken et al. (2014b), as well as present results, suggest a closer affinity to L. borzianus and L. fulvissimus, which share a similar oedotrichodermal pileipellis. Lactarius populicola J.M. Vidal, Konstantin., Setkos & Slavova, sp. nov. ― MycoBank MB828497; Fig. 9 Basidiomata similar to L. stephensii, 1–4 cm wide, angiocarpic, sessile, sub- globose or tuberiform, lactescent, smooth, at first pale orange, then reddish brown to violet brown. Hymenophore loculate, deeply coloured, reddish yellow to orange red. Columella absent or inconspicuous. Latex watery-white to white, of mild taste. Odour fruity, of pears. Microscopy as in L. stephensii, except the spores which are subglobose to broadly ellipsoid, measuring 13–15 × 11–13 μm in size, covered by strongly amyloid, conical or tooth-like warts, 1–2 μm long. Hypogeous or semi-hypogeous under Populus alba. Etymology. From Latin, populus = poplar, and -cola = inhabitant of, who lives in, in reference to its association to poplars (Populus). Holotype. GREECE, Central Macedonia, Serres, Sidirokastro, 50 m, under Populus alba, 13 Mar. 2010, G. Setkos (BCN JMV800648)*; isotype in herb. pers. G. Konstantinidis (GK4831). Basidiomata 1–4 cm wide, angiocarpic, subglobose or irregu-lar, lobate, reniform, tuberiform, sessile or with a minute sterile base, attached to soil by thin mycelial threads. Pileus smooth, viscid, persistent, open basally and exposing the locules in old specimens; at first pale orange (6A3), brownish orange (7C6) or orange-red (8A6) and finally reddish brown (8D6–9E5) to violet-brown (10E6). Hymenophore loculate, labyrinthoid, deeply coloured, reddish yellow (4A7), orange (5A6–6B7) to reddish orange (7A6), or orange-red (8A6–B6); locules very small, 0.2–0.5 × 0.02–0.1 mm (3–4 per mm), irregularly arranged, elongated, sinuous; septa 250–500 μm thick; fresh spore mass in locules pale yellow (4A4–A5); pale orange (5A3) to brownish orange (6C5) in exsiccata. Columella absent or inconspicuous. Latex watery-white in immature basidiomata and white in mature basidiomata, of mild taste. Odour strong, aromatic and fruity, reminiscent of pears. Spores 13–15 × 11–13 μm, Q = 1.07–1.2, subglobose to broadly ellipsoid, orthotropic, echinated, unigutulate, deep orange; warts deeply amyloid, robust, 1–2 μm long, conical or tooth-like, usu-ally curved, with some verrucae among them; suprahilar plage absent; hilar appendix minute, usually united to a fragment of sterigma up to 4 × 2.5 μm. Basidia 1-spored, 35–50 × 8–10 μm, cylindrical, straight or sinuous, soon collapsed; initially hyaline and filled with oleiferous guttules, then of dark orange colour; sterigmata central or eccentric, 2–4 μm long. Basidioles similar to basidia. Macrocystidia and cystidioles absent. Paraphysoid cells 20–30 × 6–8 μm, usually with 1–2 septa, cylindrical or with a clavate apex. Pseudocystidia not observed. Subhymenium ramose, formed by septate, cylindrical hyphae, 10–30 × 3–9 μm, perpendicular to the tramal hyphae. Hymenophoral trama 60–90 μm thick, homoiomerous, formed by hyaline, thin-walled, subgelatinized hyphae, 3–6 μm diam, finally dark yellow, completely gelatinized, with a prosenchymatous aspect; laticifera abundant, 8–10 μm diam, with thinner branches 3–5 μm diam; sometimes penetrating into the hymenium and terminating as pseudocystidia; thromboplera yellow, 3–5 μm diam. Pileipellis and context 300–500 µm thick; pileipellis 50–90 μm thick composed of: 1) a trichodermal suprapellis of septate hairs and dermatocystidia, 20–40 × 2.5–4 μm, straight to sinuous, rounded to acute at the apex, with yellow granular content, that soon collapses in a brown granular mass; and 2) a prosenchymatous subpellis 20–50 µm thick, of densely entangled hyphae 1.5–3 μm diam. Pileal context an ixocutis 250–400 µm thick, made by subgelatinized, entangled hyphae2–4μm diam, crossedby somethromboplera and abundant laticifera. Habitat, Distribution & Season ― Gregarious, hypogeous to semi-hypogeous in riparian forests under Populus, on alluvial soils. Spring to autumn. Distributed in temperate and Mediterranean regions, from Western to Southeastern Europe, from sea level up to 1000 m altitude. Additional material studied.BULGARIA, Stara Zagora, Parvomai, Mirovo, 130 m, in a riparian forest under Populus alba and Fraxinus ornus, 12 Dec. 2015, T. Chokova MSL1629F6516 (SOMF 29973, duplicate BCN JMV800687). – FRANCE, Franche-Comté, Jura, Hérimoncourt, Aug. 1892, L. Quélet as ʻHydnangium galatheiumʼ (UPS F013405). – GREECE, Central Macedonia, Serres, Sidirokastro, 50 m, under Populus alba, 24 Apr. 2010, G. Setkos (GK4969, duplicate BCN JMV800649); West Macedonia, Kastoria, Maniakoi, 650 m, under Populus alba, Corylus avellana and Alnus glutinosa, 7 Oct. 2010, G. Setkos (GK5223, duplicate BCN JMV800650); ibid., 25 Nov. 2011, G. Set-kos (GK5904, duplicate BCN JMV800651). – ITALY, Emilia-Romagna, Reggio Emilia, Febbio, 1050 m, under Populus alba, 30 Oct. 1991, A. Montecchi as ʻA. borzianaʼ (AM1042, duplicate BCN JMV800036); Tuscany, Lucca, Montuolo, 15 m, under Populus alba, 1 Nov. 1993, G. Bernardini & L. Gori as ʻA. stephensiiʼ (ELG931101-2, duplicate BCN JMV800171); ibid., 28 Nov. 1993, G. Bernardini & L. Gori as ʻA. stephensii ʼ (ELG931128-2, duplicate BCN JMV800169); Lucca, Nozzano Castello, 20 m, under Populus alba, Fig. 9 Lactarius populicola. a–c. MSL1629F6516. a. Mature basidiomata and cut of hymenophore bleeding white latex; b. pileipellis and context; c. section of a septum with gelatinized hymenophoral trama. ― d. GK4969. Old basidiomata. ― e–i. GK4831 (BCN JMV800648, holotype). e. Locules of hymenophore; f. spores in Melzer; g–i. SEM images of spores. ― j–l. GK5904. j–k. Young basidiomata; l. cut of hymenophore bleeding watery-white latex. ― m. UPS F013405, herb. L. Quélet (as ʻHydnangium galatheiumʼ). Spores in Melzer. ― Scale bars: a, d, j–k = 1 cm; b–c = 20 µm; e, l = 1 mm; f, m = 10 µm; g–i = 5 µm. ― Photos: a. M. Slavova; b–c, f, m. J.M. Vidal; d–e, j–l. G. Konstantinidis; g–i. UdG. 31 July 1993, G. Bernardini & L. Gori as ʻA. stephensiiʼ (ELG930731-1, duplicate BCN JMV800170). Notes ― Lactarius populicola has been confused with L. stephensii (Nuytinck et al. 2003, Gori 2005) because of their multiple similarities. Microscopically, L. populicola has the same pileal and hymenial structure, and identical spore shape with L. stephensii, but spores are intensely amyloid and ornamented with robust conical or tooth-like warts, while L. stephensii has weakly amyloid spiny spores. Macroscopically, L. populicola is distinguished by the violaceous coloration of basidiomata in mature specimens and smaller locules in the hymenophore. While L. stephensii is probably associated with Fagaceae and Betulaceae hosts, L. populicola is linked to Salicaceae plants. Genetically, L. populicola is a monophyletic taxon significantly related with L. stephensii, both forming a significantly monophyletic clade with a third species, L. soehneri. These angiocarpic species were also related (PP 1.00, BP 88) with a clade formed by samples identified as L. evosmus and L. zonarius. The only sequenced specimens of L. populicola are by now the type proposed here, and a sample from Belgium (RW2930 GENT) treated as L. stephensii by Nuytinck et al. (2003). Lactarius soehneri (Zeller & C.W. Dodge) J.M. Vidal & G. Moreno, comb. nov. ― MycoBank MB828498; Fig. 10 Basionym. Hydnangium soehneri Zeller & C.W. Dodge, Ann. Missouri Bot. Gard. 22: 372. 1935. Synonyms. Octaviania soehneri (Zeller & C.W. Dodge) Svrček in Pilát, Flora ČSR B1, Gasteromycetes: 200. 1958. Martellia soehneri (Zeller & C.W. Dodge) Singer & A.H. Sm., Mem. Torrey Bot. Club 21, 3: 31. 1960. Zelleromyces soehneri (Zeller & C.W. Dodge) Trappe et al., Mycotaxon 81: 205. 2002. Hydnangium soehneri var. ettenbergii Soehner, Z. Pilzk., N.F. 20, 3–4: 110. 1941 (nom. inval., Art. 39.1). Basidiomata 1–2 cm wide, angiocarpic, subglobose or irregular, reniform, tuberiform, sessile; sterile base not observed. Pileus smooth, viscid, at first pale orange (5A3), then greyish red (9B4) and finally reddish brown (9D4) to violet-brown (10E7), basally open and alveolate in old specimens. Hymenophore loculate, labyrinthoid, initially pinkish white (7A2), then deeply coloured, pale red (9A3), dull red (9B4) and finally dull violet (15D4); locules large, 0.4–1.5 × 0.1–0.2 mm (1–2 per mm), irregularly arranged, elongated, sinuous, empty; septa 250–500 µm thick; fresh spore mass in locules greyish ruby (12D6); J.M. Vidal et al.: Revision of sequestrate Russulaceae e. IC21091207. Basidiomata. ― f. M, herb. E. Soehner 1041 (holotype of Hydnangium soehneri var. ettenbergii). Spores in Melzer. ― g–k. M, herb. E. Soehner 1081 (holotype of Hydnangium soehneri). g. Basidia; h. spores in Melzer; i–k. SEM images of spores. ― l. JMV800677. SEM image of a spore. ― Scale bars: a, e = 1 cm; b–c = 20 µm; d, f–i = 10 µm; j–l = 5 µm. ― Photos: a. G. Moreno; b–d, f–h. J.M. Vidal; e. C. Lavoise; i–l. UdG. brownish orange (6C5) to brown (6D7) in exsiccata. Columella absent. Latex watery, scant or absent. Odour fruity. Taste earthy. Spores 14–18.5 × 12–15 μm, Q = 1.1–1.3, broadly ellipsoid to ellipsoid, orthotropic, echinate, unigutulate, pale orange; warts deeply amyloid,dense,1–2 μmlong,cylindrical,usuallycurved, sometimes fused forming short ridges, with numerous verrucae among them; suprahilar plage absent; hilar appendix 1.5–3 × 2 μm, usually attached to a fragment of sterigma. Basidia 1-spored, 50–70 × 6–10 μm, cylindrical, straight or sinuous, soon collapsed, initially hyaline, with a good number of oil drops inside, later filled with a dark orange substance that finally saturates all the hymenium; sterigmata 3–6 μm long. Basidi-oles similar to basidia. Macrocystidia and cystidioles absent. Paraphysoid cells 30–60 × 2.5–7 µm, 1–3-septate, difficult to observe. Pseudocystidia not observed. Subhymenium ramose. Hymenophoral trama 60–90 μm thick, homoiomerous, devoid of sphaerocytes, formed by hyaline, thin-walled, subgelatinized hyphae, 2–5 μm diam, displaying several yellow-brown thromboplera, 2–5 μm diam, and some laticifera, 3–10 μm diam, which sometimes penetrate into the hymenium terminating as pseudocystidia. Pileipellis and context 125–250 µm thick; pil-eipellis 70–150 μm composed of: 1) a trichodermal suprapellis of similar structure to that in L. stephensii, constituted by fragile, septate hairs and dermatocystidia, 20–30 × 3–6 μm, rounded to acute at the apex, sometimes mucronate, with dark yellow granular contents and walls up to 0.5 μm, soon collapsing into a brown granular mass formed by remnants of dermatocystidia and yellow granules; and 2) a subpellis 50–80 µm thick of prosenchymatous aspect, composed of subgelatinized hyphae 2–5 μm diam. Pileal context an ixocutis 60–120 µm thick, composed of subgelatinized, entangled hyphae, 3–6 μm diam, intersepted by abundant thromboplera and some laticifera. Habitat, Distribution & Season ― Gregarious, hypogeous to semi-hypogeous in montane forests of conifers (Abies, Pinus) or under broadleaved trees (Corylus, Quercus), on calcareous soil. Spring to autumn. Distributed in temperate regions of Germany, Italy and Spain, between 300–1400 m altitude. Material studied. GERMANY, Bavaria, ‘Ettenberg bei Berchtesgaden’, 1100 m, under Abies alba, Aug. 1925, E. Soehner (M, herb. E. Soehner 1041, holotype of H. soehneri var. ettenbergii); ‘Pupplinger Heide bei Wolfratshausen’, 19 Aug. 1919, E. Soehner as ʻHydnangium monosporumʼ (M, herb. E. Soehner 1374); ‘Pupplinger Heide bei Wolfratshausen, Frenwald’, 21 Oct. 1928, E. Soehner as ʻHydnangium carneumʼ (M, herb. E. Soehner 1081, holotype of H. soehneri; FH, herb. C.W. Dodge, isotype; NY, herb. S.M. Zeller, isotype). – ITALY, Lombardy, Bergamo, San Pellegrino Terme, 350 m, under Corylus avellana, on calcareous soil, Sept. 2016, M. Berbenni, comm. M. Morara (BCN JMV800677); Umbria, Perugia, Gubbio, 340 m, under Quercus sp. and Corylus avellana, in a truffle plantation of Tuber melanosporum, on calcareous soil, 21 Sept. 2012, A. Paz & C. Lavoise (BCN IC21091207). – SPAIN, Castilla-La Mancha, Guadalajara, Cantalojas, 1300–1400 m, in humus of Pinus sylvestris, on calcareous soil, 3 Apr. 2011, M.A. Sanz (AH 39272)*; ibid., 9 Apr. 2011, M.A. Sanz (AH 46013, duplicate BCN JMV800633)*. Notes ― Lactarius soehneri was considered a macrosporic form of L. stephensii by several authors (Mader & Mader 1992, Vidal 2004a), because of the very similar microscopic features, especially the presence of laticiferous hyphae and monosporic basidia documented in the type collection by Vidal (2004a). However, L. soehneri differs from L. stephensii and L. popu-licola by its scarce, watery latex, by the larger hymenophoral locules and by the ellipsoid spores up to 18.5 × 15 µm. The dark orange substance of hymenium that impregnates basidia and spores is another differentiating feature. While L. stephensii and L. populicola occur exclusively under broadleaved trees, L. soehneri also occurs under conifers. Genetically, L. soehneri is significantly related with L. populicola and L. stephensii, and to a lesser extent to the gymnocarpic species L. evosmus and L. zonarius, and probably belongs to Lactarius subg. Lactarius. Unfortunately, no samples from the region where the type was found (Bavaria, Germany) or the same habitat (Abies or Picea forests) could be analyzed, and therefore the identification made here relies entirely on mor- phological similarities between the newly collected samples and the type specimen. Lactarius stephensii (Berk.) Verbeken & Walleyn, Belg. J. Bot. 136, 2: 151. 2003 ― Fig. 11 Basionym. Hydnangium stephensii Berk., Ann. Mag. Nat. Hist., Ser. I, 13: 352. 1844. Synonyms. Octaviania stephensii (Berk.) Tul. & C. Tul., Fung. Hypog.: 78. 1851. Arcangeliella stephensii (Berk.) Zeller & C.W. Dodge, Ann. Missouri Bot. Gard. 18: 463. 1931. Zelleromyces stephensii (Berk.) A.H. Sm., Mycologia 54: 635. 1962. Martellia stephensii (Berk.) K. Mader & A. Mader, Österr. Z. Pilzk. 1: 4. 1992. Hydnangium galatheium Quél., Enchir. Fung.: 247. 1886 (ʻgalathejumʼ). Octaviania galatheia (Quél.) De Toni in P. A. Saccardo, Syll. Fung. 7, 1: 491. 1888 (ʻgalathejaʼ). Basidiomata 1–3.5 cm wide, angiocarpic, subglobose to tuberi-form, lobate, sessile, sometimes with a minute sterile base; basally closed, or open and radially alveolate in old specimens. Pileus smooth,viscid,atfistwhite,then maize yellow (4A6),oxide yellow (5C7) to carrod red (6B7), finally reddish brown (7D8– E8), with the same colour as L. rufus. Hymenophore loculate, labyrinthoid, initially white or ochraceous, acquiring a reddish brown colour like the pileus; locules small, 0.3–1.5 × 0.1–0.5 mm (2–3 per mm), irregularly arranged, elongated, sinuous; septa 150–300 µm thick; fresh spore mass in locules greyish orange (5B5). Columella absent or inconspicuous. Latex scant and hyaline, changing to citrine yellow in young specimens, white, abundant, and almost immutable in old basidiomata (orange in contact with KOH), with a bitter and more or less hot taste as in L. chrysorrheus. Odour mild in young basidiomata, but intensely fruity when mature. Spores (11.5–)12–14.5(–15) × (10–)11–13 µm, Q = 1.1–1.2, sub-globose to broadly ellipsoid, orthotropic, echinate; warts 0.5–2 µm high, isolated, inamyloid or weakly amyloid; suprahilar plage J.M. Vidal et al.: Revision of sequestrate Russulaceae absent; hilar appendix inconspicuous, usually united to a fragment of sterigma up to 4 µm long. Basidia 1-spored, 35–70 × 6–10 μm, cylindrical, straight to sinuose, soon collapsed; initially hyaline, with a good number of oil drops inside, that become filled with a dark orange substance; sterigmata central or lateral, 4–6 µm long. Macrocystidia absent. Cystidioles only present in sterile locules, of the same shape and size of dermatocystidia. Paraphysoid cells plenty, 20–60 × 6–8 µm, cylindrical or clavate, usually with 1–3 septa. Pseudocystidia occasionally observed. Subhymenium ramose, formed by cylindrical hyphae 10–30 × 3–9 µm, perpendicular to the tramal hyphae. Hymenophoral trama homoiomerous, formed by subgelatinized hyphae 2–5 µm diam, with abundant laticifera. Pileipellis and context 250–500 µm thick; suprapellis a trichoderm formed by yellow, straight to sinuous, septate dermatocystidia 20–60 × 3–6 µm, with rounded to acute, sometimes mucronate apex, that soon collapses in a brown granular mass. Pileal context formed by intricate gelatinized hyphae 2–6 µm diam, crossed by some thromboplera and abundant laticifera 3–10(–12.5) µm diam. Habitat, Distribution & Season ― Gregarious, hypogeous to semi-hypogeous among plant debris, in montane woods of Corylus, Carpinus, Fagus, Quercus accompanied by Acer, Buxus, Cornus, Fraxinus, Sambucus, Tilia, Ulmus, on calcareous soil. Summer to autumn. Widely distributed in temperate and submediterranean regions, from almost sea level in the British Isles to 1400 m altitude in Southern Europe. Material studied. CZECH REPUBLIC, Central Bohemia, Karlštejn, in a leafy forest, under Fagus sylvatica and Carpinus, 10 July 1949, V. Vacek as ʻHydnangium monosporumʼ (PRM 685987); ibid., in a leafy forest, under Quercus, Carpinus, Acer campester and Tilia, 17 July 1949, V. Vacek as ʻHydnangium monosporumʼ (PRM 619116); South Moravia, Dolní Věstonice, Děvičky, in Acereto-Carpineto, 28 Aug. 1955, K. Kříž, det. M. Svrček as O. stephensii (PRM 719216); Veverská Bítka, Krnovec, under Carpinus betulus and Cornus mas, 10 Aug. 1955, K. Kříž, det. M. Svrček as O. stephensii (PRM 719217); Žarošice, Aug. 1937, V. Vacek as ʻHydnangium carneumʼ (PRM 154169, coll. J. Velenovsk; ibid., in a leafy forest, 31 Aug. 1948, Vl. Vacek & V. Vacek as H. stephensii (PRM 685988); ibid., in a leafy forest, under Tilia, Quercus and Larix, 25 Aug. 1949, V. Vacek as H. galatheium (PRM 619118); ibid., in a leafy forest, under Tilia, Quercus and Larix, 25 Aug. 1949, V. Vacek as H. galatheium (PRM 685993); Ždánský les, near Žarošice, in a leafy forest, 31 Aug. 1950, V. Vacek as H. galatheium (PRM 685990); ibid., in a leafy forest, under Tilia, Quercus and Fraxinus, 9 Sept. 1950, V. Vacek as H. galatheium (PRM 685992); Zdravá Voda, near Žarošice, in a leafy forest, under Tilia, Carpinus, Fagus and Cornus mas, 24 Aug. 1949, V. Vacek as H. galatheium (PRM 685996); ibid., in a leafy forest, under Carpinus, Quercus and Tilia, 29 Aug. 1949, V. Vacek as H. galatheium (PRM 685994). – HUNGARY, Central Hungary, Budapest, Normafa, under Tilia sp., 25 Aug. 1995, I. Király (BCN JMV800071). – ITALY, Lombardy, Bergamo, San Pellegrino Terme, 350 m, under Corylus avellana, on calcareous soil, 30 Nov. 2016, M. Berbenni, comm. M. Morara (BCN IC30111603). – POLAND, Silesia, Pogórze Śląskie region, Machowa mountain, Western Carpathians, 15 June 2014, R. Rutkowski (KRA F-2014-147, duplicate BCN JMV800668)*. – SPAIN, Asturias, Somiedo, Parque Natural de Somiedo, under Corylus avellana, 8 Apr. 2009, A. Paz (BCN IC08040913); Catalonia, Girona, Albanyà, Serra de Corsavell, Can Padern, 800 m, under Corylus avellana, on calcareous soil, 9 June 2001, J.M. Vidal as A. stephensii (BCN JMV20010609-1)*; Girona, La Vall de Bianya, Vall del Bac, Bac de Mariner, 900 m, under Corylus avellana, Fagus sylvatica, Quercus humilis and Buxus sempervirens, on calcareous soil, 30 Aug. 1997, J.M. Vidal as A. stephensii (BCN JMV970830-1); Girona, Montagut, Sant Miquel de Pera, 700 m, under Corylus avellana, Quercus humilis, Populus tremula and Buxus sempervirens, on calcareous soil, 15 Oct. 1995, J.M. Vidal as A. stephensii (BCN JMV951012-1). – UNITED KINGDOM, England, Somerset, Clifton, near Bristol, Leigh Woods, 6 Aug. 1843, C.E. Broome (K(M)69330, lectotype of H. stephensii); ibid., Bristol, 1844, H.O. Stephens (K(M)69331, original material of H. stephensii); ibid., Bristol, sine dat., ex M.J. Berkeley (PRM 719218, herb. A.C.J. Corda, original material of H. stephensii). Notes ― Hydnangium stephensii is a lactescent angiocarpic fungus described by Berkeley (1844) from Leigh Woods, North Somerset (UK), which was recombined in the genus Zelleromyces by Smith (1962) based on the presence of amyloid spores and laticifera. It is macroscopically characterized by maize yellow to reddish brown basidiomata bleeding white latex (scant and hyaline when young) that turns yellow in contact with air, and microscopically, by its monosporic basidia and warty subglobose to broadly ellipsoid spores, which are inamyloid to weakly amyloid (Vidal 2004a, Fraiture & Derboven 2009, all as A. stephensii). Nuytinck et al. (2003) suggested L. stephensii belongs to Lac tarius subg. Lactarius, and considered it a prioritary synonym of Hydnangium monosporum, while Vidal (2004a) considered L. stephensii (treated as Arcangeliella stephensii) to be a prioritary synonym of Hydnangium soehneri. Following careful studying of the types and other collections in the present study, L. stephensii, H. soehneri and H. monosporum are here con-sidered to be three distinct species. Hydnangium monosporum differs by its perfectly rounded and pink-coloured spores when mature, and H. soehneri by its bigger, strongly amyloid and ellipsoid spores. It was not possible to locate original material of H. galatheium (Quélet 1875 as ‘stephensii’, 1886 as ‘galathejum’), but only a collection made later by Quélet in August 1892 at Hérimoncourt (UPS F013405),which we have identified as L. populicola, and another collection without date from Ab-bévillers (UPS F013936), which we have identified as Russula monospora. However, based on descriptions and illustrations provided by Quélet (1875: 446, pl. I, f. 9, as H. stephensii) and (1886: 247, as H. galatheium), we consider, like Patouillard (1910, 1914) and Zeller & Dodge (1937), H. galatheium to be a synonym of L. stephensii. The only genetic sequence available in GenBank before the present work from L. stephensii, was an ITS rDNA sequence of K(M)64067 specimen (Brock et al. 2009), collected by Brian Spooner in 1999 at North Somerset (UK), probably close to the locality where the type was found. This sequence clearly differs from that produced from speci-men RW2930 (GENT) by Nuytinck et al. (2003), which is here considered to represent L. populicola. Lactarius subgiennensis Loizides, J.M. Vidal & P. Alvarado, sp. nov. ― MycoBank MB828499; Fig. 12 Basidiomata 0.5–2.5 cm wide, angiocarpic, sessile, globose to subglobose or lobate, lactescent, scrobiculate-porate, ivory white. Hymenophore loculate, whitish to faintly pinkish. Latex white. Spores 8.5–11 × 6.5–8 μm, ovoid, reticulated; reticulum 0.5–1 µm high, partially interrupted. Basidia 1-spored, 30–42 × 5–9 μm, lageniform-urticiform. Cystidioles 32–40 × 5–8 µm, lageniform-urtici-form, present only in external locules. Macrocystidia absent. Hymenophoral trama homoiomerous, prosenchymatous, gelatinized, with laticifera 3–9 µm diam. Suprapellis formed by repent to semi-erect, septate hyphae 4–6 μm wide. Pileal context prosenchymatous, gelatinized, with abundant laticifera and some thromboplera 4–14 µm diam. Semi-hypogeous in Cistus maquis on the island of Cyprus. Etymology. From Latin, sub = nearly, close to, and -giennensis, referring to its close phylogenetic and morphological affinity to Lactarius giennensis. Holotype. CYPRUS, Larnaca District, Kalavasos, 170 m, under Cistus salviifolius and Cistus creticus, 25 Nov. 2012, M. Loizides (BCN JMV800627)*; isotype in herb. pers. M. Loizides (ML211152E). Basidiomata 0.5–2.5 cm wide, angiocarpic, irregularly globose, subglobose or tuberiform to weakly lobulate, occasionally more distinctly lobulate to almost irregularly-shaped, sessile, lactescent, loosely attached to the substrate by rhizoids. Pileus smooth, silky, subtomentose in parts, moderately to distinctly scrobiculate and porate, ivory white (5A2–6A2), often with ochraceous stains at maturity (5A4–A5), overripe basidiomata sometimes brownish buff (6C6); greyish orange (5B5) in exsiccata. Hymenophore loculate, labyrinthoid, white when young, at maturity becoming cream (6A2–A3) or pale ochraceous salmon (7A2); locules 0.5–1 mm long (1–3 per mm), polygonal to somewhat elongated, irregular, mostly empty. Columella ab-sent or rudimentary, but thin, sterile veins frequently intersecting the hymenophore. Latex abundant in humid conditions, white, unchanging. Odour acidic, slightly fruity. and context with one thromboplera and some laticifera; d. detail of the pileipellis; e–f. hymenium (basidia, basidioles and paraphysoid cells), subhymenium and hymenophoral trama with some laticifera; g. cystidioles of external immature locule; h. emerging necks of basidia; i. basidia; j. spores in Melzer; k–l. SEM images of spores. ― Scale bars: a–b = 1 cm; c–f = 20 µm; g–j = 10 µm; k–l = 5 µm. ― Photos: a–b. M. Loizides; c–j. J.M. Vidal; k–l. UdG. Spores 8.5–11 × 6.5–8 μm, Q = 1.2–1.4, ovoid, orthotropic, re-ticulated; reticulum strongly amyloid, formed by well-developed crests projecting 0.5–1 μm in side view, partially interrupted to nearly complete at full maturity; hilar appendix indistinct or rudimentary. Basidia 1-spored (rarely 2-spored), 30–42 × 5–9 μm (neck included), lageniform-urticiform, promptly collapsed, with a long and sinuous neck measuring 20–25 × 2.5–4 µm, protruding the hymenium; sterigmata ± 1 μm long, inconspicuous. Basidioles 22–24 × 4.5–6 µm, ventricose-rostrate. Macrocystidia absent. Cystidioles 32–40 × 5–8 µm, lageniform-urtici-form, observed only in immature external locules. Paraphysoid cells 9–25 × 5–8.5 μm, cylindrical to subclavate, entire or 1-septate. Subhymenium ramose to cellular, composed of chains of rounded, polygonal or irregularly elongated cells 5–10 μm diam. Hymenophoral trama 90–120 μm wide, homoiomerous, prosenchymatous, formed by intricate, tortuous, thin-walled hyphae 3–6 µm diam, with plenty of lipidic guttules, finally of pseudoparenchymatous aspect, moderately to strongly gelatinized, but without the presence of sphaerocytes; laticifera 3–9 μm diam, thick-walled, tortuous, often branching, sometimes forming projecting pseudocystidia. Pileipellis and context thin, 50–100 µm thick; pileipellis composed of: 1) an easily collapsable trichodermal suprapellis, comprised of terminal, septate hyphae 4–6 μm diam; and 2) a prosenchymatous subpellis of gelatinized, tightly packed, tortuous, septate hyphae 3–6 μm diam. Pileal context undifferentiated from the subpellis with abundant laticifera and some thromboplera 4–11 µm diam, locally enlarged up to 14 µm. Habitat, Distribution & Season ― Gregarious to caespitose, growing semi-hypogeously in thick litter, exclusively associated with Cistus, on siliceous soil. Autumn and winter. Xerophytic. So far restricted in the eastern Mediterranean region on the island of Cyprus, at altitudes below 400 m. Additional material studied. CYPRUS, Larnaca District, Kalavasos, 170 m, under Cistus salviifolius and Cistus creticus, 29 Nov. 2012, M. Loizides (ML211192Z); ibid., 200 m, 23 Jan. 2016, M. Loizides (ML61132Z)*; Nicosia District, Lythrodontas, 400 m, under Cistus salviifolius and Cistus creticus, 27 Nov. 2014, M. Loizides (ML411172Z)*. Notes ― Lactarius subgiennensis differs from the closely related L. giennensis in its markedly scrobiculate basidioma-ta, lageniform-urticiform basidia, and smaller, subreticulated spores with a partially incomplete or discontinuous reticulum, not exceeding 1 µm height in side view. Both species are found in xerophytic environments of Southern Europe in association J.M. Vidal et al.: Revision of sequestrate Russulaceae with Cistaceae plants, but seem to be geographically isolated. Additional collections from intermediate localities (Greece, Italy) would be needed to confirm the distribution boundaries between both taxa. Genetically, L. subgiennensis seems to be the sister-species of L. giennensis. Both taxa seem to have some intraspecific variability, but are significantly different. Russula ammophila (J.M. Vidal & Calonge) Trappe & T.F. Elliott, Fungal Systematics and Evolution 1: 231. 2018 ― Fig. 13 Basionym. Gymnomyces ammophilus J.M. Vidal & Calonge, Bol. Soc. Micol. Madrid 24: 66. 1999. Synonym. Macowanites ammophilus (J.M. Vidal & Calonge) J.M. Vidal & Calonge, Rev. Catalana Micol. 24: 70. 2002. Basidiomata russuloid, angiocarpic to pseudoangiocarpic, stipitate. Pileus 2–7 cm wide, globose to hemisphaerical, orange white to pale orange, with brownish orange and dark brown maculae; margin open, alveolate to sublamellate. Hymenophore loculate, orange-white to pale orange. Stipe-columella 1–4 × 0.7–2.5 cm, with brownish orange dots; context white. Odour fruity; taste sweetish to acrid. Spores 7–9 × 5.5–7.5 μm, Q = 1.16–1.4, subglobose to ovoid, heterotropic; warts 0.25–0.75 µm high, amyloid, some forming short ridges or connected by short lines. Basidia 2–4-spored, 35–45 × 8–11 μm, clavate. Macrocystidia 45–60 × 6–10 µm. Hymenophoral trama with sphaerocytes 6–35 µmdiam.Pileipel-lis 125–200 µm thick; suprapellis an intricate trichoderm made of sinuous, clavate, lanceolate or mucronate tips of hyphae and some poorly differentiated dermatocystidia 15–40 × 2–5 µm; subpellis an intricate cutis of gelatinized hyphae 2.5–7.5 µm diam. Pileal and stipe-columella context heteromerous. Habitat, Distribution & Season ― Gregarious, hypogeous to semi-hypogeous in sandy soils, associated with Pinus. Spring and autumn. Common in sandy pine forests of the Atlantic coast of Southern Spain and Portugal, more rarely in Northern and Central Spain. Material studied. PORTUGAL, Setubal, road from Grândola to Santa Margarita do Sado, stabilized dunes, under Pinus pinea and Pinus pinaster with Halimium halimifolium and Corema, 26 Mar. 1998, P.P. Daniëls, J.M. Vidal & F.D. Calonge as G. ammophilus (MA-Fungi 40132; duplicate BCN JMV980326-11b); Castelo Ventoso, road to Alcácer do Sal, stabilized dunes, under Pinus pinea with Corema, 26 Nov. 1999, J.M. Vidal & F.D. Calonge as M. ammophilus (MA-Fungi 51167; duplicate BCN JMV991126-1); without loc., without date, as M. ammophilus (AH 42956)*. – SPAIN, Andalusia, Cádiz, Barbate, Loma del Teniente, 120 m, stabilized dunes, under Pinus pinea with Halimium, 15 Dec. 2012, M. Becerra & E. Robles (AH 43950, as M. ammophilus)*; Cádiz, Sanlar de Barrameda, Pinar de Algaida, stabilized dunes, under Pinus pinea, 30 Oct. 2014, M. Becerra as M. ammophilus (AH 46370)*; Huelva, Matalascas, stabilized dunes, under Pinus pinea with Halimium, Corema and Juniperus phoenicea, 27 Nov. 1999, J.M. Vidal & F.D. Calonge as M. ammophilus (MA-Fungi 51165; duplicate BCN JMV991127-8); Mazag, stabilized dunes, under Pinus pinea with Halimium and Corema, 27 Nov. 1999, J.M. Vidal & F.D. Calonge as M. ammophilus (MA-Fungi 51166; dupli-cate BCN JMV991127-2); Castilla and Leon, Palencia, Dehesa de Montejo, Tosande, under Pinus nigra subsp. salzmannii and Quercus rotundifolia, in sandy basic soil, 9 Jan. 2007, A. Paz (BCN IC09010703)*. Notes ― Gymnomyces ammophilus was proposed to ac-commodate specimens of a fully angiocarpic basidiomycete without any trace of a stipe-columella, found in South-Western Spain and Portugal (Calonge & Vidal 1999). However, new specimens with an evident stipe-columella were later collected by the same authors (Vidal et al. 2002) and DNA data suggested they all represent a single species, so it was concluded that the holotype material was not fully mature and the taxon was re-combined as Macowanites ammophilus. Since its formal description, this species has been regularly collected in coastal sand dunes of the South-Western Iberian Peninsula and a high morphological variability has been observed. Recently, this species was re-combined into genus Russula by Elliott & Trappe (2018). A high genetic variability (5.8 %) among ITS sequences from these specimens can be observed (Fig. 3). This could be due in part to sequencing errors in older GenBank accessions (M.P. Martín, pers. comm.), although some variability was found also among newly produced sequences. Most interestingly, ITS sequences did not support significant differences between R. ammophila and sequences of the gymnocarpic species R. amoenolens produced in the present work and those present in public databases, including one obtained from a specimen identified byRomagnesi (MICH12838,KF245510,Bazzicalupo & Berbee, unpubl. data). These European sequences probably represent the original concept of R. amoenolens, and are significantly different from other lineages identified with the same name (Fig. 3). In addition, 28S rDNA and rpb2 data failed to support any significant difference between R. ammophila and R. amoenolens, but translation elongation factor 1-alpha (tef1) data contains diagnostic differences that can be employed to discriminate them (GenBank MK102724–MK102727). Owing to e. JMV991127-2. Spores in Melzer. ― f–g. JMV991126-1. SEM images of spores. ― Scale bars: a = 1 cm; b–c = 20 µm; d–e = 10 µm; f–g = 5 µm. ― Photos: a–e. J.M. Vidal; f–g. UdG. genetic differences in tef1 gene and the lack of a stabilized con-cept of R. amoenolens, we here still accept that R. ammophila represents an independent species. Both taxa belong to R. sect. Ingratae, and are nested within the clade of /R. amoenolens, closely related with R. cerolens, R. sororia, and several other linages identified as R. amoenolens not matching Romagnesi’s concept of this species. In R. amoenolens spores are ellipsoid, measuring 7–8.5(–9) × 5–6.7 μm and ornamented with warts upto0.75µm high(Romagnesi1985), or 6.5–8.4 × 5.2–6.4μm and with warts up to 0.8 µm high (Sarnari 1998), some warts interconnected with short ridges. Spores of R. ammophila are subglobose to ovoid, measuring 7–9 × 5.5–7.5 μm and orna-mented with warts up to 0.75 µm high, some of them fused and connected by short ridges. Russula amoenolens Romagn., Bull. Mens. Soc. Linn. Lyon 21: 111. 1952. Material studied. SPAIN, Cantabria, Castro Urdiales, Prado Maya, under Quercus sp., 23 July 2015, J.A. Cadinos (AH 46371)*; Castro Urdiales, under Quercus sp., 23 July 2015, J.A. Cadinos (AH 46372)*. Russula andaluciana T.F. Elliott & Trappe, Fungal Systematics and Evolution 1: 239. 2018 ― Fig. 14 Replaced synonym. Gymnomyces sublevisporus Mor.-Arr., Llistos. & L. Romero, Rev. Catalana Micol. 24: 179. 2002. Basidiomata 0.5–2 cm wide, angiocarpic, globose to sub-globose, somewhat lobate, sessile. Pileus smooth, pruinose, whitish, with brownish red maculae. Hymenophore loculate, whitish at first, finally brownish red. Columella percurrent, sometimes branched, inconspicuous. Odour of bitter almonds; taste mild, slightly bitter. Spores (6.5–)7.5–9.5(–11.5) × (6–)7–9(–11) µm, Q = 1.0–1.1, globose to subglobose, orthotropic; warts up to 0.3 µm high, amyloid, some connected with low ridges. Basidia 4-spored, 36–50 × 7.5–11 μm, clavate. Macrocystidia and cystidioles absent. Hymenophoral trama prosenchymatous, of dense inter- woven, tortuous hyphae about 2–6 µm diam, with some inflated ampullaceous cells up to 10 µm diam. Pileipellis and context 125–200 µm thick; suprapellis a trichoderm formed by short, cy-lindrical to clavate hairs 9–18 × 5–7 µm, and long, subulate, sometimes 1–2-septate hairs 30–49 × 2–7 µm; dermatocystidia absent; subpellis made by ampullaceous cells 8–10 µm diam. Pileal context with the same structure as the hymenophoral trama. Habitat, Distribution & Season ― Gregarious, hypogeous, associated with Cistus ladanifer, often accompanied by Quercus rotundifolia or Q. suber, on siliceous soil. Spring. Located in the Western Mediterranean region, in Central Spain, between 400–800 m altitude. Material studied. SPAIN, Andalusia, Huelva, Cortelazor, finca Galindo, 540 m, under Cistus ladanifer with Cistus crispus, Cistus salviifolius and Quercus rotundifolia, 11 Mar. 1995, L. Romero de la Osa (BCN JL5101, holotype of G. sublevisporus; BM360, isotype)*; Extremadura, Badajoz, Mirandilla, Sierra Bermeja, 400 m, under Cistus ladanifer with Quercus suber, on sili-ceous soil, 25 Mar. 2017, J.L. Becerra (BCN IC25031720); Cáceres, Cuesta de Jaraiz, under Cistus ladanifer, on siliceous soil, 2 Apr. 2011, C. Gelpi as G. sublevisporus (AH 39239)*; Cáceres, Cuesta de las Veguillas, under Cistus ladanifer with Eucalyptus camaldulensis, on siliceous soil, 2 Mar. 2011, C. Gelpi & J. Muz as G. sublevisporus (AH 39198)*. Notes―As already mentioned above, R. andaluciana is a replacement name proposed by Elliott & Trappe (2018) to re-combine G. sublevisporus into Russula. This species is distinguished by its whitish to yellowish basidiomata, and by its globose spores, ornamented with very small verrucae not d. hymenium (basidia and paraphysoid cells); e–f. spores in Melzer; g–i. SEM images of spores. ― Scale bars: a–b = 1 cm; c–d = 20 µm; e–f = 10 µm; g–i = 5 µm. ― Photos: a–b. C. Gelpi; c–f. J.M. Vidal; g–i. UdG. J.M. Vidal et al.: Revision of sequestrate Russulaceae exceeding 0.3 µm high. Sphaerocytes were not found in our observations, neither in the hymenophoral trama nor in the pileal context. Genetically, R. andaluciana is the only known European sequestrate species in Russula sect. Rigidae. It is nested within R. subsect. Amoeninae, but shows no clear relationship with any other taxon (Fig. 4). Russula bavarica J.M. Vidal, sp. nov. ― MycoBank MB828500; Fig. 15 Basidiomata 1–2 cm wide, angiocarpic, sessile, subglobose to tuberiform, subtomentose, pale orange with brown maculae. Hymenophore pale orange, forming small cells or locules. Spores 13–15(–15.5) × 12.5–14.5(–15) µm, globose to subglobose, orthotropic, intensely yellow, weakly amyloid, orna-mented with dense, curved warts, 1.5–3 µm long. Basidia 1-spored, 24–36 × 8–12 µm, clavate. Macrocystidia 30–70 × 8–16 µm, clavate, abundant, containing refringent crystals (in exsiccata). Subhymenium ramose. Hymenophoral trama homoiomerous. Suprapellis a trichoderm of repent to semierect hyphae and dermatocystidia 50–100 × 2–4 µm; pileal context a subgelatinized cutis. Etymology. From Latin, bavaricus = refers to the German state of Bavaria, where mycologist Ert Soehner collected this sequestrate fungus. Holotype. GERMANY, Bavaria, ‘Wolfratshausen bei Mchen, Pupplinger Heide in der Nähe der Aumle’, 8 July 1923, E. Soehner #724 as ʻHydnangium carneumʼ (M); isotype in FH (herb. C.W. Dodge). Basidiomata 1–2 cm wide, angiocarpic, subglobose to tuberi-form, sessile. Pileus finely tomentose, white to pale orange (5A4) in young specimens, partially covered with brownish maculae (7E6) in mature basidiomata. Hymenophore loculate, labyrinthoid, pale orange in colour (5A4–A5); locules 0.4–1 × 0.1–0.4 mm, elongated, sinuous; spore mass in locules pale orange (5A3). Columella absent. Odour and taste unknown. (Description based on herbarium material). Spores 13–15(–15.5) × 12.5–14.5(–15) µm, Q = 1.0–1.1, glo-bose to subglobose, orthotropic, echinate, hyaline when young, yellowish when mature; warts inamyloid or weakly amyloid, 1.5–3 µm long, dense, curved, sharp, with some verrucae among them; hilar appendix short, 0.5–1 µm long, often with remnants of the sterigmata attached. Basidia 1-spored, 24–36 × 8–12 µm, clavate, hollow; sterigmata 3–5 µm long, conical. Basidioles 15–28 × 6–10 µm, clavate. Macrocystidia 30–70 × 8–16 µm, clavate, abundant, filled with highly refringent hyaline needles. Pseudocystidia not observed. Subhymenium very reduced, ramose, formed by short chains of 1–3 cylindrical elements, measuring 3–10 × 3–7 µm. Hymenophoral trama 20–40 µm wide, homoiomerous, formed by an entanglement of branched, tortuous, septate, subgelatinized hyphae about 2–6 µm diam,including scattered endomacrocystidia;sphaerocytes absent. Pileipellis and context thin, 100–150 µm thick; pileipellis 25–50 µm thick, arranged as a trichoderm formed by hyphae and septate, repent or semi-erect dermatocystidia 50–100 × 2–4 µm, finally collapsing into a brownish mass. Pileal context 75–100 µm thick, formed by a cutis of intricate, subgelatinized hyphae 2–5 µm diam, connecting with the hymenophoral trama. Laticifera not observed. Gloeoplera abundant, 2–5 µm diam, present in trama and context. Notes ― Collection Soehner num. 724 was cited by Dodge & Zeller (1937) and Soehner (1941), under the name Hydnangium carneum. It was found by Soehner in July 1923 near Munich (Bavaria, Germany). After studying the specimens preserved nowadays in M and FH herbaria, amyloid spores, hymenial macrocystidia, dermatocystidia and gloeoplera were observed, suggesting this species does not belong to genus Hydnangium. These specimens resemble R. monospora, because of the presence of monosporic basidia, echinate and weakly amyloid spores, ramose subhymenium, homoiomerous hymenophoral trama, and trichodermal pileipellis. Although it has not been possible to obtain recent collections matching these specimens, we believe it necessary to accommodate the collection Soehner 724 in a new species of Russula because of its unique combination of microscopical features, very similar to those of R. monospora, but differing and being characterized by the presence of clavate macrocystidia and larger spores ornamented with longer and more acute and dense spines. Further collections are needed to confirm the ecology of this species, which is probably the same as Lactarius soehneri, whose holotype was also collected by Soehner in Pupplinger Heide (south of Munich), and also determined as H. carneum. Russula candida (Tul. & C. Tul.) J.M. Vidal, comb. nov. ― Myco-Bank MB828501; Fig. 16 Basionym. Hydnangium candidum Tul. & C. Tul., Ann. Sci. Nat., Bot., Sér. 2, 19: 376. 1843. Synonyms. Octaviania candida (Tul. & C. Tul.) Lloyd, Mycol. Writings 7, Letter 67: 1142. 1922. Sclerogaster candidus (Tul. & C. Tul.) Zeller & C.W. Dodge, Ann. Missouri Bot. Gard. 23: 570. 1937. Macowanites candidus (Tul. & C. Tul.) J.M. Vidal, Rev. Catalana Micol. 26: 84. 2004. Secotium krjukowense Bucholtz, Hedwigia 40: 314. 1901. Elasmomyces krjukowensis (Bucholtz) Sacc. & D. Sacc. in P.A. Saccardo, Syll. Fung. 17: 218. 1905. Bucholtzia krjukowensis (Bucholtz) Lohwag, Oesterr. Bot. Z. 73, 7–9: 173. 1924. Arcangeliella krjukowensis var. krjukowensis (Bucholtz) Zeller & C.W. Dodge, Ann. Missouri Bot. Gard. 22: 368. 1935. Hydnangium krjukowense var. krjukowense (Bucholtz) Svrček in Pilát, Flora ČSR B1, Gasteromycetes: 206. 1958. Macowanites krjukowensis (Bucholtz) Singer & A.H. Sm., Mem. Torrey Bot. Club 21, 3: 77. 1960. Russula krjukowensis (Bucholtz) Trappe & T.F. Elliott, Fungal Systematics and Evolution 1: 235. 2018. Secotium michailowskianum Bucholtz, Hedwigia 40: 315. 1901. Elasmomyces michailowskianus (Bucholtz) Sacc. & D. Sacc. in P.A. Saccardo, Syll. Fung. 17: 218. 1905 (ʻmichailowskjanusʼ). Arcangeliella krjukowensis var. michailowskiana (Bucholtz) Zeller & C.W. Dodge, Ann. Missouri Bot. Gard. 22: 368. 1935. Secotium krjukowense forma pleurotopsis Bucholtz, Bull. Soc. Imp. Na-turalistes Moscou 4: 463. 1908 (1907). Octaviania moravica Velen., Opera Bot. Cech. 4: 95. 1947. Hydnangium krjukowense var. moravicum (Velen.) Svrček in Pilát, Flora ČSR B1, Gasteromycetes: 208. 1958. Basidiomata russuloid, angiocarpic, stipitate. Pileus 0.5–3.5 cm wide, globose to subglobose, sometimes lobate, smooth, pruinose, white to yellowish white (4A2), with yellowish orange (4A6) to dark orange (5A8) maculae; margin closed or laterally open, alveolate. Hymenophore loculate, labyrinthoid, pale orange (5A3–A4); locules 0.7–3 × 0.1–0.4 mm (1–2 per mm), elongated, irregular, sinuous; fresh spore mass in locules pale yellow (3A3–A4). Stipe-columella 0.6–1.5 × 0.15–0.4 cm, usu-ally half stipe, half columella, cylindrical to clavate, central or lateral, sometimes deeply lateral and thus barely visible, pure white, pruinose; context white, not changing on exposure to air. Odour fruity; taste mild. Spores 8.5–11(–12.5) × 7–9(–11) μm, Q = 1.08–1.25(–1.3), subglobose to broadly ellipsoid, heterotropic to subheterotropic, echinate; warts amyloid, 0.5–1.5 µm high, isolated, cylindrical and sometimes curved, with a rounded apex; hilar appendix 1–2.5 × 1.2 μm, straight, cylindrical to conical, occasionally retaining a sterigmal appendage; suprahilar plage amyloid; immature spores sometimes with a non-functional apicular drop. Basidia 1–3(–4)-spored, 28–42 × 14–18 μm, clavate. Basidioles clavate, 20–24 × 10–16 μm. Macrocystidia 45–70 × 8–12 µm, lanceolate or fusiform, rostrate, narrowly acute or mucronate, easily collapsing. Paraphysoid cells absent. Sub-hymenium cellular, formed by subglobose cells of 8–16 µm diam. Hymenophoral trama 40–50µm wide, formed by septate hyphae 1.5–10 µm diam, with frequent enlargements up to 16 µm and abundant sphaerocytes 15–40 µm diam. Pileipel-lis and context 80–250 µm thick, thinning in the perimarginal zone; suprapellis a trichoderm of cylindrical dermatocystidia 15–70 × 3–8 µm, with tips of gelatinized hyphae soon col-lapsing into a brown, granular slimy mass; mediopellis 60–100 µm thick, that consists of an ixocutis of intricate hyphae 2–6 µm diam, with numerous gloeoplera 4–5 µm diam terminating in dermatocystidia; subpellis 50–120 µm thick, formed by an intricate cutis ofsubgelatinized hyphae 2–6 µm diam,connecting with the pileus and hymenophoral trama. Context of pileus and stipe-columella heteromerous. Stipitipellis a turf of repent to erect hyphal tips mixed with dermatocystidia. Thromboplera 3–5 µm diam, present in the trama and context. Habitat, Distribution & Season ― Solitary to gregarious, hypogeous to semi-hypogeous among plant debris, in montane broadleaved woods of Carpinus and Corylus mixed with Fagus, Populus, Quercus, Tilia, Betula, Acer, Fraxinus, Sambucus, Ulmus, on calcareous soil. Summer and autumn. Widely dis- ― Photos: a–c, f–g. J.M. Vidal; d–e. UdG. J.M. Vidal et al.: Revision of sequestrate Russulaceae tributed in temperate regions of Europe, from 100 m altitude in Northern Europe to 1100 m around the Mediterranean basin. Material studied. BULGARIA, Vraca, Chiren, under Tilia sp., 23 June 2018, M. Slavova & O. Ligabue (MSL2177F3265, duplicate BCN JMV800751). – CZECH REPUBLIC, Central Bohemia, Černošice, underCarpinus, Tilia and Acer, 9 July 1950, V. Vacek as E. krjukowensis (PRM 619106); South Moravia, Ždanice, near Bučovice, in forest, July 1941, F. Neuwirth (PRM 151648, holotype of O. moravica); Ždanice, near Zdravá Voda, in a leafy forest, July 1960, K. Kříž, det. M. Svrček as E. krjukowensis (PRM 620134). – FRANCE, Poitou-Charentes, Vienne, Couhé-Vérac, solitary under Carpinus, Oct. 1841, Tulasne (PC, herb. Tulasne, holotype of H. candidum). – GERMANY, Saarland, Gerlfangen, under Fagus, Carpinus, Quercus and Acer, 26 July 1968, G. Gross asʻE. mattiroloanusʼ (M, herb. G. Gross 171; duplicate BCN JMV799894)*; Honzrath, under Fagus, Carpinus and Corylus, on calcareous soil, 27 June 1968, G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 176). – HUNGARY, Nrád, Litke, under Carpinus, on calcareous soil, 20 Aug. 1996, Z. Lukács (BCN JMV800073); Tmez 250 m, under Carpinus, on calcareous soil, 1 Aug. 1996, Z. Lukács (BCN JMV800074). – ITALY, Lombardy, Monza e Brianza, Monza, Viale Reale, under Carpinus, 15 Oct. 1989, M. Sarasini as ʻE. mattiroloanusʼ (MS394, duplicate BCN JMV800003); Monza e Brianza, Monza, Viale Romano, under Carpinus, 15 Oct. 1992, M. Sarasini as ʻE. mattiroloanusʼ (MS467, duplicate BCN JMV800004); ibid., 19 Sept. 1994, M. Sarasini & G. Imperatori as ʻE. mattiroloanusʼ (MS619, duplicate BCN JMV800005). – RUSSIA, Moscow, Kryukovo, Kryukovo station, 18 July 1899, F. Bucholtz (FH, herb. F. Bucholtz, original material of S. krjukowense); Podolsk, Michailowskoje station, 20 July 1899, F. Bucholtz (FH, herb. F. Bucholtz 290, holotype of S. michailowskianum). – SPAIN, Cantabria, Saja, Reserva Natural del Saja, 658 m, under Fagus sylvatica, 1 Oct. 2017, J. Cabero (JC171001N)*; Catalonia, Girona, Moll Coll de la Boixeda, El Querol, 1100 m, under Corylus avellana, Quercus pubescens, Tilia platyphyllos and Populus tremula, on calcareous soil, 24 July 2010, M.A. Pérez-De-Gregorio (BCN JMV20100724b)*; ibid., 6 Sept. 2011, J.M. Vidal & F. Rodríguez (BCN JMV20110906-6b). Notes ― Tulasne & Tulasne (1843) collected in the French region of Vienne a fungus of angiocarpic appearance devoid of a stipe-columella, which they named Hydnangium candidum. Vidal (2004b) studied the holotype of H. candidum preserved in Tulasne’s herbarium at PC, as well as the holotype of Octaviania moravica preserved in Velenovsks herbarium at PRM and part of the original material of Secotium krjukowense and S. michailowskianum preserved in Bucholtz’s herbarium at FH, and concluded that all these taxa represent a single species, which he re-described as Macowanites candidus. Russula candida is characterized by its whitish angiocarpic basidiomata, which are maculated with yellowish orange, sometimes developing a minute stipe and a conspicuous columella. Spores are warty, subglobose to broadly elliptical, measuring 8.5–11(–12.5) × 7–9(–11) µm, many of them with a minute suprahilar plage in the base of hilar appendix and conserving a non-functional apicular drop, a character that is also present in other pseudo-angiocarpic species of the same clade (R. mattiroloana and R. mediterraneensis). This species is genetically nested within Russula sect. Russula, subsect. Maculatinae, and shows a significant relationship with R. maculata and R. nympharum, although it seems to represent an independent lineage not identified in previous studies of this clade (Adamčik et al. 2016). The complex around R. maculata was found by Adamčik et al. (2016) to be characterized by spores which are ornamented with low or medium-sized warts frequently connected and forming small ridges. However, R. candida has spiny spores not forming a reticulum or ridges, suggesting that this feature could be variable, or else restricted to the lineage formed by R. maculata and R. nympharum. Russula candidissima J.M. Vidal, Pasabán & Chachuła, sp. nov. ― MycoBank MB828502; Fig. 17 Basidiomata russuloid, angiocarpic to partially pseudoangiocarpic, stipitate. Pileus 1.4–2.8 cm wide, globose to lobate, subtomentose, pure white, later featuring pale yellow maculae; margin closed or laterally open, sublamel-late. Hymenophore loculate, pale yellow to yellow. Stipe-columella 1–2 × 0.2–0.4 cm, percurrent, white. Odour fruity. Spores 9.5–12.5 × 8.5–10.5 μm, subglobose to broadly ellipsoid, heterotropic, amyloid; warts 1.2–2 µm high, isolated or in groups of 2–4, commonly interconnected by low ridges. Basidia 2–4-spored, 37–55 × 14–18 μm, broadly clavate. Macrocystidia 45–65 × 6–14 µm. Hymenophoral trama with sphaerocytes measuring 20–36 µm diam. Suprapellis a cutis of repent hyphae 3–8 µm diam, and some dermatocystidia measuring 20–40 × 3–7 µm; pileal context a cutis with some sphaerocytes up to 15 µm diam. Hypogeous or semi-hypogeous between plant debris, in fir and broadleaved montane woods. Etymology. From Latin, candidissima = superlative for candidus (white), meaning very white because of its white colour more persistent than that of Russula candida. Holotype. SPAIN, Catalonia, Girona, Moll Coll de la Boixeda, El Querol, 1100 m, gregarious under Corylus avellana, Quercus pubescens, Tilia platyphyllos and Populus tremula, occurring in the same habitat as Russula candida, on calcareous soil, 6 Sept. 2011, J.M. Vidal & F. Rodríguez (BCN JMV20110906-6a)*. Basidiomata similar to Russula candida, russuloid, angiocarpic to partially hemiangiocarpic, stipitate. Pileus 1.4–2.8 cm wide, initially globose or subglobose, finally lobate, with a flattened to depressed apex, laterally open and sublamellate at the base, exposing the hymenophore; initially finely tomentose, then smooth, pure white for a long time, staining pale yellow with manipulation (4A4); greyish orange (5B5) to pale brown in ex-siccata (6D7). Hymenophore loculate, labyrinthoid, pale yellow to yellow (4A3–A6); pale brown (6D6) to brown in exsiccata (6E7); locules 0.7–1.8 × 0.15–0.3 mm (1–3 per mm), irregular, elongated, sinuous; fresh spore mass in locules pale yellow to yellow (4A3–A5); pale orange (5A4) to greyish orange (5B6) in exsiccata. Stipe-columella 1–2 × 0.2–0.4 cm, well developed, cylindrical, attenuate, central or eccentric, straight or curved, percurrent, not branched, pure white; context white, unchanging upon exposure to air. Odour fruity, taste mild. Spores 9.5–12.5(–15) × (8–)8.5–10.5(–13) μm, Q = 1.1–1.2 (–1.3), subglobose to broadly ellipsoid, heterotropic, echinate, yellow; warts amyloid, 1.2–2 μm high, cylindrical and straight, with a truncate apex, in groups of 2–4, interconnected with low inamyloid ridges forming a basal reticulum; hilar appendix 1–1.8 × 1.2 μm, conical or cylindrical, straight; suprahilar plage amyloid; immature spores with a non-functional apicular drop. Basidia (1–)2–4-spored, 37–55 × 14–18 μm, broadly clavate; sterigmata 4–5 µm long. Basidioles 30–35 × 10–15 μm, clavate. Macrocystidia scarce, 45–65(–100) × 6–14 μm, cylindrical, lanceolate or fusiform, rostrate. Paraphysoid cells 20–30 × 5–10 µm, 1-septate. Subhymenium composed of isodiametric cells 10–20 µm diam. Hymenophoral trama 30–50 μm wide, composed of hyaline hyphae 3–8 μm diam, with nests and col-umns of sphaerocytes 20–36 μm diam. Pileipellis and context 150–250 µm thick; pileipellis composed of: 1) a suprapellis soon collapsing into a yellow slimy mass, consisting of a cutis of upright to repent hyphae 2–4 µm diam, and some hyaline, cylindrical or clavate dermatocystidia 20–40 × 3–7 μm; and 2) a subpellis 100–130 μm thick, consisting of a subixocutis of densely intricate, hyaline hyphae 2–3 μm diam. Pileal context 100–110 µm thick, forming a cutis of intricate hyaline hyphae 2–4 µm diam, with some inflated hyphae and sphaerocytes up to 15 µm diam. Stipitipellis a turf of repent to erect hyphal tips mixed with dermatocystidia; context of stipe-columella het-eromerous, composed of densely interwoven narrow hyphae mixed with nests of sphaerocytes. Gloeoplera 2–5 μm diam, present in trama and context. Habitat, Distribution & Season ― Solitary to gregarious, hypogeous to semi-hypogeous among plant debris, in broadleaved montane woods of Carpinus, Corylus, Fagus, Quercus, accompanied by Acer, Betula, Castanea, Fraxinus, Populus, Tilia, but also under Abies, on calcareous soil, frequently found together with Russula candida. Summer and autumn. Widely distributed in temperate regions of Europe, from 100 m altitude in Northern countries to 1100 m in Southern countries. f. JMV800675. Basidiomata. ― g–i. JMV20100724a. g–h. Basidia, basidioles and macrocystidium; i. spores in Melzer. ― j. KRAF-2013-94. SEM image of spores. ― k. KRAF-2009-57. SEM image of sporal ornamentation. ― l. PRM 619108. Spores in Melzer. ― Scale bars: a, f = 1 cm; b, i, l = 10 µm; c, g–h = 20 µm; d–e, j = 5 µm; k = 1 µm ― Photos: a–c, g–i, l. J.M. Vidal; d–e. UdG; f. P.M. Pasabán; j–k. P. Mleczko, UJ. Additional material studied. AUSTRIA, Lower Austria, Wien-Neustadt, Näudorfl, 10 June 1937, M. Jacob, det. E. Soehner as ʻHydnangium cereumʼ (M, herb. E. Soehner 1894). – CZECH REPUBLIC, Central Bohemia, Černošice, in a leafy forest, under Carpinus, Crataegus and Cornus, 4 July 1950, V. Va-cek, det. M. Svrček as ʻElasmomyces krjukowensisʼ (PRM 619107); South Moravia, Shov, under Carpinus, Quercus and Fagus, 8 July 1953, K. Kříž, det. M. Svrček as ʻHydnangium krjukowense var. moravicumʼ (PRM 619108, herb. F. Šmarda). – GERMANY, Baden-Wttemberg, Altbach/Neckar, in a Carpinetum, 23 Oct. 1961, H. Steinnmann as ʻHydnangium krjukowenseʼ (PRM 616208); Saarland, Ballweiler, under Fagus, Carpinus and Quercus, on calcareous soil, 19 July 1968, G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 161); Bebelsheim, under Fagus, Carpinus and Quercus, on calcareous soil, 27 Aug. 1967, G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 82); ibid., 12 July 1968, G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 157); ibid., 26 Aug. 1969, G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 323); Eimersdorf, under Carpinus, Fagus and Quercus, on calcareous soil, 15 Aug 1968, Derbsch & G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 188); ibid., 6 Sept. 1969, J. Schmitt, det. G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 335); Eschringen, under Fagus, Quercus and Fraxinus, 3 Sept. 1967, G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 85); Gerlfangen, under Fagus, on calcareous soil, 20 July 1968, J. Schmitt, det. G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 164); ibid., under Fagus, Carpinus and Quercus, 20 July 1968, G. Korn, det. G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 164a); ibid., under Betula and Fagus, 10 Aug. 1968, G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 179); Sitterswald, under Fagus sylvatica, on calcareous soil, 4 July 1967, G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 79ʼ); ibid., 31 Aug. 1967, G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 84); ibid., 18 July 1968, G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 159); ibid., 10 July 1969, G. Gross as ʻE. mattiroloanusʼ (M, herb. G. Gross 303). – ITALY, Tuscany, Lucca, Vagli di Sotto, Monte Roggio, 900 m, under Corylus avellana, gregarious, 4 July 1993, G. Bernardini & L. Gori as ʻM. candidusʼ (ELG930704-1, duplicate BCN JMV800185)*. – POLAND, Lesser Poland, Krościenko nad Dunajcem, Pieniny Nat. Park, near Gródek, Western Carpathians, 480 m, under Abies alba, on calcareous soil, 19 Sept. 2009, P. Chachuła (KRA F-2009-57, duplicate BCN JMV800664)*; ibid., yellow route from Sromowce Niżne to Szopka pass, 640 m, under Abies alba, on calcareous soil, 13 July 2011, P. Chachuła (KRA F-2013-94, duplicate BCN JMV800665);Silesia,Cieszyn,Pogórze Śląskie region,Western Carpathians, 296 m, under Carpinus betulus and Quercus sp., on calcareous soil, 22 Aug. 2017, P. Chachuła (KRA F-2017-2, duplicate BCN JMV800669)*. – SPAIN, Basque Country, Guipcoa, Villabona, under Corylus avellana, Castanea sativa and Acer sp., on calcareous soil, 19 Sept. 2004, P.M. Pasabán as ʻM. candidusʼ (BCN JMV800674); Catalonia, Girona, Molló, Coll de la Boixe-da, El Querol, 1100 m, under Corylus avellana, Quercus pubescens, Tilia platyphyllos and Populus tremula, on calcareous soil, occurring in the same habitat as Russula candida, 17/24 July 2010, M.A. Pérez-De-Gregorio (BCN JMV20100724a); Navarre, Aizarotz, under Corylus avellana, on calcareous soil, 26 Sept. 2006, P.M. Pasabán as ʻM. candidusʼ (BCN JMV800675). Notes―After carefully studying all basidiomata present in collection JMV20100724 (Russula candida from Catalonia, Spain), some of them were found to have spores with intercon J.M. Vidal et al.: Revision of sequestrate Russulaceae nected warts forming low ridges, different from the spores with isolated warts found in the remaining specimens of R. candida. These basidiomata also had a well-developed stipe-columella. After revisiting all collections of Macowanites candidus studied by Vidal (2004b), similar spore ornamentation was found also in Czech, German and Italian collections. The same German collections had been previously described by Gross (1969), under the name of Elasmomyces mattiroloanus. Similar spore ornamentation was also depicted in SEM images obtained from Swiss specimens by Pegler & Young (1979). More recently, Calonge & Pasabán (2005), and Gori (2005), described and illustrated new collections (identified by them as M. candidus) found in Spain and Italy, respectively, with the same spore morphology. Russula candidissima can be discriminated macroscopically from R. candida because of its more persistently white external colour, its pileus that turns slightly yellowish with handling, and its mature basidiomata, which are opened in the base exposing the sublamellate hymenium, while in R. candida pileus becomes intensely yellow or yellowish orange with handling and its hymenium is rarely exposed. Microscopically, R. candidissima has a suprapellis composed of more or less parallel hyphae with scattered dermatocystidia and spore warts are interconnected by low ridges, while R. candida has a trichodermal suprapellis and spore warts are isolated. Russula mattiroloana differs microscopically from R. candida and R. candidissima by its bigger macrocystidia up to 130 µm long, and by its subglobose to globose spores up to 15.5(–18) × 15(–17) µm, ornamented with isolated warts up to 2.5–3 µm long. Genetically, R. candidissima is significantly related to several sequestrate species from North America such as R. xerophila, R. ellipsospora, R. mattsmithii (a replacement name for Gymnomyces compactus), R. stewartii (a replacement name for G. monosporus), and one from New Zealand, R. osphranticarpa (a replacement name for Gymnomyces redolens). All these sequestrate species constitute a significantly supported clade related to subsect. Firmiores, in concordance with previous results (Smith et al. 2006). Several species of this subsection are characterized by the yellow or yellowish orange colour of their lamellae and spores, a cuticular structure with short emergent hyphal tips, and spores ornamented with a spiny reticulum (Lebel 2002, 2003b, Smith et al. 2006). Russula cerea (Soehner) J.M. Vidal, comb. nov. ― MycoBank MB828503; Fig. 18 Basionym. Hydnangium cereum Soehner, Kryptog. Forsch. 1, 6: 394. 1924. Synonyms. Octaviania cerea (Soehner) Svrček in Pilát, Flora ČSR B1, Gasteromycetes: 197. 1958. Hydnangium carneum var. xanthosporum Hawker, Trans. Brit. Mycol. Soc. 35, 4: 281. 1952. (syn. nov.) Gymnomyces xanthosporus (Hawker) A.H. Sm., Mycologia 54: 635. 1962. Russula xanthospora (Hawker) Trappe & T.F. Elliott, Fungal Systematics and Evolution 1: 240. 2018. Lectotype of Hydnangium cereum (here designated MBT384232): GERMANY, Bavaria, ‘Erharting bei Mldorf’, under oaks, 16 Aug. 1921, E. Soehner (M, herb. E. Soehner 527); isolectotype in FH (herb. C.W. Dodge). Basidiomata 1–3 cm wide, angiocarpic, subglobose, slightly lobate, sessile, sometimes presenting an inconspicuous sterile base and attached to the substrate by a mycelial strand. Pileus dry, smooth, membranous, translucent, displaying hymenophoral locules, basally open and alveolate in old specimens; initially white, then greyish orange (5B6) to pale brown (6D6) when han- dled or upon exposure to the air,finally reddish brown (8E8). Hymenophore loculate, labyrinthoid, brown (7D7) to reddish brown (8D8); locules minute, 0.4–1.4 × 0.1–0.3 mm (2–3 per mm), elongated, sinuous; external locules always fertile; fresh spore mass in locules reddish brown (9D6); brown (6E7) in exsiccata. Columella absent or rudimentary, but thin sterile veins can be observed in some specimens. Odour fruity, taste mild. Spores (8–)9.5–12.5(–14) μm, Q = 1, variable in size, globose, orthotropic, echinate, orange to reddish in KOH; warts of irregu- lar length,1–3 μm high,intense yellow,deeply amyloid,dense, robust, conical or cylindrical, obtuse, longitudinally striate, isolated or fused in groups, often interconnected by low ridges; hilar appendix short, inconspicuous. Basidia typically 2-spored, 40–50 × 18–24 μm, pyriform to clavate, filled with minute, hyaline to dark reddish droplets, soon collapsing; sterigmata 8–10(–12) µm long, cylindrical to conical. Basidioles 22–36 × 10–13 µm, clavate. Macrocystidia abundant, (25–)30–50 × (5–)7–12(–16) μm, cylindrical to cylindro-clavate, wall up to 1 µm, with a dark yellow refractive content, rarely exceeding the basidia in length, resembling pseudocystidia because many of them originate at the deep subhymenium or hymenophoral trama. Paraphysoid cells abundant, 20–42 × 10–15 µm, vesiculose, 1-septate. Subhymenium very thick, pseudoparenchymatous, formed by 2–3 layers of prismatic to globose cells 8–20(–25) µm diam. Hymenophoral trama reduced, 10 µm thick, formed by hyphae 3–6 µm diam, and some sphaerocytes 10–20 µm diam, only present in tramal anastomoses. Pileipellis and context 70–125 µm thick; pileipellis thin, 50–100 µm, consisting of: 1) a trichodermal suprapellis of subulate to lageniform, capitulate dermatocystidia, measuring 15–30 × 3–5 µm, that soon collapses into a yellowish mass; and 2) a prosenchymatous subpellis 20–30 µm thick,formed by a dense mesh of interwoven, subgelatinized yellowish hyphae, measuring only 2–2.5 µm diam. Peridial context formed of interwoven hyaline hyphae 2–5 µm diam, lacking sphaerocytes. Gloeoplera and thromboplera present in the trama and context. Habitat, Distribution & Season ― Solitary to gregarious, hypogeous, in montane conifer (Abies, Picea, Pinus) or broadleaved (Carpinus, Castanea, Corylus, Fagus, Quercus) forests, commonly on siliceous soil. Summer and autumn. Widely dis-tributed in temperate and submediterranean regions of Europe, from almost sea level in Northern countries to 1100 m altitude in Southern countries. Additional material studied. GERMANY, Saxony, Chemnitz, Park der OdF, 300 m, under Castanea sativa, 20 June 2009, B. Mler & G. Hensel (GH20090620, duplicate BCN JMV800660)*. – HUNGARY, Veszprém, Zirc, under Fagus and Carpinus, 15 Sept. 2000, Z. Lukács (BCN JMV800164). – ITALY, Emilia-Romagna, Forli-Cesena, Santa Sofia, Foresta di Campigna, under Abies alba, 11 Oct. 1991, M. Sarasini as G. xanthosporus (MS419, duplicate BCN JMV800006); Tuscany, Lucca, Vagli di Sotto, Monte Roggio, 800 m, under Corylus avellana, 4 Dec. 1994, G. Bernardini & L. Gori as G. xanthosporus (ELG941204-3, duplicate BCN JMV800172). – POLAND, Lower Silesia, Góry Złote region, Sudety mountains, under Quercus sp., Fagus sylvatica, Picea abies, Abies alba and Acer sp., 9 Aug. 2012, M. Kozak & P. Mleczko (KRA F-2012-28, duplicate BCN JMV800667)*. – PORTUGAL, Aveiro, Luso, bosque de Buçaco, 547 m, under Castanea sativa, on siliceous soil, 11 Nov. 2015, A. Paz (BCN IC11111504). – SPAIN, Asturias, Perán, under Castanea sativa, 17 Aug. 2000, F. García (BCN JMV20000817-1)*; Catalonia, Girona, Campelles, Torrent de Prat de Jou, 1100 m, under Corylus avellana, Fraxinus, Acer and Buxus, on siliceous soil, 3 Oct. 1996, J.M. Vidal as G. xanthosporus (BCN JMV961003-8)*; Girona, Espinelves, crossroad to Viladrau, 740 m, in a Picea abies plantation, on siliceous soil, 5 July 2016, F. Rodríguez (BCN JMV20160705)*; ibid., 4 Sept. 2018, F. Rodríguez & J.M. Vidal (BCN JMV20180904-3); Navarre, Eratsun, under Pinus pinaster, 10 June 2015, P.M. Pasabán & F. Sáinz (BCN JMV800656)*. – UNITED KINGDOM, Wales, Caernarvon, ‘Bettws-y-Coed, North Wales, inter acubus emortius Piceae’, 13 Sept. 1951, L.E. Hawker (K(M)69329, herb. L. Hawker H251, holotype of H. carneum var. xanthosporum). Notes ― Soehner (1924) described an angiocarpic fungus found under oaks in the German region of Bavaria, which he called Hydnangium cereum, but did not select a type collection. In a later paper, Soehner (1941) provided new data and reference e. JMV20160705. Basidia and macrocystidia. ― f. ELG941204-3. Spores in Melzer. ― g–h. M, herb. E. Soehner 527 (original material of Hydnangium cereum). g. Basidium and macrocystidia; h. spores in KOH 5 %. ― i. JMV20160705. SEM image of spores. ― j–k. K(M)69329 (holotype of Hydnangium carneum var. xanthosporum). j. Basidia and macrocystidia; k. spores in KOH 5 %. ― l. KRAF-2012-28. SEM image of spores. ― Scale bars: a, d = 1 cm; b–c, e, g, j = 20 µm; f, h, k = 10 µm; i, l = 5 µm. ― Photos: a. G. Hensel; b–h, j–k. J.M. Vidal; i. UdG; l. P. Mleczko, UJ. codes for several collections of this species, indicating that duplicates of a collection labelled ‘Summer 1921’ were sent to several mycologists such as E. Fischer, O. Mattirolo and C.W Dodge. Dodge & Zeller (1937) considered this collection, which they labelled ʻE. Soehner 527ʼ, as the type of H. cereum. After studying several original collections of H. cereum from E. Soehner’s herbarium at Münich, we found that ʻSoehner no. 527ʼ is the only collection consisting of mature basidiomata, so it is here designated as lectotype. Later on, Hawker (1952) described from Wales, UK, an angiocarpic fungus of similar characteristics that she called Hydnangium carneum var. xanthosporum. Type collections of H. cereum and H. carneum var. xanthosporum were compared in the present work, and found to be probably conspecific because of their bisporic basidia, cylindrical thick-walled macrocystidia, and conspicuously ornamented globose spores, with H. cereum being the prioritary name. Interestingly, the presence of hyme-nial macrocystidia was not observed by Soehner (1924, 1941) in H. cereum, neither by Hawker (1952, 1954), Smith (1962), nor by Pegler et al. (1993) in H. carneum var. xanthosporum, despite them being very conspicuous. Morphological traits as well as genetic data from specimens originating from Central and Southern Europe suggest that this taxon is another se-questrate species of Russula, and therefore the combination R. cerea is here proposed. Phylogenetic inferences suggested that R. cerea belongs to sect. Ingratae, and probably subsect. Foetentinae. Russula cerea is significantly related to other European sequestrate taxa such as R. pila and R. mistiformis, being macroscopically very similar to both species, but differing because of its thinner membranous pileus, bisporic basidia and presence of thick walled, clavate macrocystidia. All samples of R. cerea seemed genetically similar to each-other, except for JMV20000817-1 from Asturias (Spain), which showed significant differences in ITS (15/564 bp = 2.65 %, including a 7 bp insertion) and 28S rDNA (4/851 bp = 0.5 %). After microscopically studying this specimen, we have been unable to find any distinct features, so it could represent a cryptic species or maybe a hybrid specimen with either R. pila or R. mistiformis. J.M. Vidal et al.: Revision of sequestrate Russulaceae a–d. J.M. Vidal; e–g. UdG. Russula galileensis (M.M. Moser et al.) Trappe & T.F. Elliott, Fungal Systematics and Evolution 1: 234. 2018 ― Fig. 19 Basionym. Macowanites galileensis M.M. Moser et al., Trans. Brit. Mycol. Soc. 68, 3: 371. 1977. Basidiomata russuloid, angiocarpic, stipitate. Pileus 2–6 cm wide, irregular and tuberous, sometimes depressed at the apex, white with cream to pale umber maculae; margin radially alveolate when mature, but not open. Hymenophore loculate, cream-ochre. Stipe-columella 0.5–3.5 × 0.5–1.7(–3) cm, white. Odour and taste mild. Spores (9–)10–11.5(–15) × (7.5–)9.5–10.5(–14) µm, Q = 1.05– 1.15, subglobose, heterotropic; reticulum 0.5 µm high, amyloid, composed of crests and warts. Basidia 4-spored, 26–37 × 11–13 µm, broadly clavate. Macrocystidia 40–60 × 10–15 µm, sometimes 1–2-septate. Hymenophoral trama heteromerous. Pileipellis and context 200–250 µm thick; suprapellis an intricate trichoderm of repent, to semi-erect hyphae and dermato cystidia 2–7 µm diam. Habitat, Distribution & Season ― Subepigeous under fallen leaves, in sclerophyllous woods of Quercus, on calcareous soil. From late autumn to spring. So far located only in the eastern Mediterranean region, in Israel, found at low altitudes. Material studied. ISRAEL, Carmel Mount, Mt Carmel National Park, near Haifa, Shajarat Al Arba’in (Horshat Ha’arbaim), under Quercus calliprinos, 12 Dec. 1972, M. Moser, N. Binyamini & Z. Avizohar-Hershenzon 72/401 (IB M72340, isotype of M. galileensis); ibid., near Pinus, 13 Jan. 2012, M. Krakhmalnyi as M. galileensis (HAI-G-91); Haifa, Technion, Dec. 2011, M. Krakhmalnyi as M. galileensis (HAI-G-59); Upper Galilee, Goren Park, Dec. 2011, Z. Shafranov, det. M. Krakhmalnyi as M. galileensis (HAI-G-83; duplicate BCN JMV800634)*; ibid., Hanita Forest, 22 Dec. 2012, Z. Shafranov, det. M.Krakhmalnyi as M. galileensis (HAI-G-166)*; Mount Meron National Park, 7 Jan. 2006, I. Shams, det. S.P. Wasser as M. galileensis (HAI-G-202)*; Nahal Kziv Nature Reserve, ‘Evolution Canyon-II’, under Quercus, 3 Jan. 2001, S. Reshetnikov, det. S.P. Wasser as M. galileensis (HAI-G-45); ibid., under Quercus, 3 Jan. 2001, T. Pavlichek & I. Shams, det. S.P. Wasser as M. galileensis (HAI-G-201; duplicate BCN JMV800635)*; ibid., under Laurus, 7 Jan. 2001, I. Duckman, det. S.P. Wasser as M. galileensis (HAI-G-46); Nahal Zalmon National Park, under Quercus, 1 Febr. 2003, Y. Ur, det. S.P. Wasser as M. galileensis (HAI-G-209)*. Notes ― Macowanites galileensis was originally described and illustrated from the specimens collected on Mt Carmel in 1972 (Moser et al. 1977, 1994). Interestingly, the authors initially regarded their collection to be M. krjukowensis, due to white colour of basidiomata. Following more careful examination, however, it became evident that new collections possessed a number of distinctive macro- and micromorphological characters, such as a very robust pileus and stipe-columella, as well as warts of spores connected by ridges and crests, features which separated them from the latter. We could not obtain im-ages to illustrate the studied collections. Recently, this species was re-combined into genus Russula by Elliott & Trappe (2018). Current molecular results indicate that R. galileensis belongs to Russula subsect. Laricinae, being significantly related to the R. laricinoaffinis clade. Subsection Laricinae is characterized by ocher to yellowish spore deposits, short basidia, rare pleuro-cystidia and multiseptate dermatocystidia (Romagnesi 1985, Li et al. 2013). Other European sequestrate taxa also related to Laricinae include R. vinaceodora and R. vidalii (a replacement name for Gymnomyces ilicis), both species apparently endemic to the Mediterranean region (Moser et al. 1977, Llistosella & Vidal 1995, Calonge & Vidal 2001, Cabero 2011). Russula galileensis and R. vinaceodora produce large basidiomata (up to 5–6 cm across), with a conspicuous stipe-columella that is sublamellate in the lower part, and spores with warts forming small interconnecting ridges, while R. vidalii is a smaller fungus (1–2.5 cm diam), with a poorly developed columella, a loculate hymenophore, and spores which are ornamented with more or less isolated warts. Russula hobartiae Loizides & J.M. Vidal, sp. nov. ― Myco-Bank MB828504; Fig. 20 Basidiomata 2–7 cm wide, angiocarpic, turbinate to pyriform, weakly lobate, tomentose, whitish to ochraceous brown, firmly rooted into the substrate. Hymenophore loculate, whitish to ochraceous orange, becoming vinaceous in FeSO4; locules minute, rounded to elongate. Columella absent or rudimentary, sometimes forming short white veins. Odour slightly fruity, with a vague hint of chocolate. Spores 7–9.5(–10.5) × 7–9(–10) μm,globose;reticulum complete to incomplete, up to 0.5 µm high. Basidia 2–4-spored, 20–35 × 8–12 µm, clavate. Macrocystidia 25–40 × 8–10 µm, clavate, scarce. Hymenophoral trama prosenchymatous, with scarce sphaerocytes. Suprapellis a palisado-trichoderm of clavate dermatocystidia 25–60 × 5–10 µm; pileal context a cutis of hyaline hyphae 1.5–3 µm diam, lacking sphaerocytes. Hypogeous or semi-hypogeous in montane woods of Pinus nigra subsp. pallasiana on the island of Cyprus. Etymology. hobartiae = in honor of mycologist Caroline Hobart, for her contribution to the study of hypogeous and sequestrate fungi in Cyprus. Holotype. CYPRUS, Nicosia District, Prodromos, 1450 m, under Pinus nigra subsp. pallasiana, 13 Oct. 2011, M. Loizides (BCN JMV800628)*; isotype in herb. pers. M. Loizides (ML110131GY). Basidiomata 2–7 cm wide, angiocarpic, turbinate, oblate-globose to pyriform and somewhat applanate at the apex, weakly lobate, distinctly tapering at base, longitudinally furrowed and firmly rooted into the substrate. Pileus dry, finely tomentose under the lens, often cracked, dirty white, cream-white (5A2–A4) or ochraceous cream (6A2), usually with darker ochraceous or brownish stains (6C5). Hymenophore loculate, labyrinthoid, cream at first, at maturity becoming deep ochraceous yellow to ochraceous orange (5A6); instantly turning vinaceous in FeSO4; locules minute, 0.2–1 × 0.1–0.3 mm (3–5 per mm), irregularly rounded to elongated, empty or filled; fresh spore mass in locules pale yellow to yellow (4A3–A4); pale orange (5A4–A5) in exsiccata. Columella absent or rudimentary, sometimes forming a sterile base or short white veins, but small sterile patches often scattered throughout the hymenophore. Odour slightly fruity, sometimes with a vague hint of chocolate when cut; taste astringent. Spores 7–9.5(–10.5) × 7–9(–10) μm, Q = 1.0–1.1, globose, orthotropic, reticulated; reticulum complete to incomplete, up to 0.5 μm high, formed by strongly amyloid crests and warts; hilar appendix usually rudimentary, exceptionally up to 2–3 μm long. Basidia 2–4-spored, 20–35 × 8–12 µm, clavate, thick-walled; sterigmata 4–6 μm long. Macrocystidia 25–40 × 8–10 µm, clavate, scarce, very scattered and immersed in the hymenium. Paraphysoid cells 11–34 × 7–12 μm, cylindrical to slightly clavate, 1–2-septate. Subhymenium ramose, composed of chains of 1–2 globose to irregularly polygonal or elongated cells 5–15 µm diam. Hymenophoral trama 40–70μm wide, prosenchymatous, formed by septate, tortuous, interwoven hyaline hyphae 3–6(–8) μm diam, with some inflated elements up to 12 µm diam and scattered sphaerocytes or nests of sphaerocytes 6–20 μm diam, especially in tramal anastomoses. Pileipellis and context 120–250(–400) µm thick; pileipellis consisting of: 1) a palisadotrichodermal suprapellis of yellow, clavate der-matocystidia 25–60 × 5–10 µm; and 2) a prosenchymatous subpellis 30–40 µm thick of branched, densely interwoven, septate hyphae 3–6 μm diam.Pileal context 80–200 µm thick, arranged in a cutis of septate, hyaline hyphae 1.5–3 µm diam; sphaerocytes absent. Gloeoplera abundant, up to 8 µm diam. Habitat, Distribution & Season ― Solitary to gregarious, hypogeous to semi-hypogeous, in montane woods of Pinus nigra subsp. pallasiana, rarely also with Pinus brutia, on siliceous, well-drained serpentine soils. Summer and autumn. Xerophytic. Known from the eastern Mediterranean region, where it is pro-bably endemic to the island of Cyprus, between 1100–1800 m altitude. Additional material studied. CYPRUS, Limassol District, Troodos, 1800 m, under Pinus nigra subsp. pallasiana, 7 Oct. 2008, M. Loizides (ML8017GY); ibid., 1700 m, 28 Sept. 2009, M. Loizides (ML90982GY); ibid., 1600 m, 4 Nov. 2012, M. Loizides (ML21114GY); ibid., 26 Aug. 2014, M. Loizides (ML41862GY); ibid., 1650 m, 3 Sept. 2014, M. Loizides (ML4193GY)*; ibid., 1700 m, 20 Sept. 2016, M. Loizides (ML61902GY); Nicosia District, Prodro-mos, 1450 m, under Pinus nigra subsp. pallasiana, 7 Oct. 2011, M. Loizides (ML11701GY); Platania, 1100 m, under Pinus brutia, 16 Nov. 2011, M. Tordelli (ML111161GY); ibid., 1120 m, 20 Nov. 2011, G. Konstantinidis & D. Klisiari (GK5889, duplicate BCN JMV800647)*; ibid., 16 Nov. 2014, M. Loizides (ML411161GY)*. Notes ― Russula hobartiae is characterized by the typically turbinate shape of its basidiomata, with a flattened upper side and a conical base that is deeply rooted into the substrate and breaks easily. Hymenial chambers are very small, only visible under a lens, cream at first but turning into an intensely orange colour when mature. Microscopically, it is characterised by its d–i. ML110131GY (BCN JMV800628, holotype). d. Pileipellis; e. hymenium, subhymenium and hymenophoral trama; f. hymenium and subhymenium; g. spores in Melzer; h–i. SEM images of spores. ― Scale bars: a–c = 1 cm; d–f = 20 µm; g = 10 µm; h–i = 5 µm. ― Photos: a–c. M. Loizides; d–g. J.M. Vidal; h–i. UdG. J.M. Vidal et al.: Revision of sequestrate Russulaceae globose or subglobose spores ornamented with a reticulum, its infrequent and very scattered clavate macrocystidia, not exceeding the hymenial layer, and the pileipellis formed by abundant clavate dermatocystidia forming a palisade. Genetically, R. hobartiae issignificantlyrelated to a North American specimen identified as Russula cf. ochrophylla (BPL231, Looney et al. 2016). Currently, there are no other records of R. ochrophylla in public databases, so the identity of this sample cannot be confirmed until addtional collections are studied. Rus-sula ochrophylla is characterized by a purple pileus, ochraceous yellow lamellae, mild taste, echinulate spores, and preference for broadleaved trees (Peck 1897, Adamčik et al. 2017), features which are remarkably diverse from those of R. hobartiae. This morphological disparity could perhaps be explained if: 1) specimen BPL231 does not actually represent R. ochrophylla; 2) the features of one of both concerned species are not typical of their lineage; or 3) R. ochrophylla and R. hobartiae actually belong to different subsections. Russula mattiroloana (Cavara) T. Lebel, Muelleria 36: 11. 2017 ― Fig. 21 Basionym. Elasmomyces mattiroloanus Cavara, Malpighia 11, 1–3: 426. 1897 (ʻmattirolianusʼ). Synonym. Macowanites mattiroloanus (Cavara) T. Lebel & Trappe, Myco-logia 92, 6: 1194. 2000. Basidiomata russuloid, pseudoangiocarpic, stipitate. Pileus 1.2–3.8 cm wide, subglobose to hemisphaerical, open in the base and exposing a marginally sublamellate hymenophore; initially finely tomentose, then smooth, whitish, pale yellow (4A4) to pale orange (5A4), maculated with brownish orange (6C8), finally brownish red (8D7) maculated with dark reddish brown (8E7–F6); greyish orange (5B5) to brown in exsiccata (7E8). Hymenophore loculate, labyrinthoid-daedaleoid, initially with an enclosed stipe-columella, soon expanding outwards and exposing the sublamellate hymenium underneath; initially whitish, yellowish white to pale yellow (4A2–A3), then pale yellow (4A5) to pale orange (5A4), finally titian red (7D6); pale orange (5A4) to dark brown in exsiccata (6F7); locules large, 1.5–3 × 0.2–0.8 mm (1–2 per mm), radially arranged, c. KRAF-2018-1. Basidiomata. ― d–l. KRAF-2012-153. d. Pileipellis; e. stipitipellis; f. hymenium and subhymenium; g–h. spores in Melzer; i–l. SEM images of spores. i. two ellipsoid young spores; j. globose mature spore; k. detached apicular drop over the ornamental warts; l. non-functional apicular drop united to the base of hilar appendix. ― Scale bars: b–c = 1 cm; d–f = 20 µm; g–h = 10 µm; i–k = 5 µm; l = 1 µm. ― Photos: b. G. Konstantinidis; c. P. Chachuła; d–h. J.M. Vidal; i–l. P. Mleczko, UJ. elongated, irregular, sinuous; fresh spore mass in locules pale yellow (4A4); pale orange (5A3) to brown (6E6) in exsiccata. Stipe-columella 0.8–2.5 × 0.3–1 cm, well developed, percur-rent, stipe 2–3 times longer than the columella, cylindrical, generally central, often curved, pruinose, pure white; context white, not changing upon exposure to air, but yellow in the cortical zone. Odour and taste mild. Spores (9–)10.5–15.5(–18) × (8–)10–15(–17) μm, Q = 1.0– 1.15, very variable in size, globose to subglobose, some ellipsoid when immature (Q = 1.1–1.2), heterotropic to subhetero tropic, echinate; warts (0.7–)1.5–2.5(–3) µm high, isolated, amyloid, conical and straight, with an acute apex; hilar appendix 1.5–2 × 1–1.5 μm, cylindrical, straight, with a minute, inconspicuous, amyloid suprahilar plage in its base and conserving a non-functional apicular drop. Basidia 2–4-spored, (25–)45–65 × 17–23 µm, ventricose to broadly clavate; sterigmata 3–5 µm long. Macrocystidia abundant, 70–125(–180) × 15–18(–25) µm, cylindrical, lanceolate or fusiform, obtuse or rostrate, with a wall 1 µm thick. Paraphysoid cells scarce, 18–28 × 6–8 µm, cylindrical or clavate, aseptate or 1-septate. Subhymenium of pseudoparenchymatous aspect, consisting of 3–4 strata of globose to prismatic cells 14–30 µm diam. Hymenophoral trama 20–30 μm wide, with scarce sphaerocytes 15–35 µm diam, more abundant in tramal anastomoses. Pileipellis and context 100–300 µm thick; pileipellis composed of: 1) a tricho dermal suprapellis consisting of upright to repent hyphae 2–6.5 µm diam, and abundant, long, cylindrical or clavate, usually 1-septate, yellow dermatocystidia, 60–140 × 3–8 μm; and 2) a subpellis 100–180 μm thick, comprised of a subixocutis of intricate, hyaline hyphae 1.5–5 μm diam, with frequent en-largements 10–20 µm at hyphal ramifications. Pileal context heteromerous, 200–1000 µm thick, formed by hyaline hyphae 3–6 µm diam, and abundant nests of sphaerocytes up to 40 µm diam. Stipitipellis a turf of repent to erect, septate hyphae, 3–5 µm diam, and clavate dermatocystidia, 50–100 × 5–12 µm, similar in shape and content to those of pileipellis; con-text of stipe-columella heteromerous, composed of densely interwoven narrow hyphae mixed with nests of sphaerocytes. Gloeoplera 2–5 μm diam, present in trama and context where they terminate as dermatocystidia. Habitat, Distribution & Season ― Solitary to gregarious, subepigeous among needles, in forests of Abies and Picea, on siliceous or calcareous soils. Summer and autumn. Distributed in temperate and submediterranean regions, from Central to Southern Europe (Carpathian, Pindus and Apennine mountains), between 500–1600 m altitude. Material studied. GREECE, Thessaly, Trikala, Pertouli, Koziakas mountain, Pindus mountains, 1180 m, under Abies borisii-regis, 5 July 2009, G. Konstantinidis as M. mattiroloanus (GK3901, duplicate BCN JMV800638)*; ibid., 1200 m, under Abies borisii-regis, 11 July 2015, G. Konstantinidis as M. mat-tiroloanus (GK8136, duplicate BCN JMV800644)*. – ITALY, Emilia-Romagna, Reggio Emilia, Civago, Abetina Reale Forest, Tuscan-Emilian Apennines, 1500 m, under Abies alba and Fagus sylvatica, on siliceous soil, 5 Oct. 1983, A. Montecchi asʻM. ellipsosporaʼ (AM1616, duplicate BCN JMV800245); ibid., 1400 m, 10 Aug. 1985, A. Montecchi as E. mattiroloanus (AM65, duplicate BCN JMV800241); ibid., 16 Aug. 1985, A. Montecchi as E. mattiroloanus (AM64, duplicate BCN JMV800240); ibid., 1600 m, under Abies alba and Fagus sylvatica, on siliceous soil, 5 Aug. 1988, A. Montecchi as E. mattiroloanus (AM825, duplicate BCN JMV800038); Tuscany, Firenze, Reggello, Vallombrosa Forest, Tuscan-Emilian Apennines, ‘ad terram sub acubus Abietis pectinatae prope S. Giovanni Gualberto Vallisumbrosae (Florentia), F. Cavara’ (FH, herb. C.W. Dodge 2087; NY, herb. S.M. Zeller 1671; all la-belled ‘Arcangeliella borziana, Italy, Etruria, Vallombrosa, forest of firs, autumn 1898, Instituto Botanico di Napoli, coll. F. Cavara, type’; isotypes of E. mat-tiroloanus); Tuscany, 24 Nov. 1899, O. Mattirolo as E. mattiroloanus, conf. F. Cavara (FH, herb. F. Bucholtz 275). – POLAND, Lesser Poland, Pieniny Nat. Park, Western Carpathians, Sokolica massif, close to Hukowa Skała rock, 490 m, under Abies alba, on calcareous soil, 13 July 2012, P. Chachuła (KRA F-2012-153, duplicate BCN JMV800666)*; ibid., Bajków Groń, 710 m, under Abies alba, on calcareous soil, 3 July 2018, P. Chachuła (KRA F-2018-1, duplicate BCN JMV800713)*; ibid., Lasek, 660 m, under Abies alba, on calcareous soil, 21 July 2018, P. Chachuła (KRA F-2018-3, duplicate BCN JMV800715); ibid., Ula mountain, 620 m, under Abies alba, on calcareous soil, 8 Nov. 2017, P. Chachuła (KRA F-2017-1, duplicate BCN JMV800670)*; ibid., 8 July 2018, P. Chachuła (KRA F-2018-2, duplicate BCN-JMV800714)*. Notes ― Elasmomyces mattiroloanus was collected under Abies alba at the beginning of September 1896 in Tuscany (Italy) by Cavara (1897), who reported it as a russuloid fungus with a pseudoangiocarpic habit, characterized by its globose spores, 14–15 µm diam, and abundant macrocystidia, 70–72 µm long. Singer & Smith (1960) observed microscopical differences between the type material kept at New York Botanical Garden (NY) and the descriptions provided by Cavara: spores were subglobose or ellipsoid, measuring 10–15 × 9.5–13.5 µm, lacked macrocystidia and showed cystidioid ʻsterile bodiesʼ, measuring 22–27 × 11–12 µm. Lebel & Trappe (2000) studied the same material at NY and confirmed the subglobose to broadly ellipsoid spore shape (although they report different spore measures, 10–12 × 9–11.5µm), andfurther discriminated between two different kinds of cystidia: ventricose to clavate, 28–37 × 7–11 µm, and clavate to cylindrical, 8–14 × 3–5 µm. Vidal (2004b) also studied the same type material kept at S.M. Zeller’s herbarium in NY (without voucher number) and concluded that microscopical features of these specimens do not match those reported in the protologue of E. mattiroloanus, but are compatible with those of Arcangeliella borziana, concluding that both species were probably erroneously labelled. The original concept of E. mattiroloanus has spores of unequal size and shape, 10.5–15.5(–18) × 10–15(–17) µm, typically globose or subglobose, sometimes ellipsoid when immature, resembling those of Martellia ellipsospora sensu Montecchi & Sarasini (2000), and numerous macrocystidia measuring 70–125(–180) × 15–18 µm. In mature spores, it is interesting to observe the intensely amyloid, non-functional apicular drop, which remains located in a basal position in the hilar appendix, immediately above a minute and residual suprahilar plage, which is not always visible, and finally becomes detached on top of orna-mental warts, or fused with the suprahilar plage, also observed in spores of R. mediterraneensis. Recently, E. mattiroloanus was re-combined into genus Russula by Lebel (2017). Genetically, R. mattiroloana belongs to the R. globispora com-plex of subsect. Maculatinae (Adamčik et al. 2016). Despite ITS rDNA sequences of R. mattiroloana sharing a 20 bp deletion not present in any other species of Maculatinae, its clade collapsed or received non-significantsupport(Fig.5),evidencing the need of additional samples and genetic markers. Subsection Macu-latinae was originally characterized by species with reddish pilei, acrid taste and a yellow spore print (Romagnesi 1985), but these features are also shared by species not directly related (Adamčik et al. 2016). Interestingly, Sarnari (1998) subsumed Maculatinae within subsect. Urentes, and highlighted the presence in this group of an amyloid suprahilar spot in spores. The clade formed by R. globispora and R. dryadicola was reported by Adamčik et al. (2016) to be characterized by spores orna-mented with large isolated spines, a feature also observed in R. mattiroloana, although also present in species not directly related with this clade, such as R. candida. Russula mattiroloa-na seems to be present in Abies sp. and Picea sp. forests of the alpine Mediterranean regions (Cavara 1897, Montecchi & Lazzari 1986, Montecchi & Sarasini 2000, Konstantinidis 2009) and Central Europe. Russula globispora has large, globose spores (Bon 1986, Llistosella 1998, Lejeune 2005) similar to those of R. mattiroloana, but spores of R. dryadicola and the new species R. heilongjiangensis (Li et al. 2018) do not share these features, suggesting that they are not representative of the entire genetic lineage. J.M. Vidal et al.: Revision of sequestrate Russulaceae Russula mediterraneensis Konstantin., J.M. Vidal, Gelardi, Papadimitriou, Tulli, Angeli & Vizzini, sp. nov. ― MycoBank MB828505; Fig. 22 Basidiomata russuloid, pseudoangiocarpic, stipitate. Pileus 2.3–4.3(–5.2) cm wide, hemispherical to spherical, sometimes lobate, smooth, whitish to yellowish with brown maculae. Hymenophore internally daedaleoid and externally lamellate, cream to ochraceous orange. Stipe-columella 1.6–4.2 × 0.7–1.8 cm, well developed, percurrent, white. Odour of pears or yoghourt and fruit. Spores (8–)10–15(–17) × (7–)9–14(–15) μm, subglobose to broadly ellipsoid, heterotropic; warts 0.6–1(–1.5) μm high, isolated or in few groups. Basidia 2–3-spored, 30–55 × 11–18 μm, clavate. Macrocystidia 70–130 × 12–20 μm. Hymenophoral trama heteromerous. Suprapellis a trichoderm of erected hyphae 2–6 µm diam, and clavate, septate dermatocystidia 30–160 × 5–10 µm; pileal context heteromerous with nests and columns of sphaerocytes 21–47 µm diam. Found in Greece and Italy, subepigeous under Quercus and Castanea. Etymology. From Latin, Mediterraneum (mare) = Mediterranean Sea, and -ensis = from, found in, for its occurrence in Mediterranean localities. Holotype. GREECE, West Macedonia, Kozani, Vouchorina, 700 m, under Quercus robur, 23 July 2013, G. Konstantinidis (BCN JMV800641)*; isotype in herb. pers. G. Konstantinidis (GK6710). Basidiomata russuloid, small to medium sized, pseudoangiocar-pic, stipitate. Pileus 2.3–4.3(–5.2) cm wide, initially globose to subglobose, sometimes lobate, almost completely closed and covering the whole stipe except its base, later convex to plane, opened basally and exposing a lamellate hymenium, with an involute, dentate margin; rugulose, finely tomentose, white or whitish at first, sometimes with translucent spots, especially at the periphery and margin, later yellowish with ochraceous areas (4A2–A3) and finally mostly bicoloured, yellowish buff (4A4) to ochraceous buff (5A4), with brown maculae (6C8–7D8). Hymenophore irregularly loculate inside, daedaleoid, lamellate in the most external parts; locules large, 2–5 × 0.2–1 mm (0.5–1 per mm), labyrinthoid or flexuous in cross-section; lamellae irregular, more or less sinuous, frequently anastomosing and bifurcate, adnate, initially white, later cream, yellowish to ochra- ceous (4A3–A7), and finally ochre-orange or ochre-buff (6C8), with concolourous, acute or obtuse edges; white remnants of a partial veil can be observed in lamellae edges of some young basidiomata; fresh spore mass in locules yellowish white (4A2). Stipe-columella 1.6–4.2 × 0.7–1.8 cm, percurrent, well developed, cylindrical, straight or rarely flexuous, central or eccentric, white, initially solid and later medullary, usually broadening but rarely also tapering at the base; context white, not changing after contact with air. Odour pleasantly fruity, of pear or remi-niscent of yoghourt and fruit; taste mild to slightly acrid. Spores (8–)10–15(–17) × (7–)9–14(–15) μm, Q = (1.0–)1.10– 1.25(–1.3), very variable in form and size, globose, subglobose, broadly ellipsoid to ovoid, heterotropic, echinate, some with one large, central oil drop; warts 0.6–1(–1.5) μm high, amy loid, irregularly conical or truncated, scattered or packed in few groups, with some verrucae among them; hilar appendix 1.2–1.8 µm long, cylindrical, with a distinct amyloid suprahilar plage in its base and conserving a non-functional apicular drop as in R. mattiroloana. Basidia (1–)2–3(–4)-spored, 30–55 × 11–18 μm, clavate, some of them with one or more oil droplets and amorphous yellowish content; sterigmata (4–)6.5–7.5 µm long. Basidioles 30–35 × 10–15 μm, clavate. Macrocystidia abundant (40–)70–130(–180) × 12–20 μm, subcylindrical, lanceolate to fusiform, thin-walled or slightly thick-walled in the central portion, 1–1.6 μm thick, with abundant amorphous yellowish content, and an obtuse to acute or subulate apex, sometimes rostrate. Paraphysoid cells scarce, 12–26 × 6–9 µm, cylindrical, entire or 1-septate. Subhymenium cellular, d. daedaleoid hymenophore; e. pileipellis; f. hymenium and subhymenium; g. spores in Melzer; h–i. SEM images of spores. ― j. GK7286. Spores in Melzer. ― Scale bars: a–b = 1 cm; c–d = 5 mm; e–f = 20 µm; g, j = 10 µm; h–i = 5 µm. ― Photos: a. M. Gelardi; b–d. G. Konstantinidis; e–g, j. J.M. Vidal; h–i. UdG. consisting of 3–4 strata of globose to prismatic cells 9–28 µm diam. Hymenophoral trama heteromerous, made of hyaline, loosely interwoven septate hyphae 3–9 μm wide, with abundant sphaerocytes 21–47 µm diam, and some gloeoplera 2–7 µm diam. Pileipellis and context 100–300 µm thick; pileipelliscomposed of: 1) a trichodermal suprapellis made of erect to semi erect, septate hyphae 2–6 µm diam, and abundant clavate, septate dermatocystidia 30–160 × 5–10 µm; and 2) a prosenchymatous subpellis of intricate hyphae 2–6 µm diam. Pileal context heteromerous, composed of septate hyphae 2–6 µm diam, mixed with several inflated hyphae and abundant nests and columns of sphaerocytes 21–47 μm diam. Stipitipellis a texture of interwoven, branched hyphae similar in shape and content to pileipellis dermatocystidia. Habitat, Distribution & Season ― Gregarious, subepigeous among plant debris, in broadleaved woods of Quercus and Castanea, on siliceous soil. Summer. Occurring in the central and east-Mediterranean regions, in Greece and Italy, from sea level up to 1000 m altitude. Additional material studied. GREECE, Central Macedonia, Kilkis, Koupa, 590 m, under Quercus sp., 8 July 2009, G. Konstantinidis (GK3930, duplicate BCN JMV800639)*; North Aegean, Lesbos Island, Agiasos, under Castanea sativa, 8 June 2014, sine leg. (GK7286, duplicate BCN JMV800642)*; Thessaly, Trikala, Logga, 1040 m, under Quercus petraea, 16 June 2012, G. Konstantinidis (GK6072, duplicate BCN JMV800640). – ITALY, Lazio, Roma, Nettuno, Torre Astura, 5 m, under Quercus robur, Quercus frainetto and Quercus cerris with presence of Pinus pinea, 2 Aug. 2014, M. Gelardi, M. Tulli & R. Polverini (MG630, duplicate MCVE 29085)*; ibid., 9 Aug. 2014, M. Gelardi, M. Tulli & R. Polverini (MG636, duplicate MCVE 29086)*. Notes ― Russula mediterraneensis belongs to the R. globi-spora complex of subsect. Maculatinae, and is therefore related with R. mattiroloana, but differs because of its larger basidio-mata, its hymenophore comprised of a daedaleoid, loculate inner part and a clearly lamellate external part, its subglobose to ellipsoid spores with lower warts and a distinct amyloid suprahilar plage, its trama and context featuring abundant sphaerocytes, and its ecological association with broadleaved trees. The large spore size (up to 17 µm) is similar to that ob-served in R. globispora and R. mattiroloana, but different from those found in R. dryadicola and R. heilongjiangensis. Russula meridionalis (Calonge et al.) J.M. Vidal, Mor.-Arr. & A. Paz, comb. nov. ― MycoBank MB828506; Fig. 23 Basionym. Zelleromyces meridionalis Calonge et al., Bol. Soc. Micol. Madrid 25: 302. 2000. Synonyms. Gymnomyces meridionalis (Calonge et al.) J.M. Vidal, Rev. Catalana Micol. 26: 78. 2004. Gymnomyces dominguezii Mor.-Arr. et al., Bol. Soc. Micol. Madrid 25: 301. 2000. (syn. nov.) Russula dominguezii (Mor.-Arr. et al.) Trappe & T.F. Elliott, Fungal Systematics and Evolution 2: 361. 2018. Basidiomata 1–2 cm wide, angiocarpic, sessile, subglobose to lobate or irregular. Pileus smooth, pale cream to ochraceous, drying dark reddish brown, intense red in contact with KOH. Hymenophore loculate, pale cream to ochraceous. Columella and sterile base present or absent. Odour fruity. Spores 8–11 × 7–10 μm, Q = 1.03–1.12, globose to subglo-bose, orthotropic; reticulum 0.4–0.6 µm high, composed of isolate warts and ridges, amyloid. Basidia 4-spored, 25–40 × 10–16 µm, clavate. Macrocystidia absent. Cystidioles like in var. messapicoides. Hymenophoral trama heteromerous. Suprapellis a trichoderm of septate hairs less developed than in var. messapicoides. Habitat, Distribution & Season ― Gregarious, hypogeous in continental sclerophyllous woods of Quercus rotundifolia, commonly on siliceous soil. Autumn and spring. Located in the western Mediterranean region, in Central Spain, between 400–1100 m altitude. Material studied. SPAIN, Andalusia, Cdoba, Cabra, under Quercus rotundifolia, 21 June 1997, B. Moreno-Arroyo & J. Gez (MA-Fungi 38502, holotype of Z. meridionalis; BM410, isotype)*; Cdoba, Cabra, La Nava, 1070 m, under Quercus rotundifolia, 24 May 2015, C. Lavoise, A. Paz & R. Molina (BCN IC24051506)*; Cdoba, Priego de Cdoba, Dehesilla Carcabuey, 700 m, under Quercus rotundifolia, 18 May 1993, J. Gez (MA-Fungi 32069, paratype of Z. meridionalis); ibid., 20 May 2014, J. Gez & A. Paz (BCN IC20051417)*; Cdoba, carretera de Los Villares, under Quercus rotundifolia with Cistus crispus and Pistacia lentiscus, Feb. 1996, B. Moreno-Arroyo & J. Gez BM412 (MA-Fungi 38572, holotype of G. do-minguezii); Castilla and Leon, Valladolid, Urue, montes Torozos, 830 m, under Quercus rotundifolia, on sandy soil, 17 June 2018, J. Cabero (JC180617NR)*; Extremadura, Cáceres, Torrej el Rubio, Parque Nacional de Monfrag, 370 m, under Quercus rotundifolia, on siliceous soil, 15 Apr. 2007, A. Paz (BCN IC15040721). Notes ― The holotype of Z. meridionalis is genetically very similar to the sequences of R. messapica and M. messapicoides analyzed in the present work. The three taxa seem to represent a monophyletic lineage, where no significant differences in ITS, 28S rDNA or rpb2 data were found between R. messapica and M. messapicoides. Both taxa were in turn significantly different from the holotype of Z. meridionalis and other sam-ples of this species, although no significant relationship was found between other collections identified as Z. meridionalis. Therefore, M. messapicoides is subsumed under the prioritary name R. messapica, while Z. meridionalis is re-combined into Russula but retained as an independent species due to the small differences found in ITS rDNA (2/715 bp different from R. messapica in sequences MK105663–MK105665 and MK105667) and rpb2 (6/639 bp different in MK102762). These differences did not produce a significant phylogenetic support for a monophyletic R. meridionalis, but this was probably caused by the lack of data, especially rpb2, from all specimens tested. Russula meridionalis has a restricted geographic distribution (under Quercus rotundifolia in Southern and Central Spain), and constitutes at least a partially supported distinct genetic lineage, while R. messapica var. messapicoides occurs in the same areas as R. messapica var. messapica (under Quercus ilex in the Mediterranean basin, from Eastern Spain to Greece), but did not receive any significant support as a distinct genetic lineage employing ITS, 28S rDNA, rpb2 and tef1 data. Basidiomata of R. meridionalis are angiocarpic and sessile or subsessile, with a completely loculate hymenophore, generally with traces of a columella and a small stipe or sterile base. The trichodermis (which is part of the universal veil) is reduced or partially lost in the substrate due to the hypogeous mode of growth, so the pileus does not always react intensely in contact with KOH 10 % as it does in R. messapica. Microscopical features of R. meridionalis are very similar to those of R. messapica, viz. tetrasporic basidia, subreticulate spores ornamented with an incomplete reticulum, and a more or less developed trichodermal suprapellis made of septate hairs with golden yellow encrusted pigment, which turns red in contact with KOH. Microscopical characters of Gymnomyces dominguezii (Moreno-Arroyo et al. 1999) also match those of R. meridionalis, viz. suprapellis a trichoderm, basidia tetrasporic and globose to subglobose spores 8–10 μm diam, ornamentedwithanincompletereticulum 0.4–0.6 µm high. Despite the fact that attempts to produce genetic data from G. dominguezii yielded no useful results, this species is here considered a synonym of R. meridionalis because of their morphological similarities and identical geographical distribution. Russula messapica Sarnari var. messapica in Sarnari, Boll. Assoc. Micol. Ecol. Romana 5, 18: 12. 1989 ― Fig. 24a left Material studied. ITALY, Apulia, Taranto, Avetrana, Bosco Modunato, under Quercus ilex, on calcareous soil, 15 Dec. 2013, C. Agnello (AH 46373)*. – SPAIN, Balearic Islands, Menorca, Alaior, Binixems, on calcareous soil, 120 m, under Quercus ilex, 18 Nov. 2011, J. Llistosella (BCN JL201111182)*; Catalonia, Girona, Rupià, Bosc Geltr 120 m, under Quercus ilex, close to R. mes-sapica var. messapicoides, 22 May 1993, J.M. Vidal (BCN JMV930522-12a). J.M. Vidal et al.: Revision of sequestrate Russulaceae (holotype of Gymnomyces dominguezii). Spores in Melzer. ― Scale bars: a b. A. Paz; c–h, l. J.M. Vidal; i–k. UdG. Russula messapica var. messapicoides (Llistos. & J.M. Vidal) J.M. Vidal, Llistos., Kaounas & P. Alvarado, comb. & stat. nov. ― MycoBank MB828507; Fig. 24 Basionym. Macowanites messapicoides Llistos. & J.M. Vidal, Rivista Micol. AMB 38, 2: 155. 1995. Synonym. Russula messapicoides (Llistos. & J.M. Vidal) Trappe & T.F. Elliott, Fungal Systematics and Evolution 1: 236. 2018. Basidiomata russuloid, pseudoangiocarpic, stipitate. Pileus 0.5–2 cm wide, rounded or occasionally bi- or trilobate, areo-late, papillose, pale yellow to orange-yellow, intensely red in contact with KOH 10 %; margin laterally open, alveolate to sub-lamellate. Hymenophore loculate to sublamellate, pale yellow to pale orange. Stipe-columella 0.3–0.7 × 0.15–0.2 cm, concolourous with pileus; context white. Odour fruity, similar to fermented fruits of Sorbus domestica. Spores 8–10 × 7.5–9.5 μm, Q = 1.0–1.1, globose to subglo-bose, subheterotropic to orthotropic; reticulum 0.6–1.2 µm high, amyloid, made by isolate warts forming ridges. Basidia 4-spored, 40–56 × 12–15 µm, clavate. Macrocystidia absent. Cystidioles 44–67 × 7–14 µm, scarce. Hymenophoral trama heteromerous. Suprapellis a trichoderm arranged in tufts of septate hairs 40–125 × 3–6 µm, with golden yellow encrusted –b = 1 cm; c, h–i, l = 10 µm; d–g = 20 µm; j–k = 5 µm. ― Photos: a. J. Cabero; pigment (soon dissolved in KOH) and sometimes also intracellular pigment (reddish brown in KOH); basal elements ampul- laceous, 8–30 µm diam; dermatocystidia absent. Habitat, Distribution & Season ― Gregarious, semi-hypogeous under fallen leaves, in sclerophyllous woods of Quercus ilex and Q. coccifera, commonly on calcareous soil, occasionally coexisting with R. messapica var. messapica. Spring to autumn. Widely distributed along the Mediterranean littoral, from Greece to Northern Spain, between 100–650 m altitude. Material studied. GREECE, Central Greece, Attica, Katsimidi, 650 m, under Quercus ilex with Pinus halepensis and Quercus coccifera, on calcareous soil, 5 June 2013, V. Kaounas (VK2998)*; ibid., 8 May 2014, V. Kaounas (VK3368)*; ibid., 28 May 2014, V. Kaounas (VK3411, duplicate BCN JMV800682)*; Crete, Retymnon, Arkadi, 520 m, under Quercus ilex, on calcareous soil, 27 Mar. 2016, V. Ramoutsakis (GK9341, duplicate BCN JMV800645)*. – SPAIN, Castilla and Leon, Le, Santibáz de la Isla, under Quercus rotundifolia, Quercus faginea and Fraxinus angustifolia, on siliceous soil, 28 Nov. 2006, A. Paz as M. messapicoides (BCN IC28110613)*; Catalonia, Girona, Rupià, Bosc Geltr 120 m, under Quercus ilex, in the same habitat as Tuber aestivum, 28 Apr. 1992, J.M. Vidal (BCN JMV920428-1, paratype of M. messapicoides); ibid., 22 May 1993, J.M. Vidal (BCN JMV930522-12b, paratype of M. messapicoides); ibid., 26 May 1993, J.M. Vidal & J. Llistosella (BCN JL1493, holotype of M. messapicoides)*; ibid., 17 Oct. 2001, J.M. Vidal (BCN JMV20011017-1). b. VK2998b/3368b. Red reaction of the pileus in contact with KOH 10 %. ― c. VK3411. Basidiomata. ― d–e. JMV920428-1 (paratype of M. messapicoides). d. Basidiomata; e. spores in Melzer. ― f. GK9341. Spores in Melzer. ― g–m. JMV20011017-1. g. Basidiomata; h. red reaction of the hairs of the suprapellis in contact with KOH 5 %; i. detail of suprapellis and subpellis; j. hymenium with one cystidiole and subhymenium; k–m. SEM images of spores. ― Scale bars: a–d, g = 1 cm; e–f = 10 µm; h–j = 20 µm; k–m = 5 µm. ― Photos: a, d–j. J.M. Vidal; b–c. V. Kaounas; k–m. UdG. Notes ― Macowanites messapicoides was firstproposed as a new species more or less identical with Russula messapica, except for its sequestrate habit (Llistosella & Vidal 1995). Martín et al. (1999) found that RFLPprofiles of mt-LrDNA, mt-SrDNA and ITS nrDNA regions were identical in R. messapica and M. messapicoides, and only the highly variable IGS region could be used to discriminate between them. However, Martín et al. (1999) only analyzed a single specimen from each species, and these were found growing at the same time only a few meters apart from one-another (Llistosella & Vidal 1995). Martín et al. (1999) suggested that M. messapicoides was recently derived from R. messapica, but noted that the status of both taxa as independent species could not be confirmed by their results. Later, Miller et al. (2006) referred to this case as an example of an angiocarpic taxon being conspecific with a gymnocarpic species of Russula, even though they did not formally propose the synonymy. In the present work, four DNA markers (ITS, 28S rDNA, rpb2, tef1) were sequenced from several specimens of M. messapicoides and R. messapica, originating from different countries of the Mediterranean basin. Results show that no significant differences exist between the two taxa (tef1 GenBank accessions MK102731–MK102735), but both taxa are here kept as distinct varieties because of their different habit. The rank of variety is the first infraspecific category recommended by Art. 4 of the International Code of Nomenclature (Turland et al. 2018), and there is currently no consensus about the application of other ranks, such as subspecies or forma (Turner & Nesom 2000). Russula messapica var. messapica, R. messapica var. messapicoides, and R. meridionalis were significantly related to subsect. Puellarinae (PP 1.00, BP 90), in concordance with previous phylogenetic re-constructions of Russula (Miller & Buyck 2002, Bazzicalupo et al. 2017). The species most closely related are R. cessans, R. odorata, R. puellaris and R. versicolor. Morphologically, R. messapica var. messapicoides is a se-questrate miniature of R. messapica var. messapica, where lamellae and stipe are progressively transformed into alveoli and columella. In the specimens from Attica (Greece) studied in the present work, the pileus is barely closed, the hymenophore is daedaleoid-sublamellate, and the stipe is well developed. On the other hand, the specimens from Catalonia (Spain) present an almost closed pileus, loculate hymenophore, and a reduced stipe. In all specimens, the suprapellis is a yellowish trichoderm that turns red in contact with KOH, the hymenium is formed by J.M. Vidal et al.: Revision of sequestrate Russulaceae tetrasporic basidia and scattered macrocystidia, and spores are subreticulated. The spores are ellipsoid and heterotropic in R. messapica var. messapica, showing a suprahilar plage, and subglobose to globose, orthotropic, without a suprahilar plage in R. messapica var. messapicoides, which produces statismospores, except in Greek collections from Attica where the spores are subheterotropic and the suprahilar plage is still present. Both taxa seem to occur under Quercus ilex across the entire Mediterranean region. Russula mistiformis (Mattir.) Trappe & T.F. Elliott, Fungal Systematics and Evolution 1: 236. 2018 ― Fig. 25 Basionym. Martellia mistiformis Mattir., Malpighia 14: 81. 1900. Synonyms. Hydnangium mistiforme (Mattir.) Zeller & C.W. Dodge, Ann. Missouri Bot. Gard. 22: 372. 1935. Octaviania mistiformis (Mattir.) Svrček in Pilát, Flora ČSR B1, Gasteromycetes: 743. 1958. Gymnomyces mistiformis (Mattir.) T. Lebel & Trappe, Mycologia 92, 6: 1199. 2000. Martellia mediterranea G. Moreno et al., Mycotaxon 42: 227. 1991. (syn. nov.) Russula mediterranea (G. Moreno et al.) Trappe & T.F. Elliott, Fungal Systematics and Evolution 1: 236. 2018. Basidiomata 0.6–2.2 cm wide, angiocarpic, subglobose to obovoid or tuberiform, sessile, lacking a sterile base, but occasionally attached to the soil by a fragile mycelial strand the same colour as the pileus. Pileus at first finely tomentose, white, then smooth, with olivaceus, pastel yellow (3A4) or cream, pale orange (5A3) tones, finally displaying brown (7E8) maculae because of contact with air and handling; basally open and alveolate in old specimens. Hymenophore loculate, laby- rinthoid, at first pale yellow (4A3) or pale orange (6A3), finally brown (6C6–7D7–7E8); locules 0.4–1.2 × 0.1–0.4 mm (2–4 per mm), radially arranged, elongated, sinuous; fresh spore mass in locules pale yellow (4A5) to brownish red (8C7); dark brown (6F8) in exsiccata. Columella absent. Odour in young specimens resembling that of an apple, very intense in mature specimens, similar to Tuber melanosporum. Spores (8.5–)9.5–11(–12.5) × (8–)8.5–10(–10.5) µm, Q = 1.1–1.2, subglobose to ovoid, orthotropic, uniguttulate, echinate, orange to reddish in KOH; warts of regular length, 0.8–1.6(–3) µm high, yellow, deeply amyloid, cylindrical with an obtuse tip, sometimes interconnected with low ridges and with some isolated verrucae among them; hilar appendix about 2 µm long. Basidia scarce, typically 4-spored, but sometimes also 1–3-spored, 30–50 × 12–18 µm, clavate to broadly clavate,filled with small dark yellowish or dark reddish oil droplets, soon collapsing, originating in the deep subhymenium; sterig-mata 3–7(–10) µm long, conical. Basidioles scarce, 21–35 × 7.5–12.5 µm, clavate. Macrocystidia absent. Cystidioles 27–40 × 6.5–9 µm; abundant in immature external locules, similar to dermatocystidia, cylindrical to fusoid, septate, sinuous, acute, capitulate, moniliform, filled with granular hyaline or yellowish content; rare in mature locules, cylindrical to fusoid or la-geniform, typically capitate, not exceeding basidia in length. Paraphysoid cells abundant, 20–40 × 8–16 µm, commonly 1-septate. Subhymenium pseudoparenchymatous, formed of 2–3 layers of polygonal cells, 7–20(–25) µm diam. Hymenophoral trama 10–20 µm wide, formed of hyphae about 3–8.5 µm, connecting with those of pileal context; inflated elements up to 12 µm can be found sometimes in tramal anastomoses as well as sphaerocytes 10–20 µm diam. Pileipellis and context 80–240 µm thick, separable from the hymenophore; pileipellis thin, consisting of: 1) a trichodermal suprapellis, soon collapsing into a brownish mass, formed of very fragile, tapered dermatocystidia 20–40 × 3–6 µm, frequently with a mucronate apex; and 2) a prosenchymatous subpellis 20–30 µm thick, formed of interwoven subgelatinized hyphae 2–5.5 µm diam, slightly differing from those of the pileal context. Pileal context 70–220 µm thick, prosenchymatous, formed by interwoven subgelatinized hyphae about 2–5.5 µm thick, lacking sphaerocytes. Thromboplera abundant in trama and context. Habitat, Distribution & Season ― Solitary to gregarious, hypogeous, associated with Quercus, Castanea and Pinus, on siliceous soil. From spring to late autumn. Found throughout the Mediterranean region, from Greece to Spain, from sea level up to 1200 m altitude. Material studied. GREECE, North Aegean, Lesbos Island, 500 m, under Castanea sativa with Crataegus monogyna and Quercus coccifera, on siliceous soil, 29 Jan. 2017, G. Fransouas (BCN JMV800652)*. – ITALY, Sar-dinia, Nuoro, Orune, under Quercus suber, May 1900, U. Martelli (FH, herb. N. Patouillard, original material of M. mistiformis); Oristano, Pau, Is Lottus, 550 m, under Quercus ilex, on siliceous soil, 1 May 1989, P. Fantini (BCN JMV800124); Oristano, Santu Lussurgiu, Rio e Messi, 650 m, under Quercus pubescens and Alnus glutinosa, on siliceous soil, 28 Oct. 1999, P. Fantini (BCN JMV800125)*. – SPAIN, Castilla and Leon, Segovia, La Granja de San Ildefonso, 1200 m, under Quercus pyrenaica, on siliceous soil, 27 Nov. 1997, F. García (AMC H-69, duplicate BCN JMV971127)*; Zamora, Tábara, under Pinus sylvestris with Quercus rotundifolia and Erica arborea, on siliceous soil, 6 Dec. 2010, J. Cabero (JC101206BT, duplicate BCN JMV800663)*; ibid., 2 Jan. 2011, J. Cabero (JC110102NR)*; ibid., 17 Nov. 2013, J. Cabero (JC131117NR, duplicate BCN JMV800662)*; Zamora, Villar del Buey, under Quercus pyrenaica, on siliceous soil, 5 Mar. 2017, J. Cabero (JC170305NR, duplicate BCN JMV800661)*; Catalonia, Girona, Quart, Sant Mateu de Mont-negre, Les Gavarres, 350 m, under Quercus ilex, Quercus suber and Pinus pinaster, 7 June 2014, F. Rodríguez (BCN JMV20140607-3)*; Girona, Sant Sadurní de l’Heura, Can Torrent, Les Gavarres, 150 m, under Pinus radiata, 23 May 1991, J.M. Vidal asʻM. pilaʼ (BCN JMV910523-3); ibid., 11 May 1992, J.M. Vidal as ʻM. pilaʼ (BCN JMV920511-3); ibid., 26 June 1992, J.M. Vidal as ʻM. pilaʼ (BCN JMV920626-1)*; ibid., Pla de Banyeres, Les Gavarres, 80 m, under Pinus pinea, on siliceous soil, 24 May 2016, F. Rodríguez & J.M. Vidal (BCN JMV20160524-1)*; Girona, Santa Cristina d’Aro, Romanyà, Can Pons, Les Gavarres, 365 m, under Pinus pinaster, 26 Dec. 1990, J.M. Vidal as ʻM. pilaʼ (BCN JMV901226-6); ibid., 20 Nov. 1993, J.M. Vidal, A. Montecchi & M.P. Martín (BCN JMV931120-9); ibid., 26 Dec. 1996, J.M. Vidal (BCN JMV961226-1); Extremadura, Cáceres, Parque Nacional de Monfrag, under Quercus suber with Cistus ladanifer, 9 Nov. 1987, A. Montecchi et al. (AH GM-RG11057, holotype of M. mediterranea; BCN JMV800632, isotype)*; Cáceres, Jarandilla de la Vera, under Quercus pyrenaica and Pinus sp., 15 Dec. 2012, A. Paz (BCN IC15121207/IC15121208/IC15121209). Notes ― Martellia mistiformis was described and illustrated by Mattirolo (1900) to accommodate several collections found by U. Martelli under Quercus ilex and Q. suber in the island of Sardinia (Italy), as well as additional collections from the island of Sicily. Mattirolo highlighted the hyphal hymenophoral trama, the predominantly tetrasporic basidia and the umber-coloured, globose to subglobose or slightly ellipsoid, echinate spores of 10 µm diam. This species was redescribed by Lebel & Trappe (2000) in a paper dedicated to the study of the generic types of sequestrate Russulales, and recombined into the genus Gymnomyces after observing sphaerocytes in the hymenophoral trama of the isotypic material at FH and NY herbaria. Recently, this species was re-combined into genus Russula by Elliott & Trappe (2018). The first genetic data of R. mistiformis obtained by Calonge & Martín (2000) from specimens collected and illustrated by Vidal (1991a) as Octaviania pila, and subsequently described by Vidal (1991b) as Martellia pila, were found to have a close relationship with R. foetens. These early sequences (AF230893 and AF230894) contained some sequencing errors and so new data from the same collections, GM-RG11057 (AH) and JMV920626-1 (BCN), were produced in this study (Table 1). Whitbeck (2003) obtained additional sequences from a specimen collected by A. Montecchi in Italy (AM1653). In the present work, new collections from Italy and Spain were tested and shown to be genetically identical to one-another. ITS and 28S rDNA obtained from the holotype of Martellia mediterranea (AH GM-RG11057), were identical to those of R. mistiformis, and both species have also very similar morphology, ecological preferences and geographi m. basidia, basidioles, paraphysoid cells and trama; n. spores in KOH 5 %. ― o. AH GM-RG11057 (holotype of Martellia mediterranea). Spores in KOH 5 %. ― Scale bars: b–c, i, l, n–o = 10 µm; d, g = 1 cm; e–f = 5 µm; h, j–k, m = 20 µm. ― Photos: b–d, g–o. J.M. Vidal; e–f. UdG. cal distribution, so they are here considered synonyms. Trappe (in Moreno et al. 1991), studied the holotype of M. mistiformis at TO and the isotype at FH, concluding that the spores of M. mediterranea are smaller than those of M. mistiformis, and the spore ornamentation thinner. These differences in spore size and ornamentation are here considered merely as part of the intraspecific variability of R. mistiformis. Russula mistiformis belongs to sect. Ingratae subsect. Foetentinae, and is closely related to other European sequestrate russuloid taxa, such as R. cerea and R. pila. It differs from these species due to the thickness of pileipellis and context (80–240 µm in R. mistiformis, 70–125 µm in R. cerea, 150–500 µm in R. pila), bisporic or tetrasporic basidia (typically bisporic in R. cerea, tetrasporic in R. pila), hymenium with capitate cystidi-oles and abundant paraphysoid cells, but lacking macrocystidia (macrocystidia thick walled and clavate in R. cerea, macrocystidia and cystidioles absent in R. pila) and subglobose to ovoid spores (globose in R. cerea and R. pila). In addition, the three species seem to be ecologically isolated, with R. mistiformis being apparently endemic to the Mediterranean region under Quercus and Pinus, while R. cerea is associated with lowland and montane trees (conifers and broadleaved trees) in Mediterranean and Central Europe, and R. pila shows preference for Fagus sylvatica stands of the Mediterranean subalpine range. J.M. Vidal et al.: Revision of sequestrate Russulaceae Russula monospora (Boud. & Pat.) Trappe & T.F. Elliott, Fungal Systematics and Evolution 1: 236. 2018 ― Fig. 26 Basionym. Hydnangium monosporum Boud. & Pat., J. Bot., Paris 2, 24: 445. 1888. Synonyms. Octaviania monospora (Boud. & Pat.) Lloyd, Mycol. Writings 7, Letter 67: 1141. 1922. Martellia monospora (Boud. & Pat.) Astier & Pacioni, Doc. Mycol. 28, 109–110: 9. 1998. Basidiomata 1.5–5.5 cm wide, angiocarpic, subglobose to ovatepyriform or tuberiform, often caespitose and deformed, irregularly depressed at the top, sometimes longitudinally plicate and basally alveolate, attached to the soil by a minute sterile base that often remains in the substrate when the basidiomata are extracted. Pileus dry, finely tomentose to minutely papillatesquamulose under the lens, evanescent (old specimens nude), pale orange (5A2) to melon (5A6) or greyish orange (5B6), covered by fulvous yellowish to ferruginous brown hairs, with wine-red (11D7) and olivaceous (3C3) maculae when handled; pale orange (5A2) to dark brown (7F7) in exsiccata. Hymenophore firm, minutely loculate, labyrinthoid, unevenly ripening, at first pure white to ochraceous, becoming pink (13A3) to purplish red (13A6) at maturity; greyish orange (5B4) to brown (7E5) in exsiccata; locules minute, 0.4–1 × 0.1–0.3 mm (2–4 per mm), elongated, compressed, flexuose, empty or filled of spores; fresh spore mass in locules initially white to yellowish, finally pink (12A3–A6) when mature; yellowish white (4A2) in immature exsiccata, and greyish red (10D5–11D5) to brownish red (10D6–11D6) in mature exsiccata. Columella absent. Odour pleasant, fruity, strongly of pineapple according to Boudier & Patouillard (1888). Spores (9–)10–13 µm, Q = 1, perfectly spherical, orthotropic, uni guttulate, echinate, hyaline, colourless for a long time, then yellowish with purplish pink hues, and finally dark pink; in contact with Melzer’s reagent the immature hyaline spores stain dark yellow and the mature pink spores stain orange; warts 0.4–1.4(–1.6) µm high, isolated, inamyloid or weakly amyloid, conical, with some verrucae among them; hilar appendix 1.5–3 × 1.5–2 μm long, sometimes retaining a sterigmal appendix. Basidia 1-spored (rarely 2-spored), 30–60 × 7–12 µm, cylindrical to oblong-clavate when young, then sinuous, constricted, lageniform-urticiform with a long neck, filled of yellow oleiferous guttules, soon collapsed; sterigmata 4–10 µm long, tapering above. Basidioles clavate, 17–38 × 7–17 µm, with numerous oleiferous guttules inside. Macrocystidia and cystidioles absent. Paraphysoid cells abundant, aseptate or 1-septate, similar to basidioles, but internally empty. Subhymenium ramose, composed of chains of 1–3 elongated cells measuring 7–15 × 4–9 µm. Hymenophoral trama 30–50 µm wide, homoiomerous, composed of loosely interwoven, hya- line, subgelatinized hyphae, 2–6 µm diam, filled of oleiferous guttules; sphaerocytes absent. Pileipellis and context 150– 350 µmthick,lacking in old specimens;pileipellisconsisting of: 1) an intricate trichodermal suprapellis of dark yellow, septate hairs and dermatocystidia, 40–100 × 2–4 µm, with a thick wall up to 1 µm, finally collapsing into a brownish mass; and 2) an undifferentiated prosenchymatous subpellis. Pileal context 125–250 µm thick, densely prosenchymatous, composed of subgelatinized hyaline hyphae 2–5 µm diam; sphaerocytes absent. Gloeoplera up to 5 µm diam, abundant in the trama and context. Thromboplera also present. Habitat, Distribution & Season ― Hypogeous to semi-hypogeous, solitary to gregarious, sometimes in compact, caespitose groups, under conifers (Pinus nigra) or deciduous trees (Quercus), on calcareous soil. Summer to winter. Found in Mediterranean and submediterranean regions of Southern Europe (Bulgaria, France and Spain), between 300–1150 m altitude. Material studied. BULGARIA, Blagoevgrad, Ilindentsi, Struma Valley, 550 m, under Pinus nigra subsp. nigra with Corylus avellana, Fraxinus ornus and Ulmus sp., on calcareous soil, 16 Jan. 2014, M. Slavova (MSL0932F1017); ibid., 31 Jan. 2016, M. Slavova (MSL1689F7395); ibid., 11 Dec. 2017, M. Slavova MSL2032F2679 (SOMF 29974, duplicate BCN JMV800686). – FRANCE, Franche-Comté, Jura, Abbévillers, sine dat., L. Quélet as ʻH. galatheiumʼ (UPS F013936, herb. L. Quélet); Provence-Alpes-Ce d’Azur, Nice, July 1885, ‘dedit D. Barla’ (PC, herb. E. Boudier, lectotype of H. monosporum); ibid., sine dat., ex E. Boudier as H. monosporum (BPI 712228, coll. C.G. Lloyd 7201); ibid., sine dat., ‘fragmentum spec. orig.’, ex E. Boudier as H. mono-sporum (UPS F016541, ex herb. G. Bresadola); Nice, Drap, Grand Bois, 12 July 1886, J.-B. Barla as ʻH. candidumʼ (NICE 2012-0-06711/06712); ibid., 23 June 1887, J.-B. Barla as ʻH. candidumʼ (NICE 2012-0-06714); ‘environs de Nice’, July 1887, as H. monosporum (FH, herb. N. Patouillard); Nice, Drap, La Bauma, 8 Aug. 1887, J.-B. Barla as ʻH. candidumʼ (NICE 2012-0-06713); Nice, Drap, Oct. 1890, J.-B. Barla as ʻH. candidumʼ (NICE 2012-0-06715-1); Nice, Drap, Grand Bois, 30 June 1891, J.-B. Barla as H. monosporum (NICE 2012-0-06715-2/6716/6717/6718). – SPAIN, Castilla and Leon, Soria, San Leonardo de Yag, Sistema Ibérico, under Quercus rotundifolia, on cal-careous soil, 15 June 2001, A. Sanz-Becerra & N. Redondo (AH 46459)*; Castilla-La Mancha, Cuenca, Área Recreativa de Los Lagunillos, Serranía de Cuenca, Sistema Ibérico, 1150 m, in Pinus nigra subsp. salzmannii forest with Buxus sempervirens and Crataegus monogyna, on calcareous soil, 10 Nov. 2017, J.A. Martínez (BCN JMV800671)*; ibid., 1120 m, in Pinus nigra subsp. salzmannii forest with Quercus faginea and Juniperus communis, on calcareous soil, 16 Dec. 2017, A. Carreres & J.A. Martínez (BCN JMV800672)*; ibid., 16 Dec. 2017, A. Carreres & J.A. Martínez (BCN JMV800673, duplicate AH 50141)*. Notes ― This rare species was collected by J.-B. Barla be-tween 1885 and 1891 in Nice (Southern France), in the vicinity of Drap, in the places called Grand Bois and La Bauma (Trim-bach 1996), and initially identified as Hydnangium candidum. Parts of these collections were sent to Boudier and also to Patouillard who, after studying them and observing the presence of monosporic basidia in all the specimens, published this species as Hydnangium monosporum (Boudier & Patouillard 1888), which was later illustrated by Boudier (1906: pl. 193) in his Icones Mycologicae. The lectotype of H. monosporum (PC) designated by Nuytinck et al. (2003), as well as original and authentic collections deposited in various public herbaria (BPI, FH, NICE, UPS), were checked for the purpose of this study. Despite being completely or partially immature, or in some cases parasitized by moulds, several microscopic features were noted, such as a trichodermal pileipellis, monosporic basidia, and globose, echinate spores measuring 9–13 µm diam, still hyaline or yellowish, but showing an intense yellow reaction in contact with Melzer’s reagent. Therefore, the specimens collected by Barla were probably too immature and led Boudier (1906) to depict them as yellowish. The sample collected by Quélet (1886) in the Jura region of France displays slightly more mature spores, with purplish pink tones in the spore wall under ammonium 10 %. Modern Spanish and Bulgarian collections studied in the present work show an intense pink colour in mature spores and hymenophore, a unique character dis-criminating this species from all other European sequestrate Russulaceae taxa. Recently, this species was re-combined into genus Russula by Elliott & Trappe (2018). Russula monospora can be confused with Lactarius stephensii, which also has monosporic basidia, a ramose subhymenium, a homoiomerous hyphal trama, and echinate spores. However, in L. stephensii, spores are distinctly subglobose instead of spherical, while numerous laticifera are present in the trama and context, which are completely absent in R. monospora. Genetically, R. monospora shows no significant relationships, but is probably close to R. consobrina (PP 0.94, BP 53) and other lineages such as subsections Russula, Viscidinae and Sardoninae (PP 0.73, BP 50). The only other sequestrate species in these clades is R. gilkeyae (Trappe 2572, listed as ‘Gymnomyces gilkeyae’, and OSC 117360, as ʻG. monosporusʼ), which probably belongs to subsect. Sardoninae (Trendel et al. 2017). h. immature spores in Melzer. ― i–j. JMV800672. i. Basidiomata initiating maturation; j. pileipellis and context. ― k–q. JMV800673. k. Basidiomata showing the non-uniform ripening of hymenophore; l. detail of monosporic basidium; m. emerging necks of basidia; n. mature spores in Melzer; o. mature spores in ammonia; p–q. SEM images of spores. ― Scale bars: b, g, j, l–m = 20 µm; c–e, h, n–o = 10 µm; f, i, k = 1 cm; p–q = 5 µm. ― Photos: b–e, g–h, j, l–o. J.M. Vidal; f. M. Slavova; i, k. A. Carreres & J.A. Martínez; p–q. UdG. J.M. Vidal et al.: Revision of sequestrate Russulaceae Russula neuhoffii (Soehner) J.M. Vidal, comb. nov. ― Myco-Bank MB828508; Fig. 27 Basionym. Hydnangium neuhoffii Soehner, Z. Pilzk., N.F. 20, 3–4: 111. 1941. Synonyms. Octaviania neuhoffii (Soehner) Svrček in Pilát, Flora ČSR B1, Gasteromycetes: 202. 1958. Hymenogaster pisiformis Velen., Opera Bot. Cech. 4: 96. 1947. (syn. nov.) Lectotype of Hydnangium neuhoffii (here designated MBT384233): GERMANY, Brandenburg, ‘Liebenthaler Wäldchen bei Marienwerder’, July 1926, W. Neuhoff (M, herb. E. Soehner 1060). Basidiomata 1–3 cm wide, angiocarpic, globose, ovoid to irregu-lar, slightly sulcate and tuberous, sessile, with a minute sterile base. Pileus pruinose, initially pure white, later orange-white to pale orange (5A2–A4), with brown maculae (7D7). Hymenophore loculate, labyrinthoid, initially yellow to orange-yellow, then ochre; pale orange in exsiccata (5A4); spore mass in locules pale orange (5A4). Stipe-columella not observed. Odour mild. (Description based on Soehner 1941, Velenovsk1947 and herbarium material). Spores 9.5–12.5 × 8.5–11 μm, Q = 1.0–1.1, globose to subglo-bose, orthotropic to subheterotropic, statismosporic, echinate, intensely yellow; warts 0.5–1 μm high, amyloid, densely packed and short, cylindrical to dentiform, or rounded at the apices, with numerous verrucae among them; hilar appendix 1–2 × 1.5 μm, straight, cylindrical to conical, sometimes retaining a sterigmal appendix; suprahilar plage present, inamyloid. Basidia 2–3-spored, 27–40 × 9–15 μm, clavate to broadly clavate; sterig-mata 3.5–6.5 µm long. Macrocystidia scarce, 55–70 × 7–9 μm, cylindrical, lanceolate, fusiform, or rostrate, with amorphous content, soon collapsed. Paraphysoid cells absent. Subhymenium made ofa layer ofglobose to prismatic cellsabout5–15 µm diam. Hymenophoral trama 40–50 μm wide, homoiomerous, made of septate, branched, tortuous hyaline hyphae 1.5–7 μm diam; inflated elements and sphaerocytes rare or absent. Pileipellis and context 100–150 µm thick; pileipellis a trichoepithelium formed by: 1) a trichodermal suprapellis made of septate hairs and dermatocystidia 15–30 × 2–5 μm,thatsoon collapses in a yellow granular mass; and 2) a pseudoparenchymatous subpellis 50–80 μm thick, made of ampullaceous cells up to 17 µm wideand globose cells 6–30(–40) µm diam. Pileal context 30–100 μm thick, formed by tortuous and intricate hyaline hyphae 2–5 μm diam, with numerous ampullaceous inflated elements up to 14 µm diam and some sphaerocytes up to 16 µm diam. Gloeoplera 3–5 μm diam, present in trama and context. Habitat, Distribution & Season ― Solitary to gregarious, hypogeous under broadleaved trees (Quercus, Carpinus, Betula), on siliceous soil. Summer and autumn. Found in temperate regions of Central Europe. Additional material studied. CZECH REPUBLIC, Central Bohemia, Jidášky, near Mnichovice, ‘in humo nigro ca 10 cm profundo, supra foliis quercinis et carpineis tecto in duobus speciminibus in betuleto, statio quarcitica’, 12 July 1945, J. Velenovsk (PRM 153797, coll. J. Velenovsk holotype of H. pisiformis). – GERMANY, Bavaria, Planegg, near Munich, in oak forest, 28 Sept. 1941, E. Soehner as H. neuhoffii (M, herb. E. Soehner 1626). Notes ― Soehner (1941) described an angiocarpic fungus found by W. Neuhoff near Berlin, naming it Hydnangium neuhoffii, characterized by whitish basidiomata, yellow to fulvous loculated hymenophore, 1–3-spored basidia and echinate spores. Some years later, Velenovsk(1947) described a similar hypogeous species collected in the south of Prague, which he named Hymenogaster pisiformis. Type collections of H. neuhoffii and H. pisiformis were compared in the present work and found to represent a single species of Russula characterized by the lack of laticifera, a pileipellis arranged in trichoepithelium, and globose to subglobose amyloid spores provided with an inamyloid plage and ornamented with densely packed short spines (< 1 µm), with rounded apices. The type material of H. neuhoffii and H. pisiformis are parasitized by moulds and, unfortunately, we were unable to obtain recent collections of these species to sequence. Russula pila (Pat.) Trappe & T.F. Elliott, Fungal Systematics and Evolution 1: 237. 2018 ― Fig. 28 Basionym. Hydnangium pila Pat., Bull. Soc. Mycol. France 26: 201. 1910. Synonyms. Octaviania pila (Pat.) Svrček inPilát, FloraČSR B1, Gasteromycetes: 199. 1958. Martellia pila (Pat.) J.M. Vidal, Butll. Soc. Catalana Micol. 14–15: 172. 1991. Gymnomyces pila (Pat.) Trappe et al., Mycotaxon 81: 200. 2002. Basidiomata 1–3 cm wide, angiocarpic, subglobose to tuberi-form, frequently caespitose, sessile, with an inconspicuous sterile base, occasionally attached to the soil by a basal my-celial strand. Pileus initially densely tomentose, greyish, then smooth, pale orange (5A4) to orange (6B6), finally dark brown (8F8) after handling or in contact with air, with aromatic viscid exudations in mature basidiomata; basally open and alveolate in old specimens. Hymenophore loculate, labyrinthoid, at first orange-white (6A2) or greyish orange (6B4), finally brownish orange (6C6); locules 0.3–1 × 0.1–0.3 mm (2–4 per mm), elongated, sinuous, minute; fresh spore mass in locules pale orange (5A4) to reddish brown (5A4–8D8); brown (6E7) in exsiccata. Columella absent. Odour fruity or similar to tuber, taste mild. Spores (9–)10–12(–13) μm, Q = 1, globose, orthotropic, echinate, reddish in KOH; warts of irregular length, 1–2.5 μm high, hyaline to yellow, deeply amyloid, cylindrical with obtuse tips, with some verrucae among them; hilar appendix short, inconspicuous. Basidia abundant, typically 4-spored, but also 2–3-spored, 35–45 × 11–17 μm, broadly clavate, filled with granulose or homogeneous dark reddish content, originating deep in the subhymenium, persistent; sterigmata conical, 3–5 µm long in tetrasporic basidia, up to 10 µm in bisporic basidia. Basidioles abundant, 25–40 × 10–14 µm, with tiny, hyaline to yellowish oily droplets. Macrocystidia absent. Cystidioles 30–60 × 3–6 µm, similar to dermatocystidia, 1–3-septate, hyaline, cy-lindrical to fusoid, sinuous, acute, capitulate, moniliform, present only in immature external hymenial locules. Paraphysoid cells scarce, 20–30 × (5–)8–14 µm, commonly 1-septate. Subhymenium formed by 2–3 layers of prismatic cells about (6–)10–20 µm diam. Hymenophoral trama 10–20 µm wide, formed by branched hyphae (2–)4–6 µm diam, with occasional inflated elements up to 12 µm and chains of sphaerocytes in tramal anastomoses 6–20 µm diam. Pileipellis and context 150–500 µm thick, homogeneous, separable from the hymenophore; pileipellis consisting of: 1) a trichodermal suprapellis formed g. section of a septum; h. spores in KOH 5 %; i. spores in Melzer; j–k. SEM images of spores. ― l–m. FH, herb. N. Patouillard (original material of Hydnangium pila). SEM images of spores. ― Scale bars: a–b = 1 cm; c–g = 20 µm; h–i = 10 µm; j–m = 5 µm. ― Photos: a–i. J.M. Vidal; j–k. UdG; l–m. AH. J.M. Vidal et al.: Revision of sequestrate Russulaceae of subulate to lageniform, capitulate dermatocystidia 20–40 × 3–6 µm, that soon collapses into a yellowish mass; and 2) a prosenchymatous subpellis 50–250 µm thick, formed by a dense mesh of interwoven, subgelatinized, yellowish hyphae 2–4 µm diam. Pileal context 100–300 µm thick, formed by inter-woven, subgelatinized, hyaline hyphae 2–4 µm diam, lacking sphaerocytes. Thromboplera present in trama and context. Habitat, Distribution & Season ― Commonly gregarious, caespitose, hypogeous to semi-hypogeous, in montane broadleaved woods of Quercus and Fagus, in different types of substrates. Summer to winter. Found in temperate regions of Southern Europe, between 200–900 m altitude. Material studied. FRANCE, Rhe-Alpes, Ain, Lepinay (Cras-sur-Reyssouze), Jura mountains, under Quercus sp., Aug. 1909, N. Patouillard (FH, herb. N. Patouillard, original material of H. pila). – SPAIN, Catalonia, Girona, Vall de Bianya, Capsacosta, Coll Pregon, 900 m, under Fagus sylvatica, on sandstone soil, 16 Aug. 1997, J.M. Vidal (BCN JMV970816-8)*; ibid., 18 Aug. 1997, J.M. Vidal (BCN JMV970818-2); Navarre, Lekunberri, road to Aralar, km. 6.5, under Fagus sylvatica and Corylus avellana, on calcareous soil, 23 Nov. 2011, P.M. Pasabán & F. Sáinz (BCN JMV800654)*. Notes ― Hydnangium pila was proposed by Patouillard (1910), to name some collections of a semi-hypogeous fungus found in oak forests of the calcareous mountain range of French Jura. He reported it as sessile, at first white becoming reddish with age, with a dense and pubescent pileipellis easily separable from the hymenophore, tetrasporic clavate basidia and echinate globose spores measuring 10–12 µm diam, suggesting a close relationship of this species with Martellia. Hydnangium pila was therefore re-combined into Martellia by Vidal (1991b), because of these features, as well as its amyloid spores and the hyphal structure of the hymenophoral trama, and later re-combined into Gymnomyces by Trappe et al. (2002), following the synonymy of the genera Martellia and Gymnomyces proposed by Lebel & Trappe (2000). Recently, this species was placed in genus Russula by Elliott & Trappe (2018). Russula pila exhibits features reminiscent of both R. mistiformis and R. cerea. Russula mistiformis also has tetrasporic basidia, but differs in its subglobose to ovoid spores, thinner pileipellis and context and presence of cystidioles in the mature hymenium. Russula cerea also has globose spores, but differs in its abundant hymenial macrocystidia, bisporic basidia, and a thinner membranous pileipellis and context. Finally, R. pila and R. cerea are commonly found in montane woods, while R. mistiformis is found in Mediterranean woods. The three species are genetically related, but R. cerea and R. pila seem to be significantly closer to one-another. All of them probably belong to Russula sect. Ingratae subsect. Foetentinae. Russula vidalii Trappe & T.F. Elliott, Fungal Systematics and Evolution 1: 234. 2018 ― Fig. 29 Replaced synonym. Gymnomyces ilicis J.M. Vidal & Llistos., Rivista Micol. AMB 38, 2: 160. 1995. Basidiomata 0.5–3.5 cm wide, angiocarpic, subglobose to tur-binate, bi-lobate, sessile or subsessile, with a residual stipe. Pileus pruinose, pure white, with pale orange to reddish brown maculae; old specimens nude, completely alveolate. Hymenophore loculate, yellowish white, then yellowish orange, pale orange, and finally deep orange. Columella white, branched. Odour fruity. Spores 9–11(–13) µm,Q= 1.0–1.1,globose to subglobose,orthotropic; warts 0.5–1 µm high, isolated, amyloid, some forming short ridges or even an incomplete reticulum. Basidia 2-spored, 30–50 × 10–15 μm, clavate. Macrocystidia 43–85 × 6–10 µm. Hymenophoral trama developing nests of sphaerocytes. Pileipellis and context 175–300 µm thick, evanescent; supra- c. Basidiomata; d. suprapellis; e. spores in Melzer; f–g. SEM images of spores. ― h–i. JC100508BT01. h. Basidiomata; i. SEM image of spores. ― Scale bars: a, c, h = 1 cm; b, d = 20 µm; e = 10 µm; f–g, i = 5 µm. ― Photos: a–e. J.M. Vidal; f–g. UdG; h–i. J. Cabero. pellis a palisadotrichoderm of cylindrical to fusiform, erect der- matocystidia 5–12 µm thick; subpellis a cutis. Habitat, Distribution & Season ― Gregarious, semi-hypogeous under leaf litter of Quercus ilex and Quercus rotundifolia, sometimes mixed with Pinus, on calcareous soil. Spring and autumn. So far known from the western Mediterranean region, in France and Spain, from sea level up to 800 m altitude. Material studied. FRANCE, Provence-Alpes-Ce D’azur, Bouches du Rhe, Saint Rémy de Provence, under Quercus ilex and Pinus halepensis, 17 Nov. 1993, L. Riousset (BCN JL1501, paratype of G. ilicis). – SPAIN, Castilla and Leon, Zamora, Toro, 800 m, under Quercus rotundifolia, on calcareous soil, 9 May 2010, J. Cabero as G. ilicis (JC100508BT01, duplicate BCN JMV800688)*; Catalonia, Barcelona, Sant Just Desvern, Can Fetj 220 m, under Quercus ilex, 13 Apr. 1992, J. Llistosella, A. Rocabruna & J. Vila (BCN JL1497, holotype of G. ilicis); Girona, Rupià, Can Candell, under Quercus ilex, 3 Apr. 2005, F. Rodríguez (BCN JMV20050403-2); Girona, Sant Sad-urní de l’Heura, Can Barris, 90 m, under Quercus ilex, on sandy basic soil, 1 Apr. 1992, J.M. Vidal (BCN JMV920401-1, paratype of G. ilicis); ibid., 17 May 2016, J.M. Vidal & F. Rodríguez (BCN JMV20160517-1)*; Girona, Viladamat, Gorners, 40 m, under Quercus ilex, 10 Apr. 1991, J.M. Vidal (BCN JMV910410-3, paratype of G. ilicis); Girona, Viladamat, Palau Borrell, under Quercus ilex, 11 May 1996, J.M. Vidal (BCN JMV960511-2). Notes ― Russula vidalii is a replacement name for Gymnomyces ilicis (Elliott & Trappe 2018), an epithet originating from the apparent association of this taxon with Quercus ilex in the Mediterranean region. However, collections BCN JL1501 (Llistosella & Vidal 1995) and JC100508BT01 (Cabero 2011) were found in mixed stands of Quercus and Pinus, so it is possible that R. vidalii has a wider host range than originally thought. It is characterized by an evanescent pileipellis and context that gives mature basidiomata an alveolate look. Genetically, R. vidalii belongs to subsect. Laricinae, displaying close affinities with R. laricinoaffinis and R. galileensis. Basidiomata of R. vidalii are relatively small (1–2.5 cm), have a poorly developed columella, a residual stipe, and a loculate hymenophore, while R. galileensis produces larger basidiomata (up to 5–6 cm), with a conspicuous stipe-columella and a sublamellate hymenium. Spore warts of R. vidalii are isolated or forming short ridges, while basidia are bisporic, and the suprapellis is a pali-sadotrichoderm. By contrast, spore warts of R. galileensis are interconnected by small ridges, basidia are tetrasporic, and the suprapellis is an intricate trichoderm. Russula vinaceodora (Calonge & J.M. Vidal) Trappe & T.F. Elliott, Fungal Systematics and Evolution 1: 240. 2018 ― Fig. 30 Basionym. Macowanites vinaceodorus Calonge & J.M. Vidal, Mycotaxon 79: 2. 2001. Basidiomata russuloid, pseudoangiocarpic, stipitate. Pileus 3–8.5 cm wide, convex to plano-convex and depressed, smooth, viscid, initially pale yellow, then pinkish white to purplish brown, darker at the center, with pale orange maculae; margin open, lamellate. Hymenophore loculate in the upper zone and sub-lamellate in the lower, pale orange. Stipe-columella 1.5–4 × 0.8–2.5 cm, white; context white. Odour vinaceous, intense; taste sweetish to slightly acrid. Spores 7–11 × 6–9.5 μm, Q = 1.05–1.17, globose to broadly el-lipsoid, heterotropic; reticulum 0.7–1.5 µm high, amyloid, made of crests and isolate warts. Basidia 2–4-spored, 30–45 × 11–16 µm, broadly clavate. Macrocystidia 50–75 × 10–16 µm, fusi-form. Hymenophoral trama with large sphaerocytes up to 50 µm diam. Subhymenium cellular. Pileipellis and context 150–250 µm thick; suprapellis arranged as an intricate ixotrichoderm of erect hyphae and cylindrical to clavate dermatocystidia 30–75 × 2.5–11 µm; subpellis an intricate ixocutis. Pileal and stipecolumella context heteromerous. Habitat, Distribution & Season ― Gregarious, hypogeous to semi-hypogeous in sandy substrates, associated with Pinus, in fixed coastal dunes. Autumn. Located in the Atlantic coast of Southern Spain. Material studied. SPAIN, Andalusia, Cádiz, Sanlar de Barrameda, Pinar de Algaida, littoral stabilized sand dunes, under Pinus pinea, 30 Oct. 2014, M. Becerra as M. vinaceodorus (AH 46374)*; Huelva, Mazag, littoral stabilized sand dunes, under Pinus pinea with Corema, Halimium and Heli-chrysum, 27 Nov. 1999, J.M. Vidal & F.D. Calonge (MA-Fungi 47416, holotype of M. vinaceodorus; BCN JMV991127-4, isotype); ibid., three young speci-mens growing fasciculate among Macowanites ammophilus, 27 Nov. 1999, J.M. Vidal & F.D. Calonge as M. vinaceodorus (MA-Fungi 46524; duplicate BCN JMV991127-5). Notes ― Recently,this specieswasre-combined into genus Russula by Elliott & Trappe (2018). Russula vinaceodora is characterized by its russuloid, pseudoangiocarpic basidiomata, with a pinkish white to purplish brown pileal colour, especially in the centre. Mature basidiomata have a very characteristic odour of fermented wine. b. Suprapellis; c. hymenium and subhymenium; d. spores in Melzer; e–f. SEM images of spores. ― g. MA-Fungi 46524 (as M. vinaceodorus). Loculated hymenophore of young basidioma. ― Scale bars: a, g = 1 cm; b–c = 20 µm; d = 10 µm; e–f = 5 µm. ― Photos: a–d, g. J.M. Vidal; e–f. UdG. J.M. Vidal et al.: Revision of sequestrate Russulaceae Genetically, R. vinaceodora belongs to subsect. Laricinae, and is closely related to R. laricina, R. murrilli and R. nauseosa. The most closely related sequestrate taxon is R. sichuanensis (Li et al. 2013), another pseudoangiocarpic species found in Picea forests of Xizang and Sichuan provinces of China, which has similar cream to yellowish lamellae, globose or subglobose spores ornamented with an incomplete reticulum, and short basidia. However, R. vinaceodora has a pinkish to violet pileus, different from the cream-coloured one of R. sichuanensis, a distinctive odour of wine, and a different habitat under Pinus pinea, in Mediterranean coastal sandy soils. IDENTIFICATION KEY TO EUROPEAN SPECIES OF SEQUESTRATE RUSSULACEAE 1. Hymenophore lactescent or with laticifera. Hymenophoral trama homoiomerous, lacking nests of sphaerocytes .... .................................... Lactarius 2 1. Hymenophore not lactescent, lacking laticifera. Hymenophoral trama heteromerous, with nests of sphaerocytes, especially in tramal anastomoses ........... Russula 8 2. Spores echinate................................ 3 2. Spores reticulate ...............................6 3. Basidia 3–4-spored. — Spores 9.5–13(–15) × 8–10(–11) µm, subglobose to broadly ellipsoid; warts 1–1.5 µm high, isolated. Basidiomata 1–4 cm, subglobose to tuberiform, with a residual stipe, pale yellow with reddish brown maculae. Hymenophore loculate, pale yellow to pale brown. Latex scant, colourless to white, changing to yellow. In subalpine conifer woods (Abies, Picea). Temperate (Alps to Rhodopes) ... . ................................... L. borzianus 3. Basidia 1-spored ...............................4 4. Spores weakly amyloid. — Spores (11.5–)12–14.5(–15) × (10–)11–13 µm, subglobosetobroadly ellipsoid; warts 0.5–2 µm high, isolated. Basidiomata 1–3.5 cm, subglobose to tuberiform, maize yellow to reddish brown. Hymenophore loculate, ochraceous to reddish brown. Latex scant and hyaline, changing to citrine yellow in young specimens, white, abundant, and almost immutable in old basidiomata. In montane woods of Carpinus, Corylus, Fagus, Quercus, Tilia. Temperate to submediterranean (British Isles to Southern Europe) . ............................ L. stephensii 4.Spores strongly amyloid .........................5 5. Spores subglobose to broadly ellipsoid. — Spores 13–15 × 11–13 μm; warts 1–2 μm high, isolated or tooth-like fused. Basidiomata 1–4 cm, subglobose to tuberiform, at first pale orange then reddish brown to violet brown. Hymenophore loculate, deeply coloured, reddish yellow to orange red. Latex white. Under Populus. Temperate to Mediterranean (Belgium to Bulgaria) ................... L. populicola 5. Spores broadly ellipsoid to ellipsoid. — Spores 14–18.5 × 12–15 μm; warts 1–2 μm high, cylindrical, isolated or for ming short ridges. Hymenium completely embebbed in a dark orange substance. Basidiomata 1–2 cm, subglobose to tuberiform, orange to reddish brown or violet brown. Hymenophore loculate, deeply coloured, dull red to dull violet. Latex watery, scant. In montane conifer woods (Abies, Pinus) or under broadleaved trees (Corylus, Quercus). Temperate (Germany, Italy, Spain) ................... L. soehneri 6. Basidia 2–4-spored. — Spores 8.5–12.5(–13.5) × 7–9.5 (–10.5) μm, subglobose to ellipsoid; reticulum 0.5–1 µm high, incomplete. Basidiomata 2–5 cm, obpyriform to tuberiform, pale orange to reddish brown, with minute depressions. Hymenophore loculate, yellowish white to orange-white. Latex scant, white, immutable. Taste sweetish, later astringent. In montane conifer woods (Cedrus, Pinus). Submediterranean (France, Morocco, Spain) . ........ L. josserandii 6. Basidia 1-spored ..............................7 7. Reticulum 1–2 μm high, complete. —Spores 10–13 × 7–9 µm, ovoid to ellipsoid. Basidiomata 1–2.5 cm, subglobose to tuberiform, whitish to pale yellowish, pileus membranous, partially evanescent, indistinctly scrobiculate or with some scattered minute openings. Hymenophore loculate, whitish to pale cream or pinkish. Latex not observed. Under Cistus and Halimium. Mediterranean (Central Spain) ........ . ................................. L. giennensis 7. Reticulum 0.5–1 μm high, incomplete. — Spores 8.5–11 × 6.5–8 µm, ovoid. Basidiomata 0.5–2.5 cm, globose to tuberiform, whitish to brownish buff, pileus membranous, partially evanescent, distinctly scrobiculate, with abundant large openings. Hymenophore loculate, whitish to cream or faintly ochraceous pink. Latex white. Under Cistus. Mediterranean (Cyprus) ........... L. subgiennensis 8. Basidiomata stipitate, pseudoangiocarpic. Spores hetero-tropic. Amyloid suprahilar plage present ............ 9 8. Basidiomata sessile, angiocarpic. Spores orthotropic. Amyloid suprahilar plage absent ................. 16 9. Spores verrucose to echinate. Warts isolated or connected by low ridges................................ 10 9. Spores subreticulated .........................14 10. Macrocystidia 45–70 µm long ................... 11 10. Macrocystidia 70–130 µm long .................. 13 11. Spores subglobose to ovoid.—Spores 7–9 × 5.5–7.5 μm; warts0.25–0.75 µm high,some forming shortridgesorconnected by short lines. Pileus 2–7 cm, orange white to pale orange, with brownish orange and dark brown maculae; margin open, alveolate to sublamellate. Hymenophore loculate, orange-white to pale orange. Stipe-columella 1–4 × 0.7–2.5 cm, with brownish orange dots. Common in coastal sand dunes, under Pinus. Mediterranean (Southern Portugal and Spain). ................. R. ammophila 11. Spores globose to broadly ellipsoid...............12 12. Warts interconnected with low ridges.—Spores 9.5–12.5 × 8.5–10.5 μm; warts 1.2–2 µm high, in groups of 2–4. Pileus 1.4–2.8 cm, pure white, belatedly maculated of pale yellow; margin closed or laterally open, sublamellate. Hymenophore loculate, pale yellow to yellow. Stipe-columella 1–2 × 0.2–0.4 cm. In montane broadleaved woods (Carpinus, Corylus, Fagus, Quercus) or conifer woods (Abies). Temperate (Eastern to Southern Europe) ..... ................................ R. candidissima 12. Warts isolated. — Spores 8.5–11(–12.5) × 7–9(–11) μm; warts 0.5–1.5 µm high. Pileus 0.5–3.5 cm, white to yellow-ish white, with yellowish orange maculae; margin closed or laterally open, alveolate. Hymenophore loculate, pale orange. Stipe-columella 0.6–1.5 × 0.15–0.4 cm. In mon-tane broadleaved woods (Carpinus, Corylus). Temperate (Eastern to Southern Europe) ............ R. candida 13. Warts 0.6–1(–1.5) μm high. — Spores (8–)10–15(–17) × (7–)9–14(–15) μm, subglobose to broadly ellipsoid. Pileus 2.3–5 cm, yellowish to yellowish buff, with dark brown maculae; margin open, lamellate. Hymenophore sublamellate-daedaleoid, cream to ochre-orange. Stipe-columella 1.6–4.2 x 0.7–1.8 cm. Under broadleaved trees (Castanea, Quercus). Mediterranean (Greece and Italy).......... . ............................ R. mediterraneensis 13. Warts (0.7–)1.5–2.5(–3) µm high. — Spores (9–)10.5– 15.5(–18) × (8–)10–15(–17) μm, globose to subglobose, some ellipsoid when immature. Pileus 1.2–3.8 cm, pale yellow to pale orange with dark brown maculae; margin open, sublamellate. Hymenophore loculate, pale yellow to titian red. Stipe-columella 0.8–2.5 × 0.3–1 cm. In mon-tane conifer woods (Abies, Picea). Temperate to submediterranean (Southern Poland, Greece and Italy) ....... ................................. R. mattiroloana 14. Reticulum 0.5 µm high. —Spores (9–)10–11.5(–15) × (7.5–) 9.5–10.5(–14) µm, subglobose; reticulum made of crests and warts. Pileus 2–6 cm, white with cream to pale umber maculae; margin radially alveolate when mature, but not open. Hymenophore loculate, cream-ochre. Stipe-columel-la 0.5–3.5 × 0.5–1.7(–3) cm. In sclerophyllous woods of Quercus. Mediterranean (Israel) ........ R. galileensis 14. Reticulum 0.7–1.5 µm high ..................... 15 15. Macrocystidia present. — Spores 7–11 × 6–9.5 μm, glo-bose to broadly ellipsoid; reticulum made of crests and isolate warts. Pileus 3–8.5 cm, pinkish white to purplish brown; margin open, lamellate. Hymenophore loculate to sublamellate, pale orange. Stipe-columella 1.5–4 × 0.8–2.5 cm. Odour intense, vinaceous. In littoral sand dunes, under Pinus. Mediterranean (Atlantic coast of Southern Spain) ................................. R. vinaceodora 15. Macrocystidia absent. — Spores 8–10 × 7.5–9.5 μm, glo-bose to subglobose; reticulum made of isolate warts and ridges. Pileus 0.5–2 cm, rounded or bi-trilobate, areolate, papillose, pale yellow to orange-yellow, intense red in contact with KOH; margin laterally open, alveolate to sublamellate. Hymenophore loculate to sublamellate, pale yellow to pale orange. Stipe-columella 0.3–0.7 × 0.15–0.2 cm, concolourous with pileus. In littoral sclerophyllous woods of Quercus ilex. Mediterranean (Greece to Spain) .................. R. messapica var. messapicoides 16. Spores reticulated ............................17 16. Spores echinate or verrucose...................18 17. Macrocystidia absent.—Spores 8–11 × 7–10 μm,globose to subglobose; reticulum 0.4–0.6 µm high, made of iso-late warts and ridges. Basidiomata 1–2 cm, subglobose to lobate or irregular, smooth, pale cream to ochraceous, drying dark reddish brown, intense red in contact with KOH. Hymenophore loculate, pale cream to ochraceous. In continental sclerophyllous woods of Quercus rotundifolia. Mediterranean (Central Spain) ........ R. meridionalis 17. Macrocystidia present but scarce.—Spores 7–9.5(–10.5) × 7–9(–10), globose; reticulum 0.5 μm high, complete to incomplete, made of low ridges and warts. Basidiomata 2–7 cm, turbinate, firmly rooted into the substrate, often cracked, cream-white to ochraceous cream, with ochra-ceous to brownish stains. Hymenophore loculate, ochra-ceous yellow to ochraceous orange, vinaceous in FeSO4. In montane woods of Pinus. Mediterranean (Cyprus) ... ................................... R. hobartiae 18. Pileipellis a trichoepithelium or an oedotrichoderm ... 19 18. Pileipellis a trichoderm.........................20 19. Pileipellis a trichoepithelium.—Spores 9.5–12.5 × 8.5–11 μm, globose to subglobose; warts dense 0.5–1 µm high, isolated. Basidiomata 1–3 cm, globose to irregular, whitish. Hymenophore loculate, yellow, orange yellow to ochre. Under broadleaved trees (Carpinus, Betula, Quercus). Temperate (Central Europe) ...................... R. neuhoffii 19. Pileipellis an oedotrichoderm. — Spores (6.5–)7.5–9.5 (–11.5) × (6–)7–9(–11) µm, globose to subglobose; warts up to 0.3 µm high, some connected with low ridges. Basidio-mata 0.5–2 cm, globose to subglobose, pruinose, whitish, with brownish red maculae. Hymenophore loculate, whitish at first, finally brownish red. Under Cistus. Mediterranean (Central Spain) . . . . . . . . . . . . . . . . . . . . R. andaluciana 20. Basidia 1-spored .............................21 20. Basidia 2–4-spored...........................22 21. Macrocystidia present. — Spores 13–15(–15.5) × 12.5– 14.5(–15) µm, globose to subglobose, weakly amyloid, yellow; warts dense, 1.5–3 µm high, isolated. Macrocystidia numerous, 30–70 × 8–16 µm, clavate. Basidiomata 1–2 cm, subglobose to tuberiform, finely tomentose, pale orange with brown maculae. Hymenophore loculate, pale orange. Temperate (Germany) ........... R. bavarica 21. Macrocystidia absent. — Spores (9–)10–13 µm, spheri- cal, weakly amyloid, intense pink at maturity; warts 0.4– 1.4(–1.6) µm high, isolated. Basidia clavate to lageniform urticiform, sometimes 2-spored. Basidiomata 1.5–5.5 cm, subglobose to tuberiform, finely tomentose to papillate-squamulose, pale orange to greyish orange with wine red and olivaceous maculae when rubbing. Old speci-mens nude. Hymenophore loculate, pink to purplish red at maturity. Under Pinus and Quercus. Mediterranean to submediterranean (Bulgaria, France and Spain)....... ................................. R. monospora 22. Warts 0.5–1µm high. —Spores 9–11(–13) µm, globose to subglobose; warts isolated, some forming short ridges or even an incomplete reticulum. Basidiomata 0.5–3.5 cm, subglobose to turbinate, with a residual stipe, pruinose, pure white, with pale orange to reddish brown maculae. Old specimens nude, completely alveolate. Hymenophore loculate, yellowish white to yellowish orange or deep orange. In sclerophyllous woods of Quercus. Mediterranean (France and Spain).................. R. vidalii 22. Warts 0.8–3 µm high ......................... 23 23. Basidia 2-spored. Macrocystidia present. — Spores (8–)9.5– 12.5(–14) μm, globose; warts variable in length, 1–3 µm high, isolated. Macrocystidia (25–)30–50 × (5–)7–12(–16) μm, cylindrical to cylindro-clavate, thick walled. Basidio- mata 1–3 cm, subglobose, smooth, greyish orange to pale brown, maculated of reddish brown. Hymenophore loculate, brown to reddish brown. In montane conifer woods (Abies, Picea, Pinus) or broadleaved woods (Carpinus, Castanea, Corylus, Fagus, Quercus). Temperate to sub-mediterranean (British Isles to Southern Europe) . . . . . . ...................................... R. cerea 23. Basidia 1–4-spored. Macrocystidia absent ......... 24 24. Spores globose. — Spores (9–)10–12(–13) μm; warts variable in length, 1–2.5 µm high, isolated. Basidiomata 1–3 cm, subglobose to tuberiform, caespitose, pubescent, greyish to pale orange or orange, maculated of dark brown and producing aromatic exudations. Hymenophore locu-late, orange-white to greyish orange or brownish orange. In montane broadleaved woods (Fagus, Quercus). Temperate (Southern Europe)......................... R. pila 24. Spores subglobose to ovoid.—Spores (8.5–)9.5–11(–12.5) × (8–)8.5–10(–10.5) µm; warts of regular length, 0.8– 1.6(–3) µm high, isolated. Basidiomata 0.6–2.2 cm, sub-globose to tuberiform, finely tomentose, pastel yellow to pale orange, maculated of brown. Hymenophore loculate, pale yellow, pale orange to brown. In woods of Castanea, Pinus and Quercus. Mediterranean (Greece to Spain) . . .................................. R. mistiformis DISCUSSION The multigene phylogenetic analysis and morphological revision of European sequestrate Russulaceae taxa carried out in this work, allowed us to clarify several taxonomic issues at specific and supraspecific levels. The overall topology of the multigenic analyses was not different from those obtained by other studies focused on gymnocarpic species (Shimono et al. 2004, Verbeken et al. 2014b, Kong et al. 2015, Looney et al. 2016, Buyck et al. 2017), except for lower support values for J.M. Vidal et al.: Revision of sequestrate Russulaceae some supraspecific lineages of Lactarius. These differences could be due to phylogenetic noise introduced by some se-quences analyzed in the present work, or differences in the alignment of the highly variable ITS rDNA region. In contrast to Lactarius, multiple significant supraspecific clades were found within Russula, in concordance with previous phylogenetic reconstructions of this genus (Kong et al. 2015, Looney et al. 2016). Most of these clades include type species of subgenera, sections and/or subsections proposed by different classical authors on the basis of morphological evidence (Romagnesi 1985, Singer 1986, Bon 1988, Sarnari 1998, 2005), but until now, no formal re-arrangement had been proposed to incorporate the current genetic knowledge into the often overlapping nomenclature available. Kong et al. (2015) informally named the genetic lineages of Russula with a representative species or a supraspecific name without rank, and discussed the different supraspecific taxa in classical literature putatively matching these lineages. In the present work, eight major clades were found in Russula (Fig. 1). The largest ones (Ingratae, Rigidae, Russula s.str.) have often been treated as subgenera or sec-tions, while the remaining were usually considered as sections or subsections. Treating all of them as sections could be a natural solution requiring few nomenclatural changes, with the major phylogenetic clades within these sections deserving the status of subsection. Most European sequestrate Russulaceae species tested in the present work nested within previously known supraspecific clades, except two: Russula monospora and R. hobartiae. Russula monospora nested within a poorly supported clade formed by subsections Russula s.str. (= Emeticinae), Sardoninae (= Firmae) and Viscidinae. This clade was recovered also in the analyses ofKong etal.(2015),butreceived no significant support (PP 0.87). Russula monospora represents a distinct lineage within this clade, and so it could deserve its own sub-section. However, a deeper study of the whole group and the other isolated lineages (e.g., those of R. consobrina and R. fellea), would be necessary to propose a stable nomenclature at the supraspecific level. The other taxon studied in the present work not matching any known supraspecific clade, is the new species R. hobartiae, which is significantly related with a putative specimen of R. ochrophylla. The identity of this specimen needs to be compared with the type collection (Adamčik et al. 2017), in order to support the introduction of a supraspecific name for the monophyletic lineage shared with R. hobartiae. All other European sequestrate Russulaceae taxa belong to existing supraspecific taxa, except for R. ammophila, which belongs to a significantly supported clade lacking a formal name, related to subsections Foetentinae, Pectinatinae and Subvelatae, within sect. Ingratae. This clade is often referred to as the /R. amoenolens clade, but genetic data support a formal name at the same rank of the other clades. European sequestrate Russulaceae taxa have unequal intra- specific variability in the DNAmarkers analyzed.Some species were found to be genetically homogeneous (e.g., R. mistifor-mis), while others presented conspicuous random differences between most individuals (e.g., R. candidissima), or had a genetically homogeneous core and one or few individuals showing some differences (e.g., R. mediterraneensis). In addition, there was one species recently evolved from a gymnocarpic taxon, which could only be discriminated from the former using tef1 marker (R. ammophila). The diverse phylogenetic status of these species suggests differences in their evolutive past and present, but these can be due to very different factors, such as their life-cycles, mating systems, metapopulation structure, hybridization and introgression events, or selection pressure, to name a few. The future of these lineages remains, of course, an open question, and so predictions based on their eventual status are risky based on current evidence. Partially or recently isolated lineages may thus: 1) genetically diverge from each other and lead to fully isolated species; 2) merge again if reproductive barriers disappear; or 3) become extinct after the conditions that favored diversifi cation change. Two varieties of Russula messapica are here recognized: the original gymnocarpic phenotype (var. messapica) and a pseudo-angiocarpic phenotype (var. messapicoides) that cannot be phylogenetically discriminated even using up to four DNA mark-ers. This is a very rare case, but a few other examples can be found: for instance, a broad morphological variation from gymnocarpic to angiocarpic habit can be observed in Hydnangium sublamellatum (Bougher et al. 1993), or Setchelliogaster te-nuipes var. rheophyllus (Martín & Rocabruna 1999, Lago et al. 2000, Martín & Moreno 2001). Hibbett et al. (1994) demon-strated that a pseudoangiocarpic forma can be triggered by a recessive allele at a single locus in Lentinus tigrinus, and so it is possible that specific matings or even environmental factors can be responsible for the apparition of sequestrate phenotypes. However, despite the evidence provided here by multigenic data, we cannot exclude the possibility that none of the markers selected reflect a hypothetical reproductive isolation between R. messapica var. messapica and R. messapica var. messapicoides. In addition, a closely related species with an exclusively angiocarpic habit (R. meridionalis) was found to have at least a partial genetic isolation, probably linked to its restricted geographic distribution. Despite the fact that ITS and 28S rDNAcannot be employed to significantly discriminate between R. messapica and R. meridionalis, the latter taxon is retained as an independent species because: 1) this is the most conservative decision in nomenclatural terms; and 2) the only rpb2 sequence successfully obtained from samples of R. me-ridionalis (MK102762) seems different enough from those of R. messapica (MK102763–MK102766), with 6/634 bp varying between both species, vs only 1/634 variable sites among rpb2 sequences of R. messapica. Therefore, a more complete rpb2 dataset would likely support the separation between R. meridionalis and R. messapica. Angiocarpic species are present in most lineages of Lactarius and Russula, confirming that transitions from gymnocarpic ancestors occurred multiple times. Such transitions were often explained as an evolutive adaption to adverse and arid environmental conditions (Thiers 1984, Buyck 1995, Bougher & Lebel 2001, Trappe & Claridge 2005, Smith et al. 2006), although an-giocarpic species could be far more frequent in tropical climates than previously assumed (Verbeken et al. 2014b). Wilson et al. (2011) showed that angiocarpic species in Agaricomycetes appeared rather recently, but evolve at similar or even faster rates than the gymnocarpic lineages they have derived from, eventually predominating over them. In the present phylogenetic reconstructions, most angiocarpic species seem to be isolated in gymnocarpic lineages, but small groups of angiocarpic species can be found too (e.g., R. candidissima and related species with-in subsect. Firmiores). Interestingly, the clade named /R. tapawera is composed mainly of Australian and South American angiocarpic species with scattered gymnocarpic taxa, such as R. purpureoflava, R. tawai or R. tricholomopsis (Lebel & Tonkin 2007, Trierveiler-Pereira et al. 2015). Such an abundance of sequestrate species in the Southern Hemisphere could be due to: 1) very frequent transitions from gymnocarpic to angiocarpic states and/or a faster divergence rate of angiocarpic lineages; 2) an angiocarpic ancestor of the entire lineage with scattered reversions to a gymnocarpic state; or 3) a fragmented picture of the biogeographical distribution of gymnocarpic russuloid taxa in the Southern Hemisphere. Additional information is critically needed to evaluate these hypotheses. Table 2 Comparison of the different European sequestrate species of Lactarius and Russula ordered by systematic sections. Basidiomata type Columella Suprapellis Hymenophoral trama Cystidia Basidia Spores (µm) Ornamentation Host Lactarius subgen. Lactarius giennensis L.angiocarpic, sessile branched, inconspicuous trichoderm homoiomerous cystidioles 1-spored 10–13 × 7–9 reticulum Cistaceae populicola L.angiocarpic, sessile absent or inconspicuous trichoderm homoiomerous absent 1-spored 13–15 × 11–13 isolated and tooth-like Salicaceae fused warts soehneri L.angiocarpic, sessile absent trichoderm homoiomerous absent 1-spored 14–18.5 × 12–15 isolated warts and Fagaceae and Pinaceae short ridges stephensii L.angiocarpic, sessile absent or inconspicuous trichoderm homoiomerous absent 1-spored 12–14.5 × 11–13 isolated warts Fagaceae (weakly amyloid) subgiennensis L.angiocarpic, sessile branched, inconspicuous trichoderm homoiomerous cystidioles 1-spored 8.5–11 × 6.5–7.8 incomplete reticulum Cistaceae Lactarius subgen. Russularia borzianus L.angiocarpic, sessile percurrent, branched trichoderm to homoiomerous cystidioles 3–4-spored 9.5–13 × 8–10 isolated warts Pinaceae oedotrichoderm josserandii L.angiocarpic, sessile absent or inconspicuous oedotrichoderm homoiomerous cystidioles 2–4-spored 8.5–12.5 × 7–9.5 incomplete reticulum Pinaceae Russula sect. Ingratae R. ammophila angiocarpic to pseudoangiocarpic, stipe-columella intricate trichoderm heteromerous macrocystidia 2–4-spored 7–9 × 5.5–7.5 isolated warts and Pinaceae stipitate short ridges cerea R.angiocarpic, sessile absent or inconspicuous trichoderm sphaerocysts only in macrocystidia 2-spored 9.5–12.5 isolated warts Fagaceae and Pinaceae tramal anastomoses mistiformis R.angiocarpic, sessile absent trichoderm sphaerocysts only in cystidioles 1–4-spored 9.5–11 × 8.5–10 isolated warts Fagaceae and Pinaceae tramal anastomoses pila R.angiocarpic, sessile absent trichoderm sphaerocysts only in cystidioles 3–4-spored 10–12 isolated warts Fagaceae tramal anastomoses Russula sect. Rigidae andaluciana R.angiocarpic, sessile inconspicuous oedotrichoderm prosenchymatous absent 4-spored 7.5–9.5 × 7–9 isolated warts and Cistaceae low ridges Russula sect. Russula candida R.angiocarpic, stipitate stipe-columella trichoderm heteromerous macrocystidia 1–3-spored 8.5–11 × 7–9 isolated warts Fagaceae candidissima R.angiocarpic to pseudoangiocarpic, stipe-columella cutis of repent hyphae heteromerous macrocystidia 2–4-spored 9.5–12.5 × 8.5–10.5 groups of 2–4 warts Fagaceae and Pinaceae stipitate interconnected by low ridges galileensis R.angiocarpic, stipitate stipe-columella intricate trichoderm heteromerous macrocystidia 4-spored 10–11.5 × 9.5–10.5 reticulum of crests Fagaceae and warts hobartiae R.angiocarpic, sessile absent or inconspicuous palisadotrichoderm prosenchymatous macrocystidia 2–4-spored 7–9.5 × 7–9 reticulum of low Pinaceae ridges and warts mattiroloana R.pseudoangiocarpic, stipitate stipe-columella trichoderm heteromerous macrocystidia 2–4-spored 10.5–15.5 × 10–15 isolated warts Pinaceae mediterraneensis R.pseudoangiocarpic, stipitate stipe-columella trichoderm heteromerous macrocystidia 2–3-spored 10–15 × 9–14 isolated warts Fagaceae meridionalis R.angiocarpic, sessile absent or present trichoderm heteromerous cystidioles 4-spored 8–11 × 7–10 reticulum of ridges Fagaceae and isolated warts messapica R.pseudoangiocarpic, stipitate stipe-columella trichoderm heteromerous cystidioles 4-spored 8–10 × 7.5–9.5 reticulum of ridges Fagaceae var. messapicoides and isolated warts monospora R.angiocarpic, sessile absent intricate trichoderm homoiomerous absent 1-spored 10–13 isolated warts Fagaceae and Pinaceae (weakly amyloid) vidalii R.angiocarpic, sessile percurrent, branched palisadotrichoderm heteromerous macrocystidia 2-spored 9–11 isolated warts and Fagaceae short ridges vinaceodora R.pseudoangiocarpic, stipitate stipe-columella intricate ixotrichoderm heteromerous macrocystidia 2–4-spored 7–11 × 6–9.5 reticulum of crests and Pinaceae isolated warts Russula not classified bavarica R.angiocarpic, sessile absent trichoderm homoiomerous with macrocystidia 1-spored 13–15 × 12.5–14.5 isolated wartsendomacrocystidia (weakly amyloid) neuhoffii R.angiocarpic, sessile absent trichoepithelium homoiomerous macrocystidia 2–3-spored 9.5–12.5 × 8.5–11 isolated warts Fagaceae J.M. Vidal et al.: Revision of sequestrate Russulaceae Speciation processes seem to be quite diverse among European sequestrate russuloid taxa. For example, L. giennensis and L. subgiennensis seem to represent independent genetic lineages maybe diverging as a result of allopatric speciation. Both species can be found associated with Cistaceae hosts in acidic soils in the Mediterranean basin, with L. giennensis apparently restricted to Western and Southern Spain, while L. subgiennensis is so far known only from the island of Cyprus at the opposite end of the Mediterranean basin. This clade apparently represents one of the few lineages within Russulaceae associated with Cistaceae hosts, although comprehensive phylogenetic assessments of fungal diversity within Cistaceae ecosystems are generally lacking (Fellner & Biber 1990, Lavo-rato 1991, Vila & Llimona 1999, 2002, 2006, 2009, Comandini et al. 2006, Torrej 2007, 2009, Campos et al. 2010, Loizides & Kyriakou 2011, Malloch & Thorn 2011, Zambonelli et al. 2014, Leonardi et al. 2016, Loizides 2016). Russulaceae species associated with Cistaceae are not monophyletic, suggesting that multiple events of host specialization took place in the evolutionary history of this family. Moreover, speciation pro-cesses through biogeographic isolation (L. josserandii vs L. bor-zianus), host specialization (L. populicola vs L. stephensii), or climatic regions (R. mistiformis vs R. cerea), seem to be un-stable in time, as no apomorphic features based solely on these factors could be identified in any section or subsection. Despite the loss of aerial spore dispersal, however, the majority of lineages do not appear to be endemic to specific continents, such as the widespread European species R. candidissima, which nests within a clade otherwise composed exclusively of American species. These observations are consistent with the ‘generalized diversification rate’ pattern found in Russula by Looney et al. (2016). Morphological features traditionally employed to discriminate between sequestrate Russulaceae species seem to have a phylogenetic basis, although some of them were in need of re-interpretation. In Table 2, all European species studied are compared according to: 1) basidiomata development (pseudoangiocarpic, angiocar-pic) and presence or absence of a stipe (stipitate, sessile); 2) presence or absence of a columella (percurrent, branched, stipe-columella); 3) suprapellis structure (trichoderm, intricate trichoderm, palisadotrichoderm, oedotrichoderm, trichoepithelium); 4) hymenophoral trama composition (heteromerous or ho-moiomerous); 5) presence or absence of cystidia (cystidioles, macrocystidia); 6) number of spores in basidia; 7) size of spores; 8) spore ornamentation (warts, reticulum, ridges, crests); and finally 9) family of putative symbiont plants (Cistaceae, Fagaceae, Pinaceae, Salicaceae). The degree of development of the sequestrate syndrome (gasteromycetation) is indeed not only poorly correlated with classical taxonomical concepts, as already evident in other groups (Hibbett et al. 1997, Peintner et al. 2001, Hosaka et al. 2006), but probably variable within some species, and maybe even between basidiomes from the same mycelium. Hymenophoral trama, suprapellis structure, and number of spores per basidia seem to be features characteristic of supraspecific clades, in concordance with observations by Calonge & Martín (2000) and Lebel & Trappe (2000), and so is ectomycorrhizal morphology (Beenken 2004), while spore size, shape, and ornamentations, ecological traits, as well as classical macroscopic features (size, colour, odor, spore print), are still useful at the species level. Acknowledgements The authors want to express their gratitude to Antoni Sánchez-Cuxart, curator of Herbarium BCN (Barcelona University, CeDocBiV, Barcelona), for his help with loan requests. To the directors and curators of the following herbaria: Universidad de Alcalá, Alcalá de Henares (AH), Asociaci Vallisoletana de Micologia, Valladolid (AVM), U.S. National Fungus Collections, Systematic Mycology and Microbiology Laboratory, Beltsville (BPI), Harvard University, Farlow Reference Library and Herbarium of Cryptogamic Botany, Cambridge (FH), University of Haifa, Institute of Evolution, Haifa (HAI), Leopold Franzens Universität, Institut f Mikrobiologie, Innsbruck (IB), Royal Botanic Gardens, Kew (K), Jagiellonian University, Institute of Botany, Krak (KRA), Botanische Staatssammlung, Mchen (M), Real Jardín Botánico, CSIC, Madrid (MA), Museo di Storia Naturalle di Venezia, Venezia (MCVE), Institut de Botanique, Montpellier (MPU), Muséum d’Histoire Naturelle de Nice, Nice (NICE), New York Botanical Garden, New York (NY), Muséum National d’Histoire Naturelle, Laboratoire de Cryptogamie, Paris (PC), National Museum, Mycological Department, Praha (PRM), Uppsala Universitet, Museum of Evolution, Botany, Uppsala (UPS) and University of Vienna, Institute of Botany, Wien (WU). The authors also thank the following collaborators for kindly providing collections and/or images or valuable information data: Carlo Agnello for the Italian collection of R. messapica, José Luis Becerra for the Spanish collections and photographs of R. andaluciana, Manuel Becerra for the Spanish collections of R. ammophila and R. vinaceodora, Alyona Biketova, Zohar Shafranov, Anatoliy Fedorenko, T. Pavlichek, S. Reshetnikov, I. Duckman, I. Shams and Y. Ur for providing collections, photographs and data of R. galileensis, José Antonio Cadinos for the Spanish collections of R. amoenolens, T. Choko-va for the Bulgarian collection of L. populicola, Paolo Fantini for the Italian collections of R. mistiformis, Giorgos Fransouas for the Greek collection and photographs of R. mistiformis, Aurelio García-Blanco, Miguel Sanz-Carazo and J.B. Del Val for the Spanish collection of L. giennensis, Faustino García for the Spanish collections and photographs of L. josserandii, R. cerea and R. mistiformis, Matteo Gelardi (MG), Maria Tulli and R. Polverini for the Italian collections and photographs of R. mediterraneensis, Celestino Gelpi and Julián Muz for the Spanish collections and photographs of R. andaluci-ana, Javier Gez for helping in collecting R. meridionalis, Lamberto Gori† (ELG) and Giampaolo Bernardini for the Italian collections of L. populicola, R. candidissima and R. cerea, Gerhard Gross† for the German collection of R. candida, Gunnar Hensel (GH) and B. Mler for the German collection and photographs of R. cerea, Pablo Juste for the Spanish collection of L. giennensis, Istvan Király for the Hungarian collection of L. stephensii, Maciej Kozak for the Polish collection of R. cerea, Claude Lavoise for the photographs of L. soehneri, Zoltán Lukács for the Hungarian collections of R. candida, José Angel Martínez and Andreu Carreres for the Spanish collections and photographs of R. monospora, Amer Montecchi (AM) for the Italian collections of L. borzianus, L. populicola and R. mattiroloana, Marco Morara and Maurizio Berbenni for the Italian collections and photographs of L. soehneri and L. stephensii, Pierre-Arthur Moreau (PAM) for the French collections and images of L. borzianus and for his help in consulting the original material of H. monosporum from Nice, Pedro María Pasabán and Francisco Sáinz for the Spanish collections and photographs of R. candidis-sima, R. cerea and R. pila, Miquel Angel Pérez-De-Gregorio for the Spanish collections of R. candida and R. candidissima, Tomás Pérez-Jarauta for the Spanish collections of L. giennensis, Vasilis Ramoutsakis for the Greek collection of R. messapica var. messapicoides, Fernando Rodríguez for his friendship and invaluable help in many Catalan collections, Ryszard Rutkowski for the Polish collection of L. stephensii, A. Sanz-Becerra and Natalio Redondo for the Spanish collection of R. monospora, Miguel Ángel Sanz for the Spanish collections of L. soehneri, Mario Sarasini (MS) for the Italian collections of R. candida and R. cerea, George Setkos for the Greek collections of L. populicola, Josep Lleonard Siquier for images and information of the Spanish collections of R. messapica and M. Tordelli and Despina Klisiari for the Cypriot collections of R. hobartiae. Finally the authors thank Caroline Loup, curator of the Herbier of the Université de Montpellier (SPH, MPU) for kindly allowing us to reproduce the original plates of Zelleromyces josserandii of G. Malençon, Olivier Gerriet, curator of the Muséum d’Histoire Naturelle de Nice (NICE), for kindly providing us information, images and original material of Hydnangium monosporum from Barla’s Herbarium, Prof. Solomon P. Wasser for providing access to HAI Herbarium and Carles Roqué and the Unitat de Microscia dels Serveis Tècnics de Recerca of the Universitat de Girona (STR, UdG) for their invaluable help with SEM imaging. Last but not least, the authors are grateful to Caroline Hobart for her help and valuable advice on sequestrate material from Cyprus, particularly R. hobartiae and also to Angelos Papadimitriou for translating into English the Greek description of L. populicola. REFERENCES Adamčik S, Caboň M, Eberhardt U, et al. 2016. Amolecular analysis reveals hidden species diversity within the current concept of Russula maculata (Russulaceae, Basidiomycota). Phytotaxa 270: 71–88. Adamčik S, Jančovičová S, Buyck B. 2017. The Russulas described by Charles Horton Peck. Cryptogamie, Mycologie 39: 3–108. Altschul SF, Madden TL, Schäffer AA, et al. 1997. Gapped BLAST and PSIBLAST: a new generation of protein database search programs. Nucleic Acids Research 25: 3389–3402. Alvarado P, Moreno G, Manj JL. 2012. Comparison between Tuber gennadii and T. oligospermum lineages reveals the existence of the new species T. cistophilum (Tuberaceae, Pezizales). Mycologia 104: 894–910. Astier J, Pacioni G. 1998. Le genre Martellia Matt. en Europe. Documents Mycologiques 109–110: 7–9. Avis PG. 2012. Ectomycorrhizal iconoclasts: the ITS rDNA diversity and nitrophilic tendencies of fetid Russula. Mycologia 104: 998–1007. Bahram M, Polme S, Koljalg U, et al. 2011. A single European aspen (Populus tremula) tree individual may potentially harbour dozens of Cenococcum geophilum ITS genotypes and hundreds of species of ectomycorrhizal fungi. FEMS Microbiology Ecology 75: 313–320. Bahram M, Polme S, Koljalg U, et al. 2012. Regional and local patterns of ectomycorrhizal fungal diversity and community structure along an altitudinal gradient in the Hyrcanian forests of northern Iran. New Phytologist 193: 465–473. Bandala VM, Montoya L, Ramos A. 2016. Two new Lactarius species in a subtropical cloud forest in eastern Mexico. Mycologia 108: 967–980. Barge EG, Cripps CL, Osmundson TW. 2016. Systematics of the ectomycorrhizal genus Lactarius in the Rocky Mountain alpine zone. Mycologia 108: 414–440. Bazzicalupo AL, Buyck B, Saar I, et al. 2017. Troubles with mycorrhizal mushroom identification where morphological differentiation lags behind barcode sequence divergence. Taxon 66: 791–810. Beaton G, Pegler DN, Young TWK. 1984. Gasteroid Basidiomycota of Victoria State, Australia. 2. Russulales. Kew Bulletin 39: 669–698. Beenken L. 2004. Les ectomycorrhizes du genre Russula. Bulletin Trimestriel de la Société Mycologique de France 120: 293–333. Beenken L, Sainge MN, Kocyan A. 2016. Lactarius megalopterus, a new angiocarpous species from tropical rainforest in Central Africa, shows adaptations to endozoochorous spore dispersal. Mycological Progress 15: 1–10. Berkeley MJ. 1844. Notices of British fungi (257–322). Annals and Magazine of Natural History, Series 1: 340–360. Binder M, Hibbett DS. 2002. Higher-level phylogenetic relationships of Homo-basidiomycetes (mushroom-forming fungi) inferred from four rDNA regions. Molecular Phylogenetics & Evolution 22: 76–90. Bon M. 1986. Novitates. Validations et taxons nouveaux. Documents Myco-logiques 65: 51–56. Bon M. 1988. Clé monographiquedes russules dʼEurope. Documents Myco- logiques 70–71: 1–120. Boudier E. 1906. Icones mycologicae ou iconographie des champignons de France, principalement Discomycètes. Tome 1. Librairie des Sciences Naturelles Paul Klincksieck, Paris, France. Boudier E, Patouillard N. 1888. Note sur deux nouvelles espèces de Cham-pignons des environs de Nice. Journal de Botanique, Paris 2: 445–446. Bougher NL. 1997. Three new sequestrate Basidiomycetes from Western Australia. Mycotaxon 63: 37–48. Bougher NL, Lebel T. 2001. Sequestrate (truffle-like) fungi of Australia and New Zealand. Australian Systematic Botany 14: 439–484. Bougher NL, Tommerup IC, Malajczuk N. 1993. Broad variation in developmental and mature basidiome morphology of the ectomycorrhizal fungus Hydnangium sublamellatum sp. nov. bridges morphologically based generic concepts of Hydnangium, Podohydnangium and Laccaria. Mycological Research 97: 613–619. Brock PM, Doring H, Bidartondo MI. 2009. How to know unknown fungi: the role of a herbarium. New Phytologist 181: 719–724. Bucholtz F. 1901. Hypogaeen aus Russland. Hedwigia 40: 304–322. Bucholtz F. 1903. Zur Morphologie und Systematik der Fungi hypogaei. Annales Mycologici 1: 152–174, taf. IV, V. Buyck B. 1995. A global integrated approach on the taxonomy of Russulales. Russulales News 3: 2–17. Buyck B, Duhem B, Das K, et al. 2017. Fungal Biodiversity Profiles 21–30. Cryptogamie, Mycologie 38: 101–146. Buyck B, Hofstetter V, Eberhardt U, et al. 2008. Walking the thin line between Russula and Lactarius: the dilemma of Russula subsect. Ochricompactae. Fungal Diversity 28: 15–40. Buyck B, Hofstetter V, Verbeken A, et al. 2010. Proposal 1919: To conserve Lactarius nom. cons. (Basidiomycota) with a conserved type. Taxon 59: 295–296. Buyck B, Ovrebo CL. 2002. New and interesting Russula species from Panama. Mycologia 94: 888–901. Buyck B, Verbeken A, Eberhardt U. 2007. The genus Lactarius in Madagas-car. Mycological Research 111: 787–798. Cabero J. 2011. Segunda cita para la península ibérica de Gymnomyces ilicis. Boletín Micolico de la FAMCAL 6: 63–68. Calonge FD. 2000. Validation or confirmation of some new taxa recently published. Boletín de la Sociedad Micolica de Madrid 25: 301–302. Calonge FD, Martín MP. 2000. Morphological and molecular data on the taxonomy of Gymnomyces, Martellia and Zelleromyces (Russulales). Mycotaxon 76: 9–15. Calonge FD, Martín MP. 2003. Zelleromyces hispanicus, the gasteroid phase of Lactarius aurantiacus. Boletín de la Sociedad Micolica de Madrid 27: 231–236. Calonge FD, Pasabán PM. 2005. Macowanites candidus (Russulales), nuevo para el catálogo micolico espal. Boletín de la Sociedad Micolica de Madrid 29: 87–90. Calonge FD, Pegler DN. 1998. Zelleromyces hispanicus sp. nov. (Russulales, Elasmomycetaceae), an orange-red species possibly related to Lactarius aurantiacus. Cryptogamie, Mycologie 19: 99–105. Calonge FD, Vidal JM. 1999. Gymnomyces ammophilus Vidal & Calonge, sp. nov., encontrado en Portugal. Boletín de la Sociedad Micolica de Madrid 24: 65–70. Calonge FD, Vidal JM. 2001. Macowanites vinaceodorus sp. nov. (Russulales), a new gasteroid fungus from coastal dunes of Spain. Mycotaxon 79: 1–6. Campos JC, Zamora JC, Vila J. 2010. Estudio de la micobiota de las comunidades de Cistaceae en el centro de la Península Ibérica. Boletín de la Sociedad Micolica de Madrid 34: 257–270. Castresana J. 2000. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Molecular Biology and Evolution 17: 540–552. Cavara F. 1897. Contributo alla conoscenza delle Podaxineae: Elasmomyces mattirolianus nov gen. et sp. Malpighia 11: 414–428, tav. VII. Cavara F. 1900. Arcangeliella borziana nov. gen. nov. sp. Nuova Imenogaste-rea delle abetine di Vallombrosa. Nuovo Giornale Botanico Italiano, Nuova Serie, 7: 117–128, tav. VII. Cheype J-L, Campo E. 2012. Russula rubropunctatissimia Cheype & Campo une nouvelle russule découverte en Guyane Française. Bulletin Trimestriel de la Société Mycologique de France 128: 127–135. Clémençon H. 2004. Cytology and plectology of the Hymenomycetes. Biblio-theca Mycologica 199. Cramer, Berlin-Stuttgart, Germany. Comandini O, Contu M, Rinaldi AC. 2006. An overview of Cistus ectomycorrhizal fungi. Mycorrhiza 16: 381–395. Corner EJH, Hawker LE. 1953. Hypogeous fungi from Malaya. Transactions of the British Mycological Society 36: 125–137. Crous PW, Wingfield MJ, Richardson DM, et al. 2016. Fungal Planet de-scription sheets 400–468. Persoonia 36: 316–458. Das K, Dowie NJ, Li GJ, et al. 2014. Two new species of Russula (Russulales) from India. Mycosphere 5: 612–622. Das K, Ghosh A, Chakraborty D, et al. 2017. Fungal Biodiversity Profiles 31–40. Cryptogamie, Mycologie 38: 353–406. De Crop E, Nuytinck J, Van de Putte K, et al. 2017. A multi-gene phylogeny of Lactifluus (Basidiomycota, Russulales) translated into a new infrageneric classification of the genus. Persoonia 38: 58–80. Desjardin D. 2003. A unique ballistosporic hypogeous sequestrate Lactarius from California. Mycologia 95: 148–155. Dodge CW. 1931. Alpova, a new genus of Rhizopogonaceae, with further notes on Leucogaster and Arcangeliella. Annals of the Missouri Botanical Garden 18: 457–464. Dodge CW, Zeller SM. 1937 ‘1936’. Hydnangium and related genera. Annals of the Missouri Botanical Garden 23: 565–598. Dring DM, Pegler DN. 1978. New and noteworthy gasteroid relatives of the Agaricales from tropical Africa. Kew Bulletin 32: 563–569. Durrall DM, Gamiet S, Simard SW, et al. 2006. Effects of clearcut logging and tree species composition on the diversity and community composition of epigeous fruit bodies formed by ectomycorrhizal fungi. Canadian Journal of Botany 84: 966–980. Dutta AK, Paloi S, Pradhan P, et al. 2015. A new species of Russula (Russulaceae) from India based on morphological and molecular (ITS sequence) data. Turkish Journal of Botany 39: 850–856. Eberhardt U. 2002. Molecular kinship analyses of the agaricoid Russulaceae: correspondence with mycorrhizal anatomy and sporocarp features in the genus Russula. Mycological Progress 1: 201–223. Eberhardt U, Oberwinkler F, Verbeken A, et al. 2000. Lactarius ectomy corrhizae on Silver fir (Abies alba): morphological description, molecular characterization, and taxonomic remarks. Mycologia 92: 860–873. J.M. Vidal et al.: Revision of sequestrate Russulaceae Eberhardt U, Verbeken A. 2004. Sequestrate Lactarius species from tropical Africa: L. angiocarpus nov. sp. and L. dolichocaulis comb. nov. Mycological Reseach 108: 1042–1052. Elliott TF, Trappe JM. 2018. Aworldwide nomenclature revision of sequestrate Russula species. Fungal Systematics and Evolution 1: 229–242. Fellner R, Biber J. 1990. Helianthemum and some agaricales: unusual case of ectomycorrhizal symbiosis. Agriculture, Ecosystems & Environment 28: 121–125. Fogel R, States J. 2001. Materials for a hypogeous mycoflora of the Great Basin and adjacent cordilleras of the Western United States IV: Zelleromy-ces rogersonii, sp. nov. (Basidiomycota, Elasmomycetaceae). Mycotaxon 80: 321–326. Fraiture A, Derboven P. 2009. Deux nouvelles récoltes belges d’Arcangeliella stephensii, espèce tubéroïde proche des lactaires. Revue du Cercle de Mycologie de Bruxelles 9: 27–42. Frank JL, Barry S, Southworth D. 2006. Mammal mycophagy and dispersal of mycorrhizal inoculum in Oregon white oak woodlands. Northwest Science 80: 264–273. Garay-Serrano E, Bandala VM, Montoya L. 2012. Morphological and mole-cular identification of the ectomycorrhizal association of Lactarius fumosibrunneus and Fagus grandifolia var. mexicana trees in eastern Mexico. Mycorrhiza 22: 583–588. Gardes M, Bruns TD. 1993. ITS primers with enhanced specificity for Basi-diomycetes —Application to the identification of mycorrhizae and rusts. Molecular Ecology 2: 113–118. Ge ZW, Smith ME, Zhang QY, et al. 2012. Two species of the Asian endemic genus Keteleeria form ectomycorrhizas with diverse fungal symbionts in southwestern China. Mycorrhiza 22: 403–408. Gelardi M, Vizzini A, Ercole E, et al. 2015. Circumscription and taxonomic arrangement of Nigroboletus roseonigrescens gen. et sp. nov., a new mem-ber of Boletaceae from tropical South–Eastern China. PLoS One 10: e0134295. Ghosh A, Das K, Bhatt RP. 2017. Russula sarnarii sp. nov. (Russulaceae, Basidiomycota) from Indian Himalaya. Current Research in Environmental and Applied Mycology 7: 64–72. Gladish S, Frank J, Southworth D. 2010. The serpentine syndrome below ground: ectomycorrhizas and hypogeous fungi associated with conifers. Canadian Journal of Forest Research 40: 1671–1679. Gori L. 2005. Funghi Ipogei della Lucchesia, di altre province italiane e dall’estero. Maria Pacini Fazzi Editore, Lucca, Italy. Gross G. 1969. Ein Saarländischer fund von Elasmomyces mattirolianus Cav. Zeitschrift f Pilzkunde, N.F., 34: 27–32. Hawker LE. 1952. Hypogeous fungi. II. A new variety of Hydnangium car-neum Wallr. from North Wales. III. Three new British records: Gautieria morchellaeformis Vitt., Hymenogaster hessei Soehner, and Elaphomyces aculeatus Vitt. Transactions of the British Mycological Society 35: 279–284. Hawker LE. 1954. British hypogeous fungi. Philosophical Transactions of the Royal Society London, Series B, 237: 429–546. Hawksworth DL, Kirk PM, Sutton BC, et al. 1995. Ainsworth & Bisbyʼs dictio-nary of the fungi. 8th edition. CAB International, Oxon, U.K. Heilmann-Clausen J, Verbeken A, Vesterholt J. 1998. The genus Lactarius. Fungi of Northern Europe. Vol. 2. The Danish Mycological Society, Copenhagen, Denmark. Heim R. 1938. Les Lactario-russulés du domaine oriental de Madagascar. Essai sur la classification et la phylogénie des Astérosporales. Prodrome à une flore mycologique de Madagascar et dépendances. I. Laboratoire de Cryptogamie du Muséum National d’Histoire Naturelle de Paris, Paris, France. Heim R. 1959. Une espèce nouvelle de Gastrolactarié en Thaïlande. Revue de Mycologie, Paris 24: 93–102. Heim R, Font-Quer P, Codina J. 1934. Fungi Iberici. Observations sur la flore mycologique Catalane. Treballs del Museu de Ciències Naturals de Barcelona, Sèr. Bot. 15: 1–146, 4 pl. Hesler LR, Smith AH. 1979. North American species of Lactarius. The University of Michigan Press, Ann Arbor, Michigan, USA. Hibbett DS, Pine EM, Langer E, et al. 1997. Evolution of gilled mushrooms and puffballs inferred from ribosomal DNA sequences. Proceedings of the National Academy of Sciences 94: 12002–12006. Hibbett DS, Tsuneda A, Murakami S. 1994. The secotioid form of Lentinus tigrinus: genetics and development of a fungal morphological innovation. American Journal of Botany 81: 466–478. Higgins D, Thompson J, Gibson T, et al. 1994. CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Research 22: 4673–4680. Horton BM. 2011. Eucalypt decline and ectomycorrhizal fungal community ecology of Eucalyptus delegatensis forest, Tasmania, Australia. PhD thesis, University of Tasmania. Hosaka K, Bates ST, Beever RE, et al. 2006. Molecular phylogenetics of the gomphoid-phalloid fungi with an establishment of the new subclass Phallomycetidae and two new orders. Mycologia 98: 949–959. Izzo AD, Meyer M, Trappe JM, et al. 2005. Hypogeous ectomycorrhizal fungal species on roots and in small mammal diet in a mixed-conifer forest. Forest Science 51: 243–254. Josserand M. 1983. La description des champignons supérieus. 2e éd. Editions Lechevalier, Paris, France. Kalchbrenner C. 1876. Zwei neue Pilzgattungen. Hedwigia 15: 115–116. Kalchbrenner C. 1882. Fungi Macowaniani. Grevillea 10: 104–109. Kleine CS, McClean T, Miller SL. 2013. Genetic divergence among disjunct populations of three Russula spp. from Africa and Madagascar. Mycologia 105: 80–89. Kong A, Cifuentes J, Estrada-Torres A, et al. 2015. Russulaceae associated with mycoheterotroph Monotropa uniflora (Ericaceae) in Tlaxcala, Mexico: a phylogenetic approach. Cryptogamie, Mycologie 36: 479–512. KonstantinidisG.2009.Μανιτάρια,φωτογραφικόςοδηγόςμανιταροσυλλέκτη (Mushrooms, a photographic collectorʼs guide). Published by the author, Athens, Greece. Kornerup A, Wanscher JH. 1978. Methuen handbook of colour. 3 ed. Eyre Methuen, London, UK. Lago M, Bougher NL, Castro ML. 2000. Morphological variability and implications for definitionof taxaintheDescolea-Setchelliogaster-Descomyces complex. Mycotaxon 18: 37–57. Larsson E, Larsson KE. 2003. Phylogenetic relationships of russuloid basidiomycetes with emphasis on aphyllophoraean taxa. Mycologia 95: 1037– 1065. Lavorato C.1991.Chiave analitica e note bibliografiche della micoflora del cisto. Bolletino dellʼAssociazione Micologica ed Ecologica Romana 24: 16–45. Lebel T. 1998. Taxonomic revision of the sequestrate relatives of Russula from Australia and New Zealand. PhD Thesis, Oregon State University, USA. Lebel T. 2001. A new species of Zelleromyces (Russulales) from Australia. Australasian Mycologist 20: 4–8. Lebel T. 2002. Sequestrate Russulales of New Zealand: Gymnomyces and Macowanites. New Zealand Journal of Botany 40: 489–509. Lebel T. 2003a. Australian sequestrate (truffle-like) fungi. XIII. Cystangium (Russulales, Basidiomycota). Australian Systematic Botany 16: 371–400. Lebel T. 2003b. Australian sequestrate (truffle-like) fungi. XIV. Gymnomyces (Russulales, Basidiomycota). Australian Systematic Botany 16: 401–426. Lebel T. 2017. Nomenclatural changes and corrections for some previously described Australasian truffle-like fungi (Basidiomycetes). Muelleria 36: 8–14. Lebel T, Castellano MA. 2002. Type studies of sequestrate Russulales II. Australian and New Zealand species related to Russula. Mycologia 94: 327–354. Lebel T, Dunk CW, May TW. 2013. Rediscovery of Multifurca stenophylla (Berk.) T. Lebel, C.W. Dunk & T.W. May comb. nov. (Russulaceae) from Australia. Mycological Progress 12: 497–504. Lebel T, Tonkin JE. 2007. Australasian species of Macowanites are sequestrate species of Russula (Russulaceae, Basidiomycota). Australian Systematic Botany 20: 355–381. Lebel T, Trappe JM. 2000. Type studies of sequestrate Russulales. I. Generic type species. Mycologia 92: 1188–1205. Lee H, Park MS, Jung PE, et al. 2017. Re-evaluation of the taxonomy and diversity of Russula section Foetentinae (Russulales, Basidiomycota) in Korea. Mycoscience 58: 351–360. Lejeune C. 2005. Russula straminea ou Russula globispora? Annales Con-federationis Europeae Mycologiae Mediterranensis 2003: 61–72. Leonardi M, Comandini O, Rinaldi AC. 2016. Peering into the Mediterranean black box: Lactifluus rugatus ectomycorrhizas on Cistus. IMA Fungus 7: 275–284. Li G-J, Zhang C-L, Zhao R-L, et al. 2018. Two new species of Russula from Northeast China. Mycosphere 9: 431–443. Li G-J, Zhao D, Li S-F, et al. 2015. Russula chiui and R. pseudopectinatoides, two new species from southwestern China supported by morphological and molecular evidence. Mycological Progress 14: 33. Li G-J, Zhao Q, Zhao D, et al. 2013. Russula atroaeruginea and R. sichuanensis spp. nov. from southwest China. Mycotaxon 124: 173–188. Liu JK, Hyde KD, Jones EBG, et al. 2015. Fungal diversity notes 1–110: Taxonomic and phylogenetic contributions to fungal species. Fungal Diversity 72: 1197. Llistosella J. 1998. Algunes espècies del gènere Russula de Catalunya i les Illes Balears. 2a Contribuci Revista Catalana de Micologia 21: 75–92. Llistosella J, Vidal JM. 1995. Due nuove specie di Russulales gasteroidi della regione mediterranea. Rivista di Micologia AMB 38: 149–162. Loizides M. 2016. Macromycetes within Cistaceae-dominated ecosystems in Cyprus. Mycotaxon 131: 255–256. Loizides M, Kyriakou T. 2011. Fungi of the Cistus maquis. Field Mycology 12: 14–22. Looney BP, Ryberg M, Hampe F, et al. 2016. Into and out of the tropics: global diversification patterns in a hyper-diverse clade of ectomycorrhizal fungi. Molecular Ecology 25: 630–647. Lutzoni F, Kauff F, Cox CJ, et al. 2004. Assembling the fungal tree of life: progress, classification, and evolution of subcellular traits. American Journal of Botany 91: 1446–1480. Mader K, Mader A. 1992. Ein Beitrag zur Kenntnis der sternsporigen Hypogäen. Österreichische Zeitschrift f Pilzkunde 1: 3–10. Malençon G. 1931. La série des Astérosporés. Recueil de travaux dédiés à L. Mangin: 377–396, 1 pl. Malençon G. 1975. Champignons hypogés du Nord de l’Afrique II. Basidiomycetes. Revue de Mycologie, Paris 39: 279–306. Malloch D, Thorn RG. 2011. The occurrence of ectomycorrhizae in some species of Cistaceae in North America. Canadian Journal of Botany 63: 872–875. Martín MP, Hberg N, Llistosella J. 1999. Macowanites messapicoides, a hypogeous relative to Russula messapica. Mycological Research 103: 203–208. Martín MP, Moreno G. 2001. Molecular data confirm Setchelliogaster tenuipes and S. rheophyllus as Cortinariales. Mycotaxon 78: 257–263. Martín MP, Rocabruna A. 1999. The taxonomic boundaries between Naucoria rheophylla and Setchelliogaster tenuipes based on morphology and molecular data. Mycotaxon 71: 141–148. Massee G. 1898. Fungi Exotici, I. Bulletin of Miscellaneous Information, Royal Botanic Gardens, Kew 1898: 113–136. Matheny PB. 2005. Improving phylogenetic inference of mushrooms with RPB1 and RPB2 nucleotide sequences (Inocybe; Agaricales). Molecular Phylogenetics and Evolution 35: 1–20. Matheny PB, Wang Z, Binder M, et al. 2007. Contributions of rpb2 and tef1 to the phylogeny of mushrooms and allies (Basidiomycota, Fungi). Molecular Phylogenetics and Evolution 43: 430–451. Mattirolo O. 1900. Gli ipogei di Sardinia e di Sicilia. Malpighia 14: 39–110, tav. I. Melera S, Ostellari C, Roemer N, et al. 2017. Analysis of morphological, ecological and molecular characters of Russula pectinatoides Peck and Russula praetervisa Sarnari, with a description of the new taxon Russula recondita Melera & Ostellari. Mycological Progress 16: 117–134. Miller SL, Buyck B. 2002. Molecular phylogeny of the genus Russula in Eu- ropewith a comparison of modern infrageneric classifications. Mycological Research 106: 259–276. Miller SL, Larsson E, Larsson KH, et al. 2006. Perspectives in the new Russulales. Mycologia 98: 960–970. Miller SL, Lebel T. 1999. Hypogeous fungi from the South-eastern United States. II. The genus Zelleromyces. Mycotaxon 72: 15–25. Miller SL, McClean TM, Walker JF, et al. 2001. A molecular phylogeny of the Russulales including agaricoid, gasteroid and pleurotoid taxa. Mycologia 93: 344–354. Montecchi A, Lazzari G. 1986. Elasmomyces mattirolianus Cav. (= Maco-wanites kriukowensis (Buch.) Sing. & Smith), forme di transizione tra Gasteromiceti e Russulaceae. Bollettino del Gruppo Micologico G. Bresadola Trento 29: 157–163. Montecchi A, Sarasini M. 2000. Funghi Ipogei d’Europa. Associazione Micologica Bresadola, Fondazione Centro Studi Micologici, Trento-Vicenza, Italy. Moreno G, Galán R, Montecchi A. 1991. Hypogeous fungi from peninsular Spain. II. Mycotaxon 42: 201–238. Moreno-Arroyo B, Gez J, Calonge FD. 1998a. Zelleromyces giennensis sp. nov. (Russulales), a gasteroid fungus from the South of Spain. Crypto-gamie, Mycologie 19: 107–111. Moreno-Arroyo B, Gez J, Calonge FD. 1998b. Zelleromyces meridionalis (Russulales, Elasmomycetaceae), a new species from Spain. Mycotaxon 69: 467–471. Moreno-Arroyo B, Gez J, Calonge FD. 1999. Gymnomyces dominguezii sp. nov. from Spain. Mycological Research 103: 215–218. Moreno-Arroyo B, Llistosella J, Romero De La Osa L. 2002. Gymnomyces sublevisporus (Russulales), una nueva especie de la regi mediterránea. Revista Catalana de Micologia 24: 179–186. Moser M, Binyamini N, Avizohar-Hershenzon Z. 1977. New and noteworthy Russulales from Israel. Transactions of the British Mycological Society 68: 371–377. Moser M, Jich W, Furrer-Ziogas C. 1994. IV Macowanites 1. In: Moser M, Jich W (eds), Farbatlas der Basidiomyceten 12. Gustav Fischer Verlag, Stuttgart, Germany. Nuytinck J, Verbeken A, Delarue S, et al. 2003. Systematics of European sequestrate lactarioid Russulaceae with spiny spore ornamentation. Belgian Journal of Botany 136: 145–153. Nuytinck J, Verbeken A, Miller SL. 2007. Worldwide phylogeny of Lactarius section Deliciosi inferred from ITS and glyceraldehyde-3-phosphate dehydrogenase gene sequences. Mycologia 99: 820–832. Nygren CM, Edqvist J, Elfstrand M, et al. 2007. Detection of extracellular protease activity in different species and genera of ectomycorrhizal fungi. Mycorrhiza 17: 241–248. Nylander JAA. 2004. MrModeltest v2. Program distributed by the autor. Evolutionary Biology Centre, Uppsala University, Uppsala, Sweden. Oberwinkler F. 1977. Das neue System der Basidiomyceten. In: Frey W, Hurka H, Oberwinkler F (eds), Beiträge zur Biologie der Niederen Pflanzen. Systematik, Stammesgeschichte, Ökologie: 59–105. Gustav Fischer Ver-lag, Stuttgart, Germany. Osmundson TW, Robert VA, Schoch CL, et al. 2013. Filling gaps in biodiver-sity knowledge for macrofungi: contributions and assessment of an herbarium collection DNA barcode sequencing project. PLoS ONE 8: e62419. Park MS, Fong JJ, Lee H, et al. 2013. Delimitation of Russula subgenus Amoenula in Korea using three molecular markers. Mycobiology 41: 191–201. Patouillard N. 1910. Note sur trois espèces d’Hydnangium de la flore du Jura. Bulletin Trimestriel de la Société Mycologique de France 26: 199–204. Patouillard N. 1914. Contribution à la flore mycologique du Jura. Bulletin Tri-mestriel de la Société Mycologique de France 30: 347–354. Paz A, Vidal JM, Lavoise C, et al. 2016. Revisi taxonica del género Octaviania (Boletales) en Europa. Boletín Micolico de la FAMCAL 11: 101–138. Peck CH. 1897. Report of the State Botanist (1896). Annual Report on the New York State Museum of Natural History 50: 77–159. Pegler DN, Spooner BM, Young TWK. 1993. British truffles. A revision of British hypogeous fungi. Royal Botanic Gardens, Kew, UK. Pegler DN, Young TWK. 1979. The gasteroid Russulales. Transactions of the British Mycological Society 72: 353–388, 134 fig. Peintner U, Bougher NL, Castellano MA, et al. 2001. Multiple origins of se-questrate fungi related to Cortinarius (Cortinariaceae). American Journal of Botany 88: 2168–2179. Peter M, Buechler U, Ayer F, et al. 2001. Ectomycorrhizas and molecular phylogeny of the hypogeous russuloid fungus Arcangeliella borziana. Mycological Research 105: 1231–1238. Quélet L. 1875. Les champignons du Jura et des Vosges. IIIe Partie. Mémoires de la Société d’Émulation de Montbéliard 5: 429–556, pl. I–IV. Quélet L. 1886. Enchiridion Fungorum in Europa media et praesertim in Gallia Vigentium. Lutetiae, Doin, France. Rehner SA, Buckley E. 2005. A Beauveria phylogeny inferred from nuclear ITS and EF1-α sequences: evidence for cryptic diversification and links to Cordyceps teleomorphs. Mycologia 97: 84–98. Romagnesi H. 1944. La cystide chez les Agaricacées. Supplément a la Revue de Mycologie, Paris 9: 4–21. Romagnesi H. 1985. Les Russules d’Europe et d’Afrique du Nord. Reprint of 1967 ed. by Cramer, Vaduz, Germany. Ronquist F, Huelsenbeck JP. 2003. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19: 1572–1574. Roy M, Schimann H, Braga-Neto R, et al. 2016. Diversity and distribution of ectomycorrhizal fungi from Amazonian lowland white-sand forests in Brazil and French Guiana. Biotropica 48: 90–100. Saba M, Khalid AN. 2015. Russula sichuanensis and its ectomycorrhizae from Himalayan moist temperate forests of Pakistan. Mycotaxon 130: 629–639. Saccardo PA, Saccardo D. 1905. Hymenomycetae-Laboulbeniomycetae. Supplementum Universale, Pars VI. In: Saccardo PA, Sylloge Fungorum 17. Padova, Italy. Sang X-Y, Li X-D, Wang Y-W, et al. 2016. Four new sequestrate species of Russulaceae found in China. Phytotaxa 289: 101–117. Sarnari M. 1998. Monografia illustrata del genere Russula in Europa. Tomo Primo. A.M.B. Fondazione Centro Studi Micologici, Trento, Italy. Sarnari M. 2005. Monografia illustrata del genere Russula in Europa. Tomo Secondo. A.M.B. Fondazione Centro Studi Micologici, Trento, Italy. Schoch CL, Seifert KA, Huhndorf S, et al. 2012. Fungal Barcoding Con-sortium. Nuclear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for Fungi. Proceedings of the National Academy of Sciences, USA 109: 6241–6246. Sene S, Avril R, Chaintreuil C, et al. 2015. Ectomycorrhizal fungal communities of Coccoloba uvifera (L.) L. mature trees and seedlings in the neotropical coastal forests of Guadeloupe (Lesser Antilles). Mycorrhiza 25: 547–559. Shimono Y, Kato M, Takamatsu S. 2004. Molecular phylogeny of Russulaceae (Basidiomycetes, Russulales) inferred from the nucleotide sequences of nuclear large subunit rDNA. Mycoscience 45: 303–316. Singer R. 1986. The Agaricales in modern taxonomy. Fourth edition. Koeltz Scientific Books, Koenigstein, Germany. Singer R, Smith AH. 1960. Studies on secotiaceous fungi. IX. The Astrogastraceous series. Memoirs of the Torrey Botanical Club 21: 1–212. J.M. Vidal et al.: Revision of sequestrate Russulaceae Smith AH. 1962. Notes on Astrogastraceous fungi. Mycologia 54: 626–639. SmithAH. 1963. NewAstrogastraceous fungi from the Pacific northwest. Mycologia 55: 421–441. Smith ME, Henkel TW, Aime CM, et al. 2011. Ectomycorrhizal fungal diversity and community structure on three co-occurring leguminous canopy tree species in a Neotropical rainforest. New Phytologist 192: 699–712. Smith ME, Henkel TW, Williams GC, et al. 2017. Investigating niche partitioning of ectomycorrhizal fungi in specialized rooting zones of the monodominant leguminous tree Dicymbe corymbosa. New Phytologist 215: 443–453. Smith ME, Trappe JM, Rizzo DM, et al. 2006. Gymnomyces xerophilus sp. nov. (sequestrate Russulaceae), an ectomycorrhizal associate of Quercus in California. Mycological Research 110: 575–582. Soehner E. 1924. Prodromus der Fungi hypogaei Bavariae. Kryptogamische Forschungen, Bayerischen Botanischen Gesellschaft, Mchen 1: 390– 398. Soehner E. 1941. Deutsche Hydnangiaceae. Zeitschrift f Pilzkunde, N.F. 20: 108–119. Song J, Chen J-J, Wang M, et al. 2016. Phylogeny and biogeography of the remarkable genus Bondarzewia (Basidiomycota, Russulales). Scientific Reports 6: 34568. Southworth D. 2016. Distribution and abundance of rare sequestrate fungi in Southwestern Oregon. FY10-15 ISSSSP Project Final Report. Southern Oregon University, Ashland. Stamatakis A. 2006. RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22: 2688–2690. Stubbe D, Nuytinck J, Verbeken A. 2010. Critical assessment of the Lactarius gerardii species complex (Russulales). Fungal Biology 114: 271–283. Stubbe D, Verbeken A. 2012. Lactarius subg. Plinthogalus: the European taxa and American varieties of L. lignyotus re-evaluated. Mycologia 104: 1490–1501. Suz LM, Barsoum N, Benham S, et al. 2014. Environmental drivers of ectomycorrhizal communities in Europe’s temperate oak forests. Molecular Ecology 23: 5628–5644. Tamura K, Peterson D, Peterson N, et al. 2011. MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Molecular Biology and Evolution 28: 2731–2739. Tao K, Chang M-C, Liu B. 1993. New species and new record of hypogeous fungi from China IV. Acta Mycologica Sinica 12: 103–106. Tedersoo L, Hallenberg N, Larsson KH, et al. 2003. Fine scale distribution of ectomycorrhizal fungi and roots across substrate layers including coarse woody debris in a mixed forest. New Phytologist 159: 153–165. Thiers HD. 1984. The secotioid syndrome. Mycologia 76: 1–8. Torrej M. 2007. Contribuci al estudio de los hongos del parque natural de la Serra Calderona y su área de influencia: CastellValència (Espa). I. Jarales (Cistion). Revista Catalana de Micologia 29: 17–28. Torrej M. 2009. A contribution to the study of fungi associated with Cistus spp. in the Sierra Calderona Nature Reserve, Castell-Valencia, Spain. II. Mycologia Balcanica 6: 111–122. Trappe JM, Claridge AW. 2003. Australasian sequestrate (truffle-like) fungi. XV. New species from tree line in the Australian Alps. Australasian Mycologist 22: 27–38. Trappe JM, Claridge AW. 2005. Hypogeous fungi: evolution of reproductive and dispersal strategies through interactions with animals and mycorrhizal plants. In: Dighton J, White JF, Oudemans P (eds), The fungal community: its organization and role in the ecosystem: 613–623. 3rd ed. Taylor & Francis, Boca Raton, USA. Trappe JM, Lebel T, Castellano MA. 2002. Nomenclatural revisions in the sequestrate russuloid genera. Mycotaxon 81: 195–214. Trendel JM, Hampe F, Verbeken A. 2017. Russula vinosoflavescens sp. nov., from deciduous forests of Northern Alsace, France. Mycotaxon 132: 707– 721. Trierveiler-Pereira L, Smith ME, Trappe JM, et al. 2015. Sequestrate fungi from Patagonian Nothofagus forests: Cystangium (Russulaceae, Basidiomycota). Mycologia 107: 90–103. Trimbach J. 1996. Barla et la mycologie. Annales du Muséum d’Histoire Naturelle de Nice 11: 179–256. Tulasne L-R, Tulasne C. 1843. Champignons hypogés de la famille des Lycoperdacées, observés dans les environs de Paris et les départaments de la Vienne et d’Indre-et-Loire. Annales des Sciences Naturelles, Botanique, 2e Série, 19: 373–381, pl. 17. Turland NJ, Wiersema JH, Barrie FR, et al. (eds). 2018. International Code of Nomenclature for algae, fungi, and plants (Shenzhen Code) adopted by the Nineteenth International Botanical Congress Shenzhen, China, July 2017. Regnum Vegetabile 159. Glashten, Koeltz Botanical Books. Turner BL, Nesom GL. 2000. Use of variety and subspecies and new varietal combinations for Styrax platanifolius (Styraceae). SIDA, Contributions to Botany 19: 257–262. Vauras J, Ruotsalainen J, Liimatainen K. 2016. Russula suecica, a new red species from Northern Fennoscandia. Karstenia 56: 5–12. VelenovskJ. 1947. Novitates mycologicae novissimae. Opera Botanica Čechica 4: 1–167, 2 pl. Verbeken A, Hampe F, Wissitrassameewong K, et al. 2014a. A new angiocarpous Lactarius species from Thailand. Phytotaxa 181: 163–170. Verbeken A, Stubbe D, Van De Putte K, et al. 2014b. Tales of the unexpected: angiocarpous representatives of the Russulaceae in tropical South East Asia. Persoonia 32: 13–24. Verbeken A, Walleyn R. 2010. Lactarius. Fungus Flora of Tropical Africa. Nationale Plantentuin België, Meise. Vidal JM. 1991a. Octaviania pila (Pat.) Svrček. Bolets de Catalunya 10: làm. 486. Vidal JM. 1991b. Contribuci al conocimiento de la flora micolica del Baix Empordà y zonas limítrofes (Catalu). IV. Hongos hipogeos (Zygomycoti-na, Ascomycotina y Basidiomycotina). Butlletí de la Societat Catalana de Micologia 14–15: 143–194. Vidal JM. 2004a. Arcangeliella borziana and A. stephensii, two gasteroid fungi often mistaken. A taxonomic revision of Lactarius-related sequestrate fungi. Revista Catalana de Micologia 26: 59–82. Vidal JM. 2004b. Macowanites candidus, a new combination for Hydnangium candidum Tul. et C. Tul. Revista Catalana de Micologia 26: 83–96. Vidal JM. 2004c. The genus Stephanospora Pat., two new combinations. Revista Catalana de Micologia 26: 97–111. Vidal JM, Calonge FD, Martín MP. 2002. Macowanites ammophilus (Russulales), a new combination based on new evidences. Revista Catalana de Micologia 24: 69–74. Vila J, Llimona X. 1999. Els fongs del Parc Natural del Cap de Creus i Serra Verdera (Girona). II. Aproximacial component fgic del Cistion. Revista Catalana de Micologia 22: 95–114. Vila J, Llimona X. 2002. Noves dades sobre el component fgic de les comunitats de Cistus de Catalunya. Revista Catalana de Micologia 24: 74–121. Vila J, Llimona X. 2006. Noves dades sobre el component fgic de les comunitats de Cistus de Catalunya. II. Revista Catalana de Micologia 28: 167–207. Vila J, Llimona X. 2009. Noves dades sobre el component fgic de les comunitats de Cistus de Catalunya. III. Addicions, correccions i claus dʼidentificació. Revista Catalana de Micologia 31: 103–137. Vilgalys R, Hester M. 1990. Rapid genetic identification and mapping of en-zymatically amplified ribosomal DNA from several Cryptococcus species. Journal of Bacteriology 172: 4238–4246. Wang Z, Binder M, Dai Y-C, et al. 2004. Phylogenetic relationships of Sparassis inferred from nuclear and mitochondrial ribosomal DNA and RNA polymerase sequences. Mycologia 96: 1015–1029. Whitbeck KL. 2003. Systematics of Pacific Northwestern species of the genus Gymnomyces inferred from nuclear ribosomal DNA internal transcribed spacer sequences. Masters thesis, Oregon State University, Corvallis. White TJ, Bruns TD, Lee SB, et al. 1990. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, et al. (eds), PCR protocols: a guide to methods and applications: 315–322. Academic Press Inc, New York-San Diego, USA. Wilson AW, Binder M, Hibbett DS. 2011. Effects of gasteroid fruiting body morphology on diversification rates in three independent clades of fungi estimated using binary state speciation and extinction analysis. Evolution 65: 1305–1322. Wisitrassameewong K, Nuytinck J, Hyde KD, et al. 2014. Lactarius subgenus Russularia (Russulaceae) in Southeast Asia: 1. Species with very distant gills. Phytotaxa 158: 23–42. Xie X-D, Liu P-G, Yu F-Q. 2010. Species diversity of russuloid mycorrhizae-forming fungi on Pinus yunnanensis seedlings and the mycorrhizal mor-phology. Acta Botanica Yunnanica 32: 211–220. Zambonelli A, Donnini D, Rana GL, et al. 2014. Hypogeous fungi in Mediterranean maquis, arid and semi-arid forests. Plant Biosystems 148: 392–401. Zeller SM, Dodge CW. 1919. Arcangeliella, Gymnomyces and Macowanites in North America. Annals of the Missouri Botanical Garden 6: 49–59. Zeller SM, Dodge CW. 1935. New species of Hydnangiaceae. Annals of the Missouri Botanical Garden 25: 365–373. Zeller SM, DodgeCW. 1937 ‘1936’. Elasmomyces, Arcangeliella and Macowanites. Annals of the Missouri Botanical Garden 23: 599–638. Zhang B-C, Yu Y-N. 1990. Two new species of gasteroid Russulales from China, with notes on taxonomy of Gymnomyces, Martellia and Zelleromy-ces. Mycological Research 94: 457–462. Zhao Q, Li Y-K, Zhu X-T, et al. 2015. Russula nigrovirens sp. nov. (Russulaceae) from southwestern China. Phytotaxa 236: 249–256.