Zoosyst. Evol. 96 (2) 2020, 345-395 | DO! 10.3897/zse.96.52420 eee BERLIN Phylogenetic relationship of catshark species of the genus Scyliorhinus (Chondrichthyes, Carcharhiniformes, Scyliorhinidae) based on comparative morphology Karla D. A. Soares', Marcelo R. de Carvalho! 1 Laboratorio de Ictiologia, Departamento de Zoologia, Instituto de Biociéncias, Universidade de SGo Paulo, Rua do Matdo, trav. 14, n° 101 Sdo Paulo, 05508-090, Brazil http://zoobank.org/54EC7875-F'785-4263-8254-SEFSE&8 D66B98 Corresponding author: Karla D. A. Soares (karlad.soares@yahoo.com.br) Academic editor: Peter Bartsch # Received 25 March 2020 Accepted 18 May 2020 @ Published 19 June 2020 Abstract The genus Scyliorhinus is part of the family Scyliorhinidae, the most diverse family of sharks and of the subfamily Scyliorhininae along with Cephaloscyllium and Poroderma. This study reviews the phylogenetic relationships of species of Scyliorhinus in the subfamily Scyliorhininae. Specimens of all Scyliorhinus species were examined as well as specimens of four of the 18 species of Cephaloscyllium, two species of Poroderma, representatives of almost all other catshark (scyliorhinid) genera and one proscylliid (Proscyllium habereri). A detailed morphological study, including external and internal morphology, morphometry and meristic data, was performed. From this study, a total of 84 morphological characters were compiled into a data matrix. Parsimony analysis was employed to generate hypotheses of phylogenetic relationships using the TNT 1.1. Proscyllium habereri was used to root the clado- gram. The phylogenetic analysis, based on implied weighting (k = 3; 300 replications and 100 trees saved per replication), resulted in three equally most parsimonious cladograms with 233 steps, with a CI of 0.37 and an RI of 0.69. The monophyly of the subfamily Scyliorhininae is supported as well as of the genus Scyliorhinus, which is proposed to be the sister group of Cephaloscyllium. The phylogenetic relationships amongst Scyliorhinus species are presented for the first time. Key Words Scyliorhinus, catsharks, Scyliorhininae, Cephaloscyllium, Poroderma, phylogeny, morphology Introduction cation of 11 catshark genera to the family Pentanchidae, elevated in rank from subfamily (Compagno 1988a). Ac- Contrasting hypotheses on the classification of catsharks are widespread in literature and divide opinions of many authors (e.g. White 1936, 1937; Compagno 1973, 1988a; Maisey 1984; Nakaya 1975; Iglésias et al. 2005; Human et al. 2006; Naylor et al. 2005, 2012a; Nelson et al. 2016; Weigmann 2016; Weigmann et al. 2018). On the basis of morphological data, Compagno (1988a) proposed that the family Scyliorhinidae is composed of 17 genera, fol- lowing the traditional arrangement for the group (Nakaya 1975; Springer 1979). Posteriorly, Iglésias et al. (2005), analysing molecular data, hypothesised that the family Scyliorhinidae is paraphyletic and proposed the re-allo- cording to Iglésias et al. (2005), both families could be morphologically distinguished by the presence/absence of the supraorbital crest on the neurocranium. Although the paraphyly of Scyliorhinidae has been corroborated by later works (Human et al. 2006; Naylor et al. 2012a, 2012b), recent molecular analysis, including a larger sample of taxa, recovered three different para- phyletic lineages of catsharks instead of two and species of Parmaturus were placed in distinct clades (Naylor et al. 2012a, 2012b). No cladistic analysis considering mor- phological data has been performed to elucidate the phy- logenetic relationships of catshark species and enlarge Copyright Karla D. A. Soares, Marcelo R. de Carvalho. This is an open access article distributed under the terms of the Creative Commons Attribution License (CC BY 4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. 346 our knowledge about the evolution and distribution of morphological characters, such as the supraorbital crest. Thus, we decided to adopt here Scyliorhinidae sensu lato (Compagno 1988a; Weigmann et al. 2018) until a more extensive evaluation of morphological characters is done, thus providing a better definition for Scyliorhinidae sensu stricto and Pentanchidae (Nelson et al. 2016). Compagno (1988a) united the genera Scyliorhi- nus, Cephaloscyllium and Poroderma in the subfamily Scyliorhininae, following Gill (1862) and on the basis of muscle and neurocranial characters. Herman et al. (1990) proposed the same arrangement, based on dental charac- ters. Later, studies using molecular data corroborated the monophyly of the subfamily (Iglésias et al. 2005; Human et al. 2006; Naylor et al. 2012a, 2012b), although diver- gences in phylogenetic relationships amongst its taxa have been observed between morphological and molecular data (cf. Compagno 1988a; Naylor et al. 2012a, 2012b). Doubts concerning the monophyly of the genus Scyliorhinus are found in many works and focus mainly on the relationships amongst S. canicula and its congeners (Springer 1966, 1979; Compagno 1988a). Scyliorhinus canicula presents unique characteristics in the nasoral re- gion, such as the presence of nasoral grooves and anterior nasal flaps very close to each other. Similar features are also found in the catshark genera Ate/omycterus and Hap- loblepharus (Compagno 1988a). These differences would be, according to Springer (1979), sufficiently great and unique to guarantee the allocation of the other species of Scyliorhinus to a distinct genus, as was proposed by Jor- dan & Evermann (1896) and Danois (1913). Compagno (1988a) even suggested the adoption of the name Betas- cyllium Leigh-Sharpe, 1926, if this new arrangement should prove to be necessary. Bell (1993) pointed out the importance of cautiously analysing the characters of the nasoral region and examining a representative number of taxa to better comprehend the evolution of these charac- ters amongst scyliorhinids. Scyliorhinus presents a unique configuration of the la- bial furrows comprised of the absence of an upper furrow concomitant with the presence of a narrow lower furrow (Compagno 1988a). The presence of a projecting flap ventral to and covering the lower labial furrow, cited by some authors as a reliable character to identify species belonging to Scyliorhinus (Bigelow and Schroeder 1948; Springer 1966, 1979), was not considered as synapomor- phy for the genus by Compagno (1988a). Yet, according to some authors (Springer 1979; Compagno 1988a), the labial furrows observed in Poroderma and in some spe- cies of Cephaloscyllium are poorly developed or absent and could be easily confused with the configuration pres- ent in Scyliorhinus species (Compagno 1988a). Detailed descriptions of all Scyliorhinus species, mainly based on external morphology, neurocranium and claspers, were provided in the generic revision of Soares and de Carvalho (2019). The morphological characters raised and analysed in that study, as well as additional morphological characters and broader comparisons with zse.pensoft.net Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus other scyliorhinid and proscyllid genera, are included in the present paper, which aims to provide a phylogenetic hypothesis amongst scyliorhinine species. As mentioned, the most recent phylogenetic hypotheses to infer relation- ships amongst catsharks are based on molecular evidence (Iglésias et al. 2005; Human et al. 2006; Naylor et al. 2012a, 2012b); we set out to provide a phylogenetical ap- praisal, based on a renewed examination of morphologi- cal characters. The main objective of the present study is to clarify the phylogenetic significance of the interspe- cific morphological variation in Scyliorhinus and shed light on the relationships amongst its species and other scyliorhinines. Material and methods Selection of taxa Thirty-five taxa were included as terminals in the phylo- genetic analysis. Species representing the three genera as- signed to the Scyliorhininae by Compagno (1988a) were included. Specimens of all 16 valid species of Scyliorhi- nus were examined (Soares and de Carvalho 2019) and, amongst these, 41 specimens were dissected for anatom- ical investigation corresponding to 14 species of this ge- nus. Specimens of S. comoroensis and S. garmani were not dissected due to the lack of available material for study. Data on meristics, external morphology and inter- nal anatomy of S. cabofriensis, S. haeckelii and S. ugoi were extracted from Soares et al. (2015, 2016). Speci- mens of the other genera of the subfamily Scyliorhininae were examined and dissected, including four of the 18 species of Cephaloscyllium and the two species of Poro- derma. For comparative taxa, we examined Proscyllium habereri (family Proscylliidae) and other representatives of Scyliorhinidae sensu lato (Table 1). Specimens of Bythaelurus were not available for dissection and not in- cluded in the analysis. The other genus not included here, Pentanchus, is only known from two specimens; one is the holotype of P. profundicolus (USNM 70260; in poor preservational condition) and the other is a specimen cit- ed by Nakaya and Séret (2000) (MNHN 1999-0270) that could not be found. In any case, Pentanchus may not be valid (Compagno 1988a). All material examined and col- lection data are listed in Appendix 1. Specimen preparation and characters examined This study was based on the examination of 84 morpho- logical characters (79 qualitative and five quantitative) that included external morphology, branchiomeric and hypobranchial cranial muscles, clasper morphology, der- mal denticles and skeleton. External morphological char- acters were observed directly or with the aid of a stereomi- croscope. Anatomical preparation was performed through manual dissections. For the examination of clasper anato- Zoosyst. Evol. 96 (2) 2020, 345-395 347 Table 1. List of species examined (except Scyliorhinus), data available for each species and institutions where the material is depos- ited. Abbreviations for institutions follow Sabaj (2016). Species examined Scyliorhininae Data available Origin of material Cephaloscyllium isabella External morphology, dermal denticles, musculature, skeleton AMNH, USNM C. sufflans External morphology, dermal denticles, musculature, skeleton, clasper SAIB C. umbratile External morphology, dermal denticles, musculature, skeleton USP C. variegatum Poroderma africanum P. pantherinum Comparative taxa Apristurus longicephalus Asymbolus rubiginosus Atelomycterus fasciatus Aulohalaelurus labiosus Cephalurus cephalus Figaro boardmani Galeus antillensis External morphology, dermal denticles, musculature, skeleton External morphology, dermal denticles, musculature, skeleton, clasper External morphology, dermal denticles, musculature, skeleton, clasper External morphology, dermal denticles, musculature, skeleton, clasper External morphology, dermal denticles, musculature, skeleton, clasper External morphology, dermal denticles, musculature, skeleton, clasper External morphology, neurocranium, clasper External morphology, dermal denticles, External morphology, dermal denticles, External morphology, dermal denticles, musculature, skeleton, clasper musculature, skeleton, clasper musculature, skeleton, clasper UF AMS SAIAB SAIAB HUMZ AMS CSIRO, MZUSP ZMH USNM CSIRO, MZUSP Halaelurus natalensis External morphology, dermal denticles, musculature, skeleton, clasper SAIAB Haploblepharus edwardsii External morphology, dermal denticles, musculature, skeleton, clasper AMNH, BMNH Holohalaelurus regani External morphology, dermal denticles, musculature, skeleton, clasper SAIAB Parmaturus xaniurus External morphology, dermal denticles, musculature, skeleton, clasper CAS Proscyllium habereri External morphology, musculature, skeleton, clasper CAS Schroederichthys saurisqualus External morphology, dermal denticles, musculature, skeleton, clasper UERJ, ZMH my, the left clasper was chosen to study the external mor- phology and the right clasper for the internal anatomy. Neurocrania and musculature of adult specimens were examined through dissection. Skin samples were taken for examination of dermal denticles from the right side of the body above the pectoral fin, below the origin of the first dorsal fin and below the insertion of the second dor- sal fin. Dermal denticles were photographed using scan- ning electron microscopes (DSM 940 and ZEISS SIGMA VP), housed in the Departamento de Zoologia of the Uni- versidade de Sao Paulo. Data of intestinal valves, tooth and vertebral counts were obtained directly from the ex- amined specimens or taken from Compagno (1988a) and other works (Compagno and Stevens 1993a, 1993b; Last et al. 1999; Human 2006a, 2006b, 2007; Gledhill et al. 2008; Last and White 2008; Last et al. 2008; Sato et al. 2008; Nakaya et al. 2013). Radiographs were taken in the Faculdade de Medici- na Veterinaria e Zootecnia da Universidade de Sao Pau- lo (FMVZ-USP) and in the radiology facilities of the following institutions: BMNH, HUMZ, MCZ, NRM, NSMT, USNM and ZMUC (according to Sabaj, 2016). Counts of monospondylous and diplospondylous verte- brae were based on Compagno (1988a). The vertebral centra present in the transition zone between monospon- dylous and diplospondylous vertebrae is generally small- er than the last monospondylous centrum and larger than the diplospondylous one and is included in the counts of monospondylous vertebrae. Terminology for neurocranium and jaws follows Com- pagno (1988a) and Motta and Wilga (1995), respectively. Terminology for gill arches follows de Beer (1937) and Shirai (1992a). Clasper terminology for external anatomy and skeletal components are based on Jungersen (1899) and Compagno (1988a). Terminology for neurocranial, hyoid and hypobranchial musculature follows Huber et al. (2011). Nomenclature for dermal denticles follows Herman et al. (1990) and Cappetta (2012). Character descriptions, related to meristic data, are presented first, followed by characters of external mor- phology, myology, skeleton and clasper. Skeletal char- acters are grouped into character complexes, such as neurocranium, jaws, hyoid and gill arches and pectoral girdle. The number preceding each character in the de- scription corresponds to its number presented in the char- acter matrix. A brief summary of each character and its states is followed by its recovered consistency and reten- tion indices (CI and RI, respectively) which reflect their ACCTRAN optimisations (chosen because it maximises initial homology hypotheses). Multistate qualitative char- acters (6, 22, 43 and 49) and quantitative characters (1-5) were analysed as ordered. Characters were illustrated with photographs and schematic drawings made from digital photographs. Pho- tographs were taken with a digital camera (Canon Power Shot SX610 HS). Characters and their states are indicated by arrows and numbers in the figures. Figures were dig- itised and edited with the aid of Adobe Photoshop CS6. Whenever a character is described in the text for a genus without a species citation, that citation refers only to the species examined in the present study and does not imply that the character is present in all congeners. Phylogenetic procedures Hypotheses of phylogenetic relationships were proposed using the cladistic method formalised by Hennig (1950, 1965, 1966) and operationally detailed in other works (Farris 1969; Nelson and Platnick 1981; Goloboff 1993, 1995, 1999; Goloboff et al. 2006, 2008). Qualitative and quantitative characters were considered in the analysis; zse.pensoft.net 348 values for meristic data were normalised and concat- enated with the other characters (Goloboff et al. 2006). Character polarity was determined by outgroup compari- son (Nixon and Carpenter 1993); the outgroups are com- posed of Proscyllium habereri (Proscylliidae) and taxa from the subfamilies Atelomycterinae, Pentanchinae and Schroederichthyinae. Proscyllium habereri was chosen to root the cladogram, as it was recovered as closely related to scyliorhinids in previous studies (Compagno 1988a; Human et al. 2006; Naylor et al. 2012). The data matrix (Appendix 2) was assembled and analysed with the aid of TNT 1.1 (Goloboff et al. 2003, 2008). Parsimony analysis was performed with implied weighting (k = 3) and the “Traditional Search’ option, using the TBR (tree bisection reconnection) algorithm, with 500 replications and 100 trees retained per replica. A strict consensus cladogram was used to summarise the equally most-parsimonious hypotheses obtained from the different topologies yielded by the analysis. Tree edition was performed with the aid of Figtree version 1.4.3 and Adobe Photoshop CS6. CI and RI values and synapomorphies of the various nodes were obtained from the set of equally most-parsi- monious trees. Relative Bremer support was calculated for each clade using TBR and retaining suboptimal trees by seven steps. Missing entries were used to represent two different instances where characters could not be deter- mined: (1) lack of appropriate study material; and (2) in- applicable character state. For Scyliorhinus comoroensis and S. garmani, it was not possible to extract information on internal anatomical characters (e.g. musculature, neu- rocranium). Adult males of the terminal taxa Cephaloscyl- lium isabella, C. umbratile, C. variegatum and Scyliorhi- nus garmani were not available for dissection and, thus, claspers were not examined for these species. Autapomor- phies were not included in the phylogenetic analysis, but are detailed in the section “Non-informative characters’. The section “Description and character analysis’ pre- sents the description of each character, its variation within the subfamily Scyliorhininae and other taxa of Scyliorhi- nidae. Character optimisation and character transforma- tions are presented in Appendices 4 and 5, respectively. Results Character descriptions and analysis Meristics 1 Counts of monospondylous vertebrae: minimum = 28; maximum = 54. (CI = 26; RI = 43-46). Springer and Garrick (1964) pointed out the relevance of vertebral counts to elucidate phylogenetic relationships in Carcharhinidae and other shark families, stating that the values would increase in less inclusive taxonomic levels. These authors considered only precaudal and cau- dal vertebral counts. Springer (1966, 1979) highlighted zse.pensoft.net Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus the importance of counts of monospondylous vertebrae in distinguishing species of Scyliorhinus distributed in the Western Atlantic. According to our results, Scyliorhi- nus torrei presents the lowest values for counts (30-35) in Scyliorhinus, followed by S. torazame (32-37). Most species of this genus present a similar range, from 38 to 44 vertebrae, whereas higher values were observed for S. capensis (44-46), S. garmani (48), S. meadi (46-48) and S. stellaris (43-47). In Cephaloscyllium spp., val- ues range from 44 to 54 and in Poroderma, from 32 to 46. The lowest value for all taxa was found in Holoha- laelurus regani (28). 2 Counts of diplospondylous vertebrae: minimum = 67; maximum = 131. (CI = 34; RI = 31). Compagno (1988a) reported an increase in counts of total vertebrae from Scyliorhinidae to Carcharhinidae, tndi- cating a possible transformation series for this character in Carcharhiniformes. In the present study, we opted to analyse counts of diplospondylous vertebrae, aiming to exclude the influence of monospondylous vertebrae in counts. In Scyliorhinus spp., counts of diplospondylous vertebrae range from 73 to 97. Similar values were ob- served in species of Cephaloscyllium and Poroderma, ex- cept for C. umbratile (110-131). In the outgroups, counts range from 67 to 131, with the lowest values found in Cephalurus cephalus (67-71). 3 Upper tooth row counts: minimum = 33; maximum = 110. (CI = 41; RI = 56). 4 Lower tooth row counts: minimum = 29; maximum = 102. (CI = 36; RI = 49). Tooth row counts are presented for many species in de- scriptions or taxonomic reviews, but never used as a phy- logenetic character. Regarding the differences between upper and lower jaws, we considered upper and lower tooth row counts as distinct characters. In Scyliorhinus spp., tooth row counts range 33 to 85, considering both jaws; S. torrei (33-42) presented the lowest values and S. capensis and S. torazame (45-81) the greatest ones. Amongst scyliorhinines, Cephaloscyllium umbratile (77-110) presented the highest values. In the outgroups, counts range from 35 to 102 with the lowest values found in Apristurus longicephalus (35-45). 5 Counts of intestinal valves: minimum = 5; maximum = 17. (CI = 37; RI = 42). White (1937) divided intestinal valves into three types, considering the numbers of valves observed: i) 2—4; 11) 5—10; 111) 11-30. The family Scyliorhinidae would be classified in the ‘intermediary’ type, presenting 5—10 valves, although White (1937) reported 16 valves for Atelomycterus. Compagno (1988a) suggested that low variation ranges in counts of intestinal valves in carchar- hiniforms would comprise a useful character for system- Zoosyst. Evol. 96 (2) 2020, 345-395 atic studies. Thus, counts of intestinal valves of examined taxa are considered here and analysed as quantitative characters. Amongst Scyliorhinus species, counts ranged from 6 to 11, with higher values for S. capensis (10-11) and lower for S. torazame and S. torrei (6-7); similar values were found in species of Cephaloscyllium. In Po- roderma spp., higher values (11-13) were found. Other scyliorhinids present a similar range for counts of intesti- nal valves, with the exception of Apristurus spp. (14—20) and Atelomycterus spp. (14-16). Nasoral region 6 Extension of anterior nasal flap: (0) entirely covering excurrent nasal aperture and posterior nasal flap, but not covering the upper lip; (1) partially covering ex- current nasal aperture and not covering posterior nasal flap nor upper lip; (2) entirely covering excurrent ap- erture, posterior nasal flap and upper lip. (ordered; CI = 29; RI = 64). The anterior nasal flap is a triangular or subrectangular structure and is situated medial to the incurrent aper- ture and lateral to the excurrent one. This flap can cover partially or entirely the excurrent aperture and poste- rior nasal flap, which is situated on the posterior bor- der of the excurrent aperture (Compagno 1988a, 1999). Scyliorhinines present a nasal flap that entirely covers the excurrent aperture and posterior nasal flap and is separated from the mouth by a short distance (state 0; Fig. 1); the same condition is found in Aulohalaelurus, Asymbolus, Holohalaelurus, Parmaturus and Proscyl- lium. In S. canicula and S. duhamelii, the anterior na- sal flap is longer and covers the upper lip and laterally the lower jaw, as in Atelomycterus and Haploblepha- rus (state 2; Fig. 1A). In Schroederichthys, Halaelurus, Figaro, Galeus, Apristurus and Cephalurus, the anteri- or nasal flap partially covers the excurrent aperture and ends at a considerable distance from the mouth (state 1; Fig. 1B). 7 Distance between anterior nasal flaps: (0) distant by one-half or more of the width of the flap; (1) distant by less than one-half. (CI = 33; RI = 33). In relation to the distance between anterior nasal flaps, these are separated by one-half or more of the width of the flaps in Cephaloscyllium, Poroderma and Scyliorhi- nus (state 0; Figs 1-3), except in S. canicula and S. du- hamelii. In these species, anterior flaps are separated by a short distance, shorter than one-half of their width, as in Atelomycterus (state 1; Figs 1-3). In the other taxa exam- ined, anterior flaps are separated by a similar to slightly larger distance than the width of the flaps. 8 Configuration of anterior nasal flap: (0) flap consisting of a single structure; (1) flap separated into lateral and medial two portions. (CI = 33; RI = 33). 349 Most scyliorhinids present a single anterior nasal flap. In Poroderma, the anterior nasal flap is divided into two portions (Fig. 2C) as in Cephalurus and Schroederichthys (Fig. 1A). 9 Mesonarial crest: (0) inconspicuous; (1) prominent. (CI = 50; RI = 92). The presence of a mesonarial crest was observed and described by Compagno (1988b) for Scyliorhinus como- roensis. The same structure was also found in the other Scyliorhinus species, Cephaloscyllium and Schroederich- thys (state 1; Figs 1 and 2). In S. stellaris, this crest is well developed and extends beyond the posterior border of the anterior nasal flap; this condition is considered here as an autapomorphy for this species (Soares and de Car- valho 2019). In Poroderma, a nasal barbel is found in the same position as the mesonarial crest (Fig. 2). Compag- no (1988a) proposed a hypothesis of homology between the barbel of Poroderma and the mesonarial crest of Scyliorhinus which was followed by Human et al. (2006). This hypothesis is rejected here; the nasal barbel in Po- roderma is composed by muscle fibres which overlap the external nasal cartilage, whereas, in Scyliorhinus, the me- sonarial crest corresponds to extensions of the external nasal cartilage. Therefore, we considered that Poroderma presents an inconspicuous mesonarial crest as in the other examined taxa (state 0; Fig. 2C). 10 Muscular nasal barbel on anterior nasal flap: (0) ab- sent; (1) present. (CI = 100; RI = 100). The presence of a muscular nasal barbel 1s observed in Poroderma (state 1; Fig. 2C) and its extension varies be- tween the two species of the genus, P. africanum and P. pantherinum. In the latter, the nasal barbel is much longer and reaches the upper lip, while, in the former, it is short- er and distant from the mouth. This barbel originates on the ventromedial surface of each anterior nasal flap and is totally separated from the posterior tip of the nasal flap (Compagno 1988a). Considering other carcharhiniforms, only the genus Furgaleus presents a similar nasal barbel. In Leptocharias, the lateral portion of the nasal flap is well developed and long, but does not form a muscular barbel. The nasal barbel present in some orectolobiforms (Chiloscyllium, Ginglymostoma, Hemiscyllium, Orectol- obus and Stegostoma) originates on the rostral surface, medial and partially anterior to the anterior nasal flap (Compagno 1988a) and has a cartilaginous base (Goto 2001), differing from the condition observed in Poroder- ma and Furgaleus. 11 Posterior nasal flap: (0) present; (1) absent. (CI = 33; RI=0). A posterior nasal flap, associated with the excurrent nasal aperture, is present in all scyliorhinines, most scyliorhi- nids and Proscyllium (state 0; Fig. 2). This flap is absent zse.pensoft.net 350 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus Figure 1. Ventral view of the head. A, Scyliorhinus canicula, MNHN 1999-1732, female, 418.5 mm TL; B, Schroederich- thys saurisqualus, UERJ uncatalogued, female, 564 mm TL; C, Holohalaelurus regani, SAIAB 25717, male, 610 mm TL. anf, anterior nasal flap; Ilf, lower labial furrow; mrd, mesonarial crest; ulf, upper labial furrow. Scale bar: 20 mm. pnf [13:0] llf [18:1] Figure 2. Nasoral region with lifted anterior nasal flap and exposed posterior nasal flap. A, Scyliorhinus ugoi, USNM 221611, male, 432 mm TL; B, Cephaloscyllium isabella, USNM 320594, female, 390 mm TL; Poroderma pantherinum, SAIAB 34577, male, 640 mm TL. anf, anterior nasal flap; ful, flap on the upper lip margin; IIf, lower labial furrow; mrd, mesonarial crest; nb, nasal barbel; pnf, posterior nasal flap; pog, postoral groove; ulf, upper labial furrow. Scale bar: 20 mm. A anf Figure 3. Nasoral region with lifted anterior nasal flap and exposed posterior nasal flap. A, Scyliorhinus canicula, USNM 221470, female, 438 mm TL; B, Ate/omycterus fasciatus, CSIRO H1298-7, male, 370 mm TL; C, Haploblepharus edwardsii, AMNH 40988, male, 480 mm TL. anf, anterior nasal flap; ful, flap on the upper lip margin; IIf, lower labial furrow; ng, nasoral groove; pnf, posterior nasal flap; pog, postoral groove; ulf, upper labial furrow; umf, upper mesonarial flap. Scale bar: 20 mm. zse.pensoft.net Zoosyst. Evol. 96 (2) 2020, 345-395 in Apristurus, Atelomycterus, Aulohalaelurus and Hap- loblepharus (state 1; Fig. 3B, C). 12 Degree of development of posterior nasal flap: (0) corresponding to one-half of the area of the anterior nasal flap; (1) reduced and only bordering the posteri- or tip of the excurrent aperture. (CI = 100; RI = 100). Scyliorhinines present a well-developed posterior nasal flap, corresponding to one-half of the area of anterior nasal flap, as do the genera Asymbolus and Halaelurus (state 0; Fig. 2). The posterior flap is rudimentary in other taxa, corresponding to one fourth or less of the area of the anterior nasal flap (state 1). 13 Position of the posterior nasal flap: (0) situated on the posterior border of the excurrent aperture; (1) laterally situated to the excurrent aperture. (CI = 100; RI = 100). In relation to the position of the posterior nasal flap, Scyliorhinus canicula and S. duhamelii present a unique condition, 1.e. the posterior nasal flap is anteroposteriorly elongated and laterally situated at the excurrent aperture (state 1; Fig. 3A). In other species of Scyliorhinus and carcharhiniforms, this flap is situated along the posteri- or margin of the excurrent aperture (state 0; Fig. 2). In orectolobiforms, a posterior nasal flap is also laterally situated at the excurrent aperture (Goto 2001). The sim- ilarity to the position of the posterior flap in S. canicula, S. duhamelii and in orectolobiforms may be related to the presence of a nasoral groove (Bell 1993). 14 Nasoral grooves: (0) absent; (1) present. (CI = 33; RI = 33). A nasoral groove, which links the excurrent aperture and the mouth, is observed only in Scyliorhinus canicula and S. duhamelii, amongst scyliorhinines (state 1; Fig. 3A). Nasoral grooves are also observed in the scyliorh- inids, Atelomycterus and Haploblepharus (Fig. 3B, C). White (1937) pointed out that the occurrence of nasoral grooves, as well as the distance of nasal flaps from the mouth, may be directly related to the environment in her Catuloidea (= Scyliorhinidae). However, species that present the same habitats and same geographic range as S. canicula and S. duhamelii (e.g. S. stellaris) lack these structures. Shirai (1996) and de Carvalho (1996) listed and coded the occurrence of nasoral grooves in their analyses, but did not comment on the differences found in the nasoral region of scyliorhinids and orec- tolobiforms. In carcharhiniforms, such flaps are shallow and wide, distinguishing them from the deep nasoral grooves of orectolobiforms that are flanked by a com- plex arrangement of flaps and projections. According to Bell (1993), the sporadic occurrence of nasoral grooves and associated features suggest that such structures have evolved independently at least three times in scyliorhi- nids and once in triakids. 351 15 Upper labial furrow: (0) present; (1) absent. (CI = 33; RI = 83). An upper labial furrow is absent in Scyliorhinus, Ceph- aloscyllium, Holohalaelurus and Poroderma africanum (state 1; Figs 1-3). This furrow is present in Poroderma pantherinum, Proscyllium habereri and other scyliorhi- nids (state 0; Figs 1-3). 16 Lower labial furrow: (0) present; (1) absent. (CI = 50; RI = 75). A lower labial furrow is present in Scyliorhinus, Poro- derma, Proscyllium and other scyliorhinids (state 0; Figs 1-3) and absent in Cephaloscyllium and Holohalaelurus (state 1; Figs 1-3). In Cephaloscyllium sufflans and C. variegatum, we observed small notches close to the com- missure of the mouth, which do not, however, correspond to labial furrows. 17 Projected flap on the upper lip margin: (0) absent; (1) present. (CI = 50; RI = 94). In Scyliorhinus species and Poroderma africanum, there is a projected flap on the upper lip margin that laterally covers the lower labial furrow and its external margin does not extend anteriorly (state 1; Figs 1A, 2A, 3A); this is a unique condition in the family Scyliorhinidae. The pres- ence of this flap was considered a diagnostic character of Scyliorhinus by Springer (1979), Compagno (1988a) and Soares and de Carvalho (2019) and may be related to the position of the upper labial cartilage, which is internal to the preorbitalis muscle anteriorly and ventral and external to the m. adductor mandibulae posteriorly. 18 Configuration of labial furrows: (0) continuous and fused laterally; (1) discontinuous and upper furrow ventral to the lower one. (CI = 33; RI = 0). In taxa, where both labial furrows are found, there is a difference concerning their configuration. In Poroderma pantherinum, these furrows are narrow and discontinu- ous and the posterior tip of the upper furrow is ventrally situated at the lower one (state 1; Fig. 2C). The same con- dition is also found in Parmaturus and Schroederichthys. In Proscyllium and all other scyliorhinids, labial furrows are continuous and laterally fused (state 0; Fig. 3B, C). The configuration of labial furrows may be related to the presence or absence of a fusion between labial cartilages; taxa, in which the furrows are continuous, also presented fused labial cartilages, whereas the furrows are discontin- uous in taxa with separated labial cartilages. 19 Number of upper labial cartilages: (0) two; (1) one. (CI = 50; RI = 92). Compagno (1988a) pointed out that the reduction or loss of labial furrows and flaps may be related to loss or reduc- zse.pensoft.net S52 tion of labial cartilages. Shirai (1992a) divided labial car- tilages into four states, according to their number: (1) three cartilages present (two upper and one lower); (2) two car- tilages (one upper and one lower); (3) only one upper car- tilage present (lower absent); (4) both cartilages absent. In this study, only the number of upper labial cartilages was considered, ranging between one and two. In Scyliorhi- nus, aS well as in Cephaloscyllium and Schroederichthys, only one upper cartilage was observed (state 1; Fig. 4A). In Proscyllium and the other catsharks, two upper carti- lages were found (state 0; Fig. 4B), with the exception of Holohalaelurus, which presents no upper labial cartilage, but only a lower one (an autapomorphy for this genus). 20 Postoral groove: (0) absent; (1) present. (CI = 100; RI = 100). In species of Cephaloscyllium, a postoral groove is found, consisting of a slit extending from the oral commissure by an extension of up to one-fifth of the width of the mouth (state 1; Fig. 2B); the extension of this groove is variable amongst species of this genus. In other taxa examined, this groove is absent (state 0; Figs 1-3). In Holohalaelurus, in which labial furrows are absent as in Cephaloscyllium, no notch or postoral groove is observed (Fig. 1C). Fins 21 Pelvic apron: (0) absent; (1) present. (CI = 33; RI = 87). The fusion of the pelvic inner margins, known as the pel- vic apron and defined by Compagno (1988a), 1s observed in males of species of Scyliorhinus, covering their clasp- ers (state 0; Fig. 5). In Cephaloscyllium and Poroderma, the pelvic inner margins are fused only at their origin; however, this condition is not considered a true pelvic apron because it has also been observed in females where no claspers are found. This fusion 1s absent in most other scyliorhinids, except in Asymbolus and Holohalaelurus. 22 Extension of pelvic apron: (0) fusion extending up to one-half the length of pelvic inner margins; (1) fu- sion extending up to two thirds of the length of pelvic inner margins; (2) pelvic inner margins almost entire- ly fused. (ordered; CI = 100; RI = 100). Amongst the taxa presenting the pelvic apron, there is some variation in its extension. In Asymbolus and Holo- halaelurus, the pelvic apron may be present only in the proximal portion of the pelvic inner margins, correspond- ing to less than one-half of the length of the inner margins. Species of Scyliorhinus present a more developed pelvic apron, ranging from up to two thirds of the length of the pelvic inner margins (most of species; Fig. 5A) to almost their entire length (S. canicula, S. capensis, S. duhamelii, S. torazame and S. torrei). In these species, claspers of ju- veniles are totally concealed ventrally by the pelvic apron and visible only when it is lifted (state 2; Fig. 5B). The zse.pensoft.net Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus more developed pelvic apron, extending up to two thirds of the length of pelvic inner margins, was considered by Soares and de Carvalho (2019) as a synapomorphy for Scyliorhinus species. 23 Origin of the first dorsal fin: (0) closer to the vertical line that passes through the insertion of pelvic fins; (1) closer to the vertical line that passes through the origin of pelvic fins. (CI = 33; RI = 0). The posteriormost origin of the first dorsal fin is the main character used to diagnose the family Scyliorhini- dae (Springer 1979; Compagno 1988a; Compagno et al. 2005; Ebert et al. 2013). Most of its genera present the or- igin of the first dorsal fin well posterior to the vertical line that passes through the origin of the pelvic fins and closer to their insertion, ranging from opposite the insertion to half-length of pelvic inner margins. In general, dorsal fins are more posteriorly situated in males than in females. The exception is observed in Cephalurus and Parmaturus, in which the origin of the first dorsal fin is slightly anterior or opposite to the origin of pelvic fins. In Proscyllium and other carcharhiniforms, the first dorsal fin is completely anterior to pelvic fins and its origin may be opposite to the posterior tip or to the half-length of the pectoral inner margins. According to White (1937) and Nakaya (1975), the relative position of the dorsal fins is more anterior in more derived carcharhiniforms and is a character of great phylogenetic relevance. By this criterion, the Scyliorh- inidae would be considered the most basal clade within carcharhiniforms. However, Compagno (1988a) pointed out that this character should be cautiously interpreted and better investigated. Regarding the fossil record and the widespread occurrence amongst diverse groups, the anterior position of the first dorsal fin would be primitive in carcharhiniforms, whereas posterior dorsal fins might be a secondary condition, correlated with a more derived benthic habit (Compagno 1988a). 24 Origin of second dorsal fin: (0) posterior to the vertical line that passes through half-length of anal fin base; (1) anterior to the vertical line that passes through half-length of anal fin base. (CI = 25; RI = 50). Two conditions were observed concerning the origin of the second dorsal fin: posterior (most scyliorhinids; state 0, Fig. 6) or anterior to the vertical line that passes through half-length of anal fin base (Cephaloscyllium, Cephal- urus, Parmaturus and Proscyllium, state 1, Fig. 6). Dermal denticles 25 Cusplets of dermal denticles on dorsolateral body surface: (0) present; (1) absent. (CI = 33; RI = 0). The crown of the dermal denticles on the dorsolateral sur- face of the body varies from ‘teardrop’ to ‘trident’ shape due to the presence or absence of cusplets lateral to the Zoosyst. Evol. 96 (2) 2020, 345-395 A 353 Icl~ Figure 4. Labial cartilages. A, detail of labial cartilages in Scyliorhinus ugoi, USNM 221611, male, 432 mm TL; B, detail of labi- al cartilages in Asymbolus rubiginosus, AMS 1.30393-004, male, 527 mm TL. uel, upper labial cartilage; Iel, lower labial cartilage. B Figure 5. Pelvic apron. A, detail of pelvic apron in Scyliorhinus comoroensis, MNHN 1984—0701, male, 457.2 mm TL; B, detail of pelvic apron in S. capensis, SAIAB 27577, male, 863 mm TL. imp, inner margins of pelvic fins; pa, pelvic apron. Modified from Soares and de Carvalho (2019). principal cusp of the crown. Cusplets are present in most scyliorhinids (state 0; Fig. 7), but absent in Cephalurus, Parmaturus and Schroederichthys (state 1; Fig. 7). Ac- cording to Reif (1985), a greater number of cusplets and ridges would contribute to drag reduction during swim- ming, improving hydrodynamics. 26 Extension of ectodermal pits in dorsal surface of the crown denticles: (0) extending through more than half the length of the crown; (1) restricted to anterior portion of the crown (CI = 25; RI = 50). Ectodermal pits were observed and illustrated by Reif (1982) in dermal denticles of carcharhiniforms and named by Mufioz-Chapuli (1985). In scyliorhinines, Afe- lomycterus, Cephalurus, Haploblepharus and Schroed- erichthys, only the proximal portion of the crown den- ticles is covered by these pits (state 0; Fig. 7). In other scyliorhinids, ectodermal pits extend through more than half or almost the entire length of the crown, mainly in denticles on anteriormost regions of the body (Apristurus, Asymbolus, Figaro, Galeus, Halaelurus, Holohalaelurus and Parmaturus) (state 1; Fig. 7). 27 Median ridges on dermal denticles: (0) two ridges; (1) one ridge. (CI = 33; RI = 50). White (1937) categorised the dermal denticles of elasmo- branchs according to the features observed in the crown. According to her, scyliorhinids have dermal denticles with flat crowns presenting incomplete median ridges not extending to the distal tip of the crown (e.g. Scyliorhinus retifer and Schroederichthys bivius) or complete medi- an ridges, extending to the distal tip of the crown (e.g. zse.pensoft.net 354 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus Figure 6. Caudal region. A, detail of dorsal fins in Scyliorhinus haeckelii, UERJ 2202, male, 444 mm TL; B, detail of dorsal fins in Parmaturus xaniurus, CAS 232152, female, 450 mm TL. Scale bar: 25 mm. Atelomycterus spp., Halaelurus burgeri and Parmaturus spp.). We observed that the extension and degree of de- velopment of median ridges present variation according to taxa and body region examined. However, the number of ridges varies 1n a consistent manner, making it possible to separate the dermal denticles into two categories: one or two median ridges present on the dorsal surface of the crown, extending from its base to the distal tip or close to it. In scyliorhinines, Apristurus, Cephalurus, Galeus and Holohalaelurus, only one median ridge, more prominent than lateral ridges, is observed (state 1; Fig. 7). In other scyliorhinids, two prominent median ridges, forming a gutter in between them, are present (state 0; Fig. 7). 28 Caudal crest of enlarged dermal denticles: (0) absent; (1) present. (CI = 100; RI = 50). The presence of a caudal crest of dermal denticles distinct from the denticles on dorsolateral surfaces and situated on the upper lobe of caudal fin, 1s found in Figaro, Gale- us, Parmaturus and some species of Apristurus (state 1; Fig. 8), varying widely amongst species of these genera. This crest is absent in other taxa examined. In Figaro, a crest of enlarged dermal denticles on the lower lobe of the caudal fin was also observed. The occurrence of a cau- dal crest of dermal denticles is widely used 1n taxonomic studies of the family Scyliorhinidae (Linnaeus 1758; Re- gan 1908; Garman 1913; Bigelow and Schroeder 1948; Springer 1966, 1979), but has never been analysed in a cladistic study until now. zse.pensoft.net Musculature 29 Muscle depressor palpebrae nictitantis: (0) present; (1) absent. (CI = 100; RI = 100). The postorbital musculature is composed of three mus- cles: m. depressor palpebrae nictitantis, m. levator pal- pebrae nictitantis and m. retractor palpebrae nictitantis. These muscles are responsible for elevation and depres- sion of the nictitating lower eyelid, which is a diagnostic character for carcharhiniforms. These muscles were found in most of the taxa examined, except in scyliorhinines, in which only the muscles /evator palpebrae nictitantis and retractor palpebrae nictitantis are present (state 1; Fig. 9A). The absence of the depressor palpebrae nictitantis in the subfamily Scyliorhininae was already reported by Compagno (1988a) and is one of the characters used by that author to diagnose it. Specimens of Aulohalaelurus labiosus, Scyliorhinus comoroensis and S. garmani were not available for dissection (these taxa are scored with a question mark in the matrix). 30 Insertion of the muscle coracomandibularis: (0) on the articular region of the antimeres of Meckel’s car- tilage; (1) near the mid-length of the lower jaws, on their anteromedial borders. (CI = 50; RI = 0). In most taxa examined, the m. coracomandibula- ris inserts on the articular region of the antimeres of Meckel’s cartilage (state 0; Fig. 1OA). In Haploblepha- Zoosyst. Evol. 96 (2) 2020, 345-395 355 Figure 7. Dermal denticles above the origin of the first dorsal fin. A, Apristurus longicephalus, HUMZ 170382, male, 475 mm TL; B, Asymbolus rubiginosus, AMS 1.30393-004, male, 527 mm TL; C, Atelomycterus fasciatus, CSIRO H1298-7, male, 370 mm TL; D, Cephaloscyllium sufflans, SALAB 6242, male, 800 mm TL; E, Cephalurus cephalus, USNM 221527, female, 285 mm TL; F, Galeus antillensis, UF 77853, female, 370 mm TL; G, Halaelurus natalensis, SALAB 26951, male, 400 mm TL; H, Haploblepharus edwardsii, AMNH 40988, male, 480 mm TL; I, Holohalaelurus regani, SALIAB 25717, male, 610 mm TL; J, Parmaturus xaniurus, CAS 232152, female, 450 mm TL; K, Poroderma africanum, SAIAB 25343, male, 920 mm TL; L, Schro- ederichthys saurisqualus, UERJ uncatalogued, female, 564 mm TL. Scale bar: 250 um. rus edwardsii and Poroderma africanum, this muscle is divided into two portions anteriorly, each of them inserting on the anteromedial borders of the antimeres of Meckel’s cartilage and not occupying the symphysial region of the jaws (state 1; Fig. 10B). In these taxa, a basimandibular cartilage is observed at the symphysial region of jaws, connecting the antimeres of Meckel’s cartilage. 31 Insertion of the muscle coracohyoideus: (0) on the ventral surface of the basihyal cartilage; (1) on con- nective tissue adjacent to the basihyal cartilage. (CI = 50; RI = 0). In most taxa examined, the muscle coracohyoideus in- serts on the ventral surface of the basihyal (state 0; Fig. 11). In Apristurus and Holohalaelurus this muscle inserts on lateral (Apristurus longicephalus) or anterior (Holo- halaelurus regani) connective tissue projections of the basihyal (state 1; Fig. 11D). 32 Configuration of muscles bundles of the m. coraco- hyoideus: (0) juxtaposed muscle bundles; (1) separat- ed muscle bundles. (CI = 33; RI = 60). The presence of m. coracohyoideus composed of two distinct muscle bundles originating in the fascia of the zse.pensoft.net 356 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scy/liorhinus Figure 8. Detail of the caudal crest in Galeus antillensis, UF 77853, female, 370 mm TL. A, general view, B, closed view. Scale bar:10 mm (A); 200 um (B). A Figure 9. Detail of the postorbital musculature. A, Schroederichthys maculatus, UF 65846, male, 313 mm TL; B, Poroderma africanum, AMNH 43134, male, 505 mm TL. mdp, muscle depressor palpebrae nictitantis, mlp, m. levator palpebrae nictitantis; mrp, m. retractor palpebrae nictitantis; msp, m. espiracularis, ob, orbit, sp, spiracle. m. coracoarcualis was observed in Proscyllium and all scyliorhinids. These bundles can be juxtaposed (most taxa examined) or separated by a distance of at least one- half the width of each bundle (Cephaloscyllium, Cephal- urus and Halaelurus, state 1, Fig. 11A). 33 Origin of the muscles coracobranchialis Il, Il and IV: (0) on the coracoid bar; (1) on the pericardial membrane. (CI = 50; RI= 75). zse.pensoft.net In Proscyllium and some scyliorhinids, the origin of the muscles coracobranchialis I, 1 and IV is on the cora- coid bar (Apristurus, Asymbolus, Atelomycterus, Cepha- loscyllium, Figaro, Galeus, Parmaturus, Poroderma and Scyliorhinus, state 0, Fig. 12A). In other scyliorhinids, these muscles originate from the pericardial membrane, a layer of connective tissue anterior to the coracoid bar and ventral to the heart (Cephalurus, Halaelurus, Haploblepharus, Holo- halaelurus and Schroederichthys, state 1, Fig. 12B). Zoosyst. Evol. 96 (2) 2020, 345-395 The insertion of the m. coracobranchialis presents the following pattern in the taxa examined: cora- cobranchialis II, on the medial border of the cera- tobranchial II cartilage and anterolateral border hy- pobranchial II; coracobranchialis Ill, on the medial border of the ceratobranchial III cartilage and antero- lateral border of hypobranchial III; coracobranchial- is IV, on the medial border of the ceratobranchial IV cartilage and anterolateral border of hypobranchial IV. Muscle coracobranchialis V presents the same pattern in the taxa examined, originating from the anterolateral borders of the coracoid bar and inserting on the medial border of ceratobranchial V and lateral border of the ba- sibranchial copula. Neurocranium 34 Rostral cartilages: (0) fused; (1) united only by con- nective tissue. (CI = 100; RI = 100). In scyliorhinines, Atelomycterus and Aulohalaelurus, the rostrum is formed by three rostral cartilages ante- riorly united only by connective tissue (state 1; Fig. 13A—D), whereas in other taxa examined, these carti- lages are fused anteriorly, sometimes forming or not a rostral node (state 0; Fig. 13E—G). Compagno (1988a) proposed that the absence of fusion between rostral cartilages could be an independently derived and the secondary condition for scyliorhinids and proscylliids based on their proximity to taxa in which the fused con- dition is present. 35 Relation between lateral rostral cartilages and anterior fontanelle: (0) rostral cartilages distant from anterior 357 fontanelle; (1) rostral cartilages confluent with lateral borders of anterior fontanelle. (CI = 100; RI = 100). The distance between lateral rostral cartilages may vary, positioned medially or laterally to the lateral borders of the anterior fontanelle. In some cases, the lateral rostral cartilages are confluent with the lateral borders of the an- terior fontanelle, connected to it through ridges that ex- tend from the base of the rostral cartilages to the border of the fontanelle; this condition was observed in Apristurus, Figaro, Galeus and Parmaturus (state 1; Fig. 13F, G). In other taxa examined, the lateral rostral cartilages are dis- tant from the anterior fontanelle and do not present ridges in between both structures (state 0; Fig. 13). 36 Relationship between median rostral cartilage and anterior fontanelle: (0) median rostral cartilage and anterior fontanelle separated by internasal space; (1) median rostral cartilage confluent with anterior fon- tanelle. (CI = 50; RI = 50). The distance between the median rostral cartilage and ante- rior fontanelle varies amongst taxa examined. In scyliorhin- ines and other taxa examined, the median rostral cartilage and the anterior fontanelle are separated by the internasal space, distant by at least two thirds of the length of the me- dian rostral cartilage (state 0; Fig. 13A). In Cephalurus, Haploblepharus and Holohalaelurus, the anterior border of the anterior fontanelle is adjacent to the base of the me- dial rostral cartilage, without an internasal space separating them (state 1; Fig. 14). Compagno (1988a) listed the mea- surement “distance from the ventral border of the anterior fontanelle to the base of the median rostral cartilage’ as a way of measuring the space between these structures. Figure 10. Detail of the insertion region of the muscle coracomandibularis. A, Cephaloscyllium umbratile, USP uncatalogued, male, 409 mm TL; B, Haploblepharus edwardsii, AMNH 40988, male, 480 mm TL. cbm, basimandibular cartilage; cor, coracoid bar; mea, muscle coracoarcualis, mech, m. coracohyoideus, mck, Meckel's cartilage; mem, m. coracomandibularis. zse.pensoft.net 358 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus Minin KenN\ cor Figure 11. Hypobranchial musculature. A, Cephaloscyllium umbratile, uncatalogued, male, 409 mm TL; B, Galeus antillensis, UF 77853, female, 370 mm TL; C, Holohalaelurus regani, SATAB 25717, male, 610 mm TL; D, Apristurus longicephalus, HUMZ 170382, male, 475 mm TL; bh, basihyal; bhf, adjacent flap of the basihyal; ch, ceratohyal; cor, coracoid bar; mca, muscle coracoarcualis; mch, m. coracohyoideus; mem, m. coracomandibularis, ocm, origin of the m. coracomandibularis. 37 Orientation of nasal capsules: (0) nasal capsules per- the nasal capsules are orientated perpendicularly and lat- pendicular to the anteroposterior axis of the neurocra- erally expanded (state 0; Fig. 15). nium; (1) nasal capsules oblique. (CI = 50; RI = 0). 38 Relative position between nasal apertures: (0) in- In Apristurus and Galeus, the nasal capsules are oblique- current aperture anterior to excurrent one; (1) na- ly orientated to the anteroposterior axis of the neurocrani- sal apertures at the same level. (CI = 50; RI = 90). um (state 1; Fig. 15), whereas in the other scyliorhinids, zse.pensoft.net Zoosyst. Evol. 96 (2) 2020, 345-395 359 A B eel hb Ill hb Il hb Ill hb Il [33:0] St { xe i v4 N cb I~ Sore cbb ZB aM cb Il ~ mcb Ill mcb Iil ) Lif \ cb Ill cbb mcb IV cor ~cb IV mcb IV ~ebV— [33: 1] mpc Figure 12. Detail of muscle coracobranchialis. A, Cephaloscyllium umbratile, USP uncatalogued, male, 409 mm TL; B, Schro- ederichthys saurisqualus, UERJ uncatalogued, female, 564 mm TL. bbq, basibranchial; cb I-V, ceratobranchials I-V; cbb, basibranchial copula; cor, coracoid bar; hb H-II, hypobranchials M—fI; meb H-IV, muscles coracobranchialis |-IV; mpc, pericardial membrane. Nasal apertures may be positioned at the same level (Cephaloscyllium, Poroderma, Schroederichthys and Scyliorhinus, state 1, Fig. 15A—C) or at distinct levels with the incurrent aperture anterior to the excurrent one (most taxa examined; Fig. 15). 39 Fusion of the external nasal cartilage to the dorsal position of the nasal capsule: (0) present: (1) absent. (CI = 50; RI = 89). The external nasal cartilage, situated anteriorly to the nasal apertures and ventrally to the nasal capsules, may or may not be fused to the anterodorsal portion of the nasal capsules (Goto 2001). The fusion is observed in Apristurus, Asymbolus, Cephalurus, Figaro, Galeus, Ha- laelurus, Haploblepharus, Holohalaelurus, Parmaturus and Proscyllium (state 0; Fig. 15E—G). In scyliorhinines, Atelomycterus, Aulohalaelurus and Schroederichthys, the external nasal cartilage is not fused to the nasal capsules and a narrow strip of connective tissue separates the ex- ternal cartilage from the dorsal portion of the nasal cap- sules (state 1; Fig. 1SA—D). 40 Degree of development of the subnasal plate: (0) re- stricted to the medial portion of the nasal capsules and ventral to the internasal septum; (1) laterally ex- panded and united to the lateral border of the nasal capsule. (CI = 33; RI = 0). The subnasal plate is the ventral floor of the nasal cap- sules, generally associated with a cavity posteromedial to the incurrent aperture and covered by a layer of connec- tive tissue (nasal fontanelle of Compagno 1988a, 1999). In most taxa examined, the subnasal plate is restricted to the medial portion of the nasal capsules and ventral to the internasal septum and the nasal fontanelle occupies the entire region posterior to the excurrent aperture (state 0; Fig. 15). In Apristurus, Galeus and Proscyllium, the sub- nasal plate is laterally expanded, occupying almost the entire region posterior to the excurrent aperture and the nasal fontanelle is reduced to a narrow strip at the pos- terior border of the excurrent aperture, divided into two portions (state 1; Fig. 15G). Compagno (1988a) suggest- ed a tendency concerning the enlargement of the subnasal plate in derived taxa and consequent substitution of the nasal fontanelle by cartilage. 41 Epiphyseal notch: (0) absent; (1) present. (CI = 25; RI = 79). The anterior fontanelle, the anterodorsal aperture of the neurocranium covered by a layer of connective tissue, presents different shapes amongst species and also varies between sexes (Soares et al. 2015, 2016). This fontanelle may present a notch or an indentation on its posterior bor- der, the epiphyseal notch to the pineal body, as observed in Atelomycterus, Halaelurus, Holohalaelurus, Schroed- erichthys and Scyliorhinus (state 1; Figs 13A, C, D, 14) or a straight and continuous border, as in Cephaloscyllium, Poroderma and in the other taxa examined (state 0; Figs 13B, E and G). In Holohalaelurus, this notch 1s well de- veloped, corresponding to two thirds of the length of the anterior fontanelle (Fig. 14). 42 Supraorbital crest: (0) present; (1) absent. (CI = 50; RI = 87. The occurrence of a supraorbital crest on the neurocranium is widely used for identification and separation of shark genera and families. The presence of this crest is con- sidered primitive for elasmobranchs and its absence sec- ondary in some sharks and rays (Compagno 1988a). This zse.pensoft.net 360 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus Figure 13. Neurocranium; dorsal view. A, Scyliorhinus ugoi, USNM 221611, male, 432 mm TL; B, Cephaloscyllium varie- gatum, AMS 1.43762-001, female, 670 mm TL; C, Schroederichthys saurisqualus, UERJ uncatalogued, female, 564 mm TL; D, Atelomycterus fasciatus, CSIRO H1298-7, male, 370 mm TL; E, Asymbolus rubiginosus, AMS 1.30393-004, male, 527 mm TL; F, Figaro boardmani, CSIRO H989-5, female, 465 mm TL; G, Apristurus longicephalus, HUMZ 170382, male, 475 mm TL. af, anterior fontanelle; asc, anterior semicircular canal; eph, epiphyseal notch; foe, external foramen of preorbital canal; ins, internasal septum; iof, infraorbital canal of the lateral line; Ire, lateral rostral cartilage; mre, medial rostral cartilage; ne, nasal capsule; pep, preorbital process; prf, parietal fossa; psec, posterior semicircular canal; pt, pterotic process; ptp, postorbital process; sc, supraorbital crest. structure 1s situated dorsally to the orbits and continuous to pre- and postorbital processes in scyliorhinines, Ate/omyc- terus, Aulohalaelurus, Proscyllium and Schroederichthys (state 1; Fig. 13 A—D). In scyliorhinids of the subfamily Pentanchinae (sensu Compagno, 1988a), the supraorbital crest is absent (state 0; Fig. 13E-G). Iglésias et al. (2005) zse.pensoft.net used the occurrence of this crest to distinguish the families Scyliorhinidae and Pentanchidae (= subfamily Pentanchi- nae of Compagno 1988a), although these authors did not provide further information about the condition found in other families of carcharhiniforms. According to Compag- no (1988a), the loss of the supraorbital crest in Hemigalei- Zoosyst. Evol. 96 (2) 2020, 345-395 pore ms wa ages isagsaghtes owe f ~pep Figure 14. Detail of rostral region and anterior fontanelle of Holohalaelurus regani, SALAB 25717, male, 610 mm TL. af, anterior fontanelle; eph, epiphyseal notch; Ire, lateral ros- tral cartilage; mre, medial rostral cartilage; ne, nasal capsule; pep, preorbital process. dae, Carcharhinidae and Sphyrnidae may be related to the anterior expansion of the muscle /evator palatoquadrati dorsal to the orbital wall and neurocranial roof. However, in scyliorhinids without a crest, the muscle /evator pala- toquadrati originates in the ventral surface of the postor- bital process and is situated entirely posterior to the orbit. 43 Distance between internal carotid foramina: (0) greater than the distance between internal carotid and stapedial foramina; (1) smaller than the distance between internal carotid and stapedial foramina; (2) equal to the distance between internal carotid and sta- pedial foramina. (ordered; CI = 20; RI = 27). Four foramina are present on the posterior portion of the basal plate: two for the medial internal carotid arter- ies and two for the lateral stapedial arteries. Compagno (1988a) reported the median position of the foramina of the internal carotid artery on the basal plate in scyliorh- inids, proscylliids and Pseudotriakis, but did not pro- pose any distinction between the patterns observed. The distance between these foramina varies widely amongst scyhorhinids. In Cephaloscyllium, Holohalaelurus, Proscyllium, Schroederichthys and Scyliorhinus, fo- ramina to the internal carotid artery are very close to each other and separated by a shorter distance than the distance between the internal carotid and stapedial fo- ramina, which are fused in some cases (state 1; Fig. 15A-C). In Atelomycterus, Aulohalaelurus, Cephal- urus, Halaelurus and Parmaturus, the internal carotid foramina are separated by a distance similar to that be- tween internal carotid and stapedial foramina (state 2; Fig. 15D). In Apristurus, Asymbolus, Figaro, Galeus, Haploblepharus and Poroderma, the distance between the internal carotid foramina is greater than the distance between the internal carotid and stapedial foramina (state 0; Fig. 1SE-G). 361 44 Relative size of postorbital groove: (0) groove corre- sponds to more than one-half the height of the hyo- mandibular facet; (1) groove corresponds to less than one-half the height of the hyomandibular facet. (CI = 33; RI=0). The postorbital groove is situated posteriorly to the orbits and ventral to the postorbital processes, limited dorsally by the opisthotic process and ventrally by the hyoman- dibular facet; the lateral vein of the head passes along this groove (Compagno 1988a). In most scyliorhinids, the postorbital groove corresponds to more than one-half of the height of the hyomandibular facet, resulting in a prominent and laterally visible structure (state 0; Fig. 16A, B). In Apristurus, Aulohalaelurus and Cephalurus, this groove is very narrow and shallow, corresponding to one-third or less of the height of the hyomandibular facet (state 1; Fig. 16C). 45 Fenestra for the infraorbital canal of the lateral line: (0) present; (1) absent. (CI = 50; RI = 0). The pre- and postorbital processes are laterally expand- ed from the neurocranial roof, as wide as or wider than the nasal capsules. In most scyliorhinids, the distal tip of the postorbital process has a large fenestra through which passes the infraorbital canal of the lateral line (Compagno 1988a; state 0, Fig. 13). In Apristurus and Schroederich- thys, this fenestra is absent. In the latter, the infraorbital canal passes through a bifurcation situated at the distal tip of the postorbital process (Fig. 13C). In Apristurus, the postorbital process is narrow and rod-like, not presenting any bifurcation or fenestra and the infraorbital canal of the lateral line passes posteriorly to it (Fig. 13G). Jaws 46 Labial ridge of the quadrate process: (0) present; (1) absent. (CI = 17; RI= 17). Compagno (1999) pointed out that, primitively, the palatoquadrate of sharks presents elevated quadrate processes with prominent ridges on the labial surface; this configuration was observed in Notorynchus cepedi- anus (Daniel 1934: fig. 48). The presence of a prominent ridge on the labial surface and about one-half the length of the quadrate process is found in some scyliorhinids (Asymbolus, Atelomycterus, Cephalurus, Galeus, Ha- laelurus, Haploblepharus and Parmaturus, state 0, Fig. 17B) and may be related to the region of insertion of the m. preorbitalis and the division between dorsal and ven- tral portions of the m. adductor mandibulae. This ridge is absent in scyliorhinines, Apristurus, Figaro, Holo- halaelurus, Proscyllium and Schroederichthys (state 1; Fig. 17A). 47 Position of the orbital processes of the palatoquad- rate: (0) at the anterior one-fourth of each antimere; zse.pensoft.net 362 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus Figure 15. Neurocranium; ventral view. A, Scyliorhinus ugoi, USNM 221611, male, 432 mm TL; B, Cephaloscyllium varie- gatum, AMS 1.43762-001, female, 670 mm TL; C, Schroederichthys saurisqualus, UERJ uncatalogued, female, 564 mm TL; D, Atelomycterus fasciatus, CSIRO H1298-7, male, 370 mm TL; E, Asymbolus rubiginosus, AMS 1.30393-004, male, 527 mm TL; F, Figaro boardmani, CSIRO H989-5, female, 465 mm TL; G, Apristurus longicephalus, HUMZ 170382, male, 475 mm TL. bp, basal plate; enc, external nasal cartilage; exc, excurrent aperture; hf, hyomandibular facet; icf, internal carotid foramen; ine, incurrent aperture; Irc, lateral rostral cartilage; mre, medial rostral cartilage; nf, nasal fontanelle; sbp, subnasal plate; sf, stapedial foramen; ss, suborbital shelf. (1) closer to the half-length of each antimere. (CI = 100; RI = 100). The palatoquadrate articulates to the neurocranium by ethmopalatine ligaments, which are inserted on the postorbital processes of the palatoquadrates and origi- zse.pensoft.net nate from the orbital notches; these notches are situated between the posteroventral region of the nasal capsules and the preorbital wall. Orbital processes are situated in variable positions in the dorsal border of each antim- ere of the palatoquadrate, delimitating the extension of palatine and quadrate processes. In most scyliorhinids, Zoosyst. Evol. 96 (2) 2020, 345-395 363 A | [44:1] Figure 16. Neurocranium; lateral view. A, Scyliorhinus ugoi, USNM 221611, male, 432 mm TL; B, Figaro boardmani, CSIRO H989-5, female, 465 mm TL; C, Apristurus longicephalus, HUMZ 170382, male, 475 mm TL. enc, external nasal cartilage; hf, hyomandibular facet; Irc, lateral rostral cartilage; mre, medial rostral cartilage; ne, nasal capsule; pog, postorbital groove. He a Tag pL B - K Pn 46:0 ed) Figure 17. Detail of jaws; lighter portion of Meckel’s cartilage is less calcified portion. A, Scyliorhinus ugoi, USNM 221611, male, 432 mm TL; B, Apristurus longicephalus, HUMZ 170382, male, 475 mm TL. Irq, labial ridge of the quadrate process; mck, Meckel’s cartilage; opq, orbital process of the palatoquadrate; pq, palatoquadrate. orbital processes are situated at the anterior one-fourth scribed by Compagno (1988a), more posterior orbital (state 0; Fig. 17A). In Apristurus, Cephalurus, Gale- processes are situated at a greater distance from the or- us, Haploblepharus, Holohalaelurus and Parmaturus, bital notches and connected to them through elongated they are more posteriorly situated, close to one-half the ethmopalatine ligaments (except in Haploblepharus), length of the palatoquadrate (state 1; Fig. 17B). As de- this arrangement is probably an adaptation to increase zse.pensoft.net 364 jaw protusibility. In Haploblepharus, a unique condi- tion is found as the ethmopalatine ligaments are short and the articulation occurs directly between orbital pro- cesses and notches. 48 Degree of calcification of the medial portion of Meckel’s cartilage: (0) similar calcification through- out; (1) medial portion less calcified than the rest of Meckel’s cartilage. (CI = 100; RI = 100). Meckel’s cartilages present distinct degrees of calcifica- tion in some scyliorhinids. In Apristurus, Cephalurus, Figaro, Galeus and Parmaturus, the medial portion of the antimeres is less calcified than the rest of the cartilage (state 1; Fig. 17B). In the other taxa examined, the degree of calcification is equal throughout Meckel’s cartilage (state 0; Fig. 17A). 49 Articular region of the quadratomandibular joint of Meckel’s cartilage: (0) posterior lingual condyle sit- uated between anterior labial condyle and facet; (1) anterior and posterior condyles forming a unit and distant from the facet; (2) posterior lingual condyle opposite to the facet. (ordered; CI = 50; RI = 60). Moss (1972) described two articular regions between the jaws in carcharhinid sharks, relating that both regions correspond to the ‘ball and socket’ type of articulation: palatoquadrate with a convex posterior region and situ- ated more laterally and Meckel’s cartilage with a more anterior and medial condyle. Motta and Wilga (1995), describing the jaw anatomy of Negaprion brevirostris, proposed the terms ‘medial quadratomandibular joint’ (QJM) and ‘lateral quadratomandibular joint’ (QJL) to refer to the articular regions between palatoquadrate and Meckel’s cartilage. We observed a greater complex- ity and wider variation in relation to the arrangement of condyles and facets of the quadratomandibular region of Meckel’s cartilage; three patterns were identified: 1) “me- dial quadratomandibular joint’ composed of two condyles (labial and lingual), forming a unit (state 1, Fig. 18A); ii) a lingual condyle situated posteriorly to the ‘lateral quadratomandibular joint’ (Asymbolus, Atelomycterus, Figaro, Galeus, Halaelurus and Proscyllium; state 0, Fig. 18B); it1) labial condyle more internally positioned and lingual condyle more posterior and opposite to the facet (Apristurus; state 2, Fig. 18C). The first pattern is found in scyliorhinines, Cephalurus, Haploblepharus, Holo- halaelurus, Parmaturus and Schroederichthys, in which the condyles are significantly separated from the facet. In relation to the articular region of the palatoquadrate, an anterior facet and a posterior condyle have the same morphology in all taxa examined. Hyoid and gill arches 50 Thyroid foramen: (0) present; (1) absent. (CI = 33; RI = 60). zse.pensoft.net Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus The basihyal cartilage, situated ventromedially to other components of the hyoid arch, is a structure that presents variable dimensions amongst taxa. This cartilage may or may not present an opening in its anterior portion, the thyroid foramen (de Beer 1937), which is the entrance to the duct of the thyroid gland. This foramen is pres- ent in scyliorhinines, Asymbolus, Atelomycterus, Figaro, Halaelurus, Holohalaelurus and Schroederichthys (state 0; Fig. 19), but absent in other scyliorhinids (state 1). In taxa without a thyroid foramen, the duct of the thyroid gland passes anteriorly to the anterior border of the basi- hyal cartilage. 51 Internal surface of the hyomandibular cartilage: (0) smooth; (1) concave. (CI = 50; RI = 0). In Apristurus and Parmaturus, we observed a prominent concavity on the internal surface of the posterior region of the hyomandibular cartilage, close to the articular region between the ceratohyal and Meckel’s cartilages (state 1; Fig. 20B). This concavity is situated in the region of the insertion of the muscles constrictor superficialis dorsalis and /evator hyomandibulae in Parmaturus, but only of the m. constrictor superficialis dorsalis in Apristurus. In other examined taxa, the internal surface is smooth and no concavity is present (state 0; Fig. 20A). 52 Anterior border of the basihyal cartilage: (0) not bi- furcated; (1) bifurcated. (CI = 33; RI = 33). The occurrence of a bifurcation on the anterior border of the basihyal cartilage, anterior to the thyroid foramen and not confluent with it, was observed in Ate/omycterus, Ha- laelurus, Holohalaelurus and Schroederichthys (state 1; Fig. 19B). This bifurcation was reported and illustrated for Schroederichthys chilensis by Leible et al. (1982). In scyliorhinines and other taxa examined, the basihyal car- tilage has a smooth and slightly convex anterior border (state 0; Fig. 19A). 53 Lateral processi rastriformis: (0) present; (1) absent. (CI = 33; RI = 60). Processi rastriformis were observed and illustrated in Squalus acanthias by Marinelli and Strenger (1959) and defined as anteriorly directed cartilaginous projections situated on the internal borders of the cerato- and epibranchial cartilages. Compagno (1988a; fig. 2.7) used the term ‘dermal papillae’ to refer to short structures without cartilaginous support observed in scyliorhinids, proscylliids and some carcharhinids, distinguishing these from the processi rastriformis found in squaliforms, hexanchiforms and Megachasma pelagios. The presence and distribution of these papillae vary widely in the taxa examined, whereas processi rastriformis sensu strictu were observed only in some taxa in which they occupy specific positions in relation to the gill arches. Processi rastriformis greater than the dermal papillae and situated Zoosyst. Evol. 96 (2) 2020, 345-395 365 A —_ _lactlic B Figure 18. Detail of the articular region of the quadratomandibular joint of Meckel’s cartilage. A, Cephaloscyllium sufflans, SAIAB 6242, male, 800 mm TL; Galeus antillensis, UF 77853, female, 370 mm TL; C, Apristurus longicephalus, HUMZ 170382, male, 475 mm TL. lac, labial condyle; lic, lingual condyle; mfa, mandibular facet. A a. B Figure 19. Detail of the basihyal cartilage. A, Scyliorhinus haeckelii, UERJ 1691, male, 522 mm TL; B, Schroederichthys sauris- qualus, UERJ uncatalogued, female, 564 mm TL. thf, thyroid foramen. only on the anterior surface of the articular region between are absent. The term ‘gill rakers’ used by Daniel (1934) cerato- and epibranchial cartilages were observed in and Compagno (1988a) is not used here, as the structures Asymbolus, Atelomycterus, Halaelurus, Poroderma and observed in elasmobranchs are formed by cartilage, Schroederichthys (state 0; Fig. 21). In Cephaloscyllium, — whereas gill rakers of bony fishes have a dermal origin and Scyliorhinus and other taxa examined, processirastriformis are, therefore, not homologous to processi rastriformis. zse.pensoft.net 366 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus [51:1] Figure 20. Detail of the hyomandibular cartilage; internal surface. A, Scyliorhinus haeckelii, UERJ 1691, male, 522 mm TL; B, Apristurus longicephalus, HUMZ 170382, male, 475 mm TL. 2358. 54 Oropharyngeal denticles: (0) small and not forming rows on internal face of gill components; (1) large and forming rows on internal face of gill compo- nents. (CI = 25; RI = 25). Nelson (1970) described macroscopic features, such as shape, distribution and abundance of oropharyngeal den- ticles in Rhizoprionodon terraenovae. Later, Ciena et al. (2016) and Rangel et al. (2017) described the ultra- structure of oropharyngeal denticles and their disposition amongst dermal papillae in Rhizoprionodon lalandii and Prionace glauca, respectively. In all of these species, the denticles are distributed on the entire ventral surface of the oropharyngeal cavity. Heemstra (1997) observed rows of oropharyngeal denticles greater than the denti- cles around the fifth ceratobranchial of Mustelus norrisi. Here, we report a different condition for oropharyngeal denticles in Apristurus longicephalus, Cephaloscyllium sufflans, C. variegatum, Halaelurus natalensis and Par- maturus xaniurus (state 1; Fig. 22). Besides the denticles of the ventral surface, rows of denticles, greater than the surrounding ones and similar in shape and size to the der- mal denticles of dorsolateral surfaces of the body, were found on the internal surface of gill components (cerato- and/or epibranchial) in these taxa. Two rows parallel to the gill arches and composed of 5 to 13 denticles were observed. In other taxa examined, these denticles were absent (state 0). 55 Shape of the gill pickax: (0) elongated and sling-like; (1) short and triangular. (CI = 50; RI = 80). The fusion between the dorsal tips of gill arches [TV and V, forming a unique plate known as the gill pickax (Shirai zse.pensoft.net 1992a), is observed in many neoselachians, with the ex- ception of some rays (Shirai 1996). Despite the presence of this structure having been listed by Shirai (1992a) as a diagnostic character for modern elasmobranchs (ex- cept Heterodontus and Trigonognathus), no mention of its morphological variations was provided. Amongst the taxa examined, we observed some differences regarding the shape of the gill pickax. In Cephaloscyllium and Po- roderma, the gill pickax is short and triangular in shape (state 1; Fig. 23B). In Proscyllium and other scyliorhi- nids, this structure is distally elongated and sling-like (state 0; Fig. 23). 56 Ventral extrabranchial cartilages: (0) four; (1) three. (CI = 100; RI = 100). Ventral extrabranchial cartilages are present amongst muscle bundles of the coracobranchialis and on the pos- terior border of the four anteriormost gill openings. In most taxa examined, four cartilages are observed, where- as in scyliorhinines only three are present (state 1). The number of ventral extrabranchial cartilages was listed by Compagno (1988a) as a diagnostic character of the sub- family Scyliorhininae. Pectoral skeleton 57 Medial projection of the coracoid bar: (0) present; (1) absent. (CI = 50; RI = 0). The presence of a medial projection on the coracoid bar was observed in many orders of elasmobranchs by Silva and de Carvalho (2015), who reported and illustrated its presence in the following galeomorph taxa: Heterodontus Zoosyst. Evol. 96 (2) 2020, 345-395 Figure 21. Detail of the lateral processi rastriformis (ras) in Schroederichthys saurisqualus, UERJ uncatalogued, female, 564 mm TL. francisci (Heterodontiformes), Ginglymostoma cirratum, Rhincodon typus, Stegostoma fasciatum (Orectolobi- formes), A/opias superciliosus, Carcharias taurus, Isurus oxyrhinchus, Mitsukurina owstoni, Pseudocarcharias ka- moharai (Lamniformes), Carcharhinus galapagensis and Mustelus canis (Carcharhiniformes). This projection was observed in Proscyllium and most scyliorhinid taxa ex- amined (state 0; Fig. 24A, B, D), except Haploblepharus and Schroederichthys. In these taxa, the coracoid bar has a Straight anterior border with no projections (state 1; Fig. 24C). 58 Degree of development of the medial projection of the coracoid bar: (0) reduced to less than twice the size of the lateral portion of the coracoid bar; (1) well developed, more than twice the size of the lateral por- tion of the coracoid bar. (CI = 50; RI = 90). In taxa in which a medial projection of the coracoid bar is present, differences concerning its shape and degree of development were observed. In Asymbolus, Apristurus, Atelomycterus, Galeus, Halaelurus, Poroderma and Pros- cyllium, the medial projection has an anterior border that is slightly convex and not very prominent (state 0; Fig. 24B). In Cephalurus, Cephaloscyllium and Scyliorhinus, the projection is well developed and corresponds to more lew, than twice the lateral portion of the coracoid bar. In these taxa, the medial projection entirely covers the heart ven- trally (state 1; Fig. 24A, D). 59 Lateral processes on pectoral girdle: (0) present; (1) absent. (CI = 25; RI = 57). Lateral processes on the coracoid bar, medial to the ar- ticular region between pectoral girdle and fins, were ob- served in Cephaloscyllium, Halaelurus, Haploblepharus, Schroederichthys and Scyliorhinus (state 0; Fig. 24A, B). The processes observed in scyliorhinids correspond to two thirds of the length or similar in size to the medial projection of the coracoid bar. In the illustrations provid- ed by Silva and de Carvalho (2015), projections similar to the lateral processes are present in A/opias supercilio- sus (p. 17; fig. 14) and Pseudocarcharias kamoharai (p. 30; fig. 27); the authors briefly mentioned the presence of these processes in the latter species. Clasper 60 Dermal denticles on the dorsal surface of clasper glans: (0) present; (1) absent. (CI = 25; RI = 40). Dermal denticles on the dorsal surface of clasper glans were observed in most taxa examined. In Apristurus lon- gicephalus, Cephalurus cephalus, Halaelurus natalensis, Haploblepharus edwardsii, Parmaturus xaniurus and Proscyllium habereri, the dorsal surface is totally smooth (state 1). The absence of dermal denticles on the dorsal surface of clasper glans may be related to the degree of development of the cover rhipidion and/or the exorhipidi- on and the presence of an open clasper groove. 61 Distribution of dermal denticles on dorsal surface of clasper glans: (0) denticles present only on the ex- orhipidion; (1) denticles over all of the dorsal surface except on rhipidion and terminal dermal cover. (CI = 25; RI = 73). Leigh-Sharpe (1926b) subdivided the genus Scyliorhinus (as Scyllium) into four ‘pseudogenera’ based on charac- ters of the external morphology of the claspers, including the distribution of dermal denticles on dorsal surface of the clasper glans. Only two of the four groups proposed (Alphascyllium and Betascyllium) included species cur- rently valid for Scyliorhinus. According to Leigh-Sharpe (1926b), species allocated to Alphascyllium presented claspers totally covered by dermal denticles, whereas in Betascyllium denticles are restricted to certain areas; no further details were provided by the author. In Cephalos- cyllium sufflans, C. ventriosum, Scyliorhinus spp. (except S. boa, S. cervigoni, S. hesperius and S. retifer) and Ho- lohalaelurus regani, the dermal denticles are present on most of the dorsal surface of the clasper glans with the exception of the rhipidion and the terminal dermal cover (state 1; Fig. 25B, C, E, F). In S. boa, S. cervigoni, S. hes- zse.pensoft.net 368 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus Figure 22. Detail of the oropharyngeal denticles in Cephaloscyllium sufflans, SAIAB 6242, male, 800 mm TL. cb I-V, cerato- branchials IV. Figure 23. Detail of the gill pickax. A, Scyliorhinus haeckelii, UERJ 1691, male, 522 mm TL; B, Cephaloscyllium sufflans, SAIAB 6242, male, 800 mm TL. perius and S. retifer, denticles are absent on cover rhipid- ion, rhipidion and the terminal dermal cover, as well as in Asymbolus rubiginosus, Atelomycterus fasciatus, Auloha- laelurus labiosus, Figaro boardmani, Galeus antillensis and Schroederichthys saurisqualus (state 0; Fig. 25A, D). In Poroderma africanum and P. pantherinum, dermal den- ticles are present only on the ventrolateral and posterior margins of the exorhipidion and medial to the rhipidion. 62 Dermal denticles on the medial border of the exorhi- pidion: (0) absent; (1) present. (CI = 25; RI = 25). Compagno (1988a) mentioned the presence of hook-like dermal denticles arranged in rows on the ventral sur- zse.pensoft.net face of the free margin of the exorhipidion in species of Cephaloscyllium, Halaelurus, Parmaturus, Apristurus, Poroderma and Scyliorhinus. Here, we observed spe- cialised hooks in the claspers of S. torazame (Soares and de Carvalho 2019; Fig. 25F), Figaro boardmani, Galeus antillensis, Halaelurus natalensis, Poroderma panther- inum and Proscyllium habereri, slightly greater dermal denticles, on the ventral surface at the posterior border of the exorhipidion and on the medial margin of the cover rhipidion (state 1). These denticles were not observed in the other taxa examined (state 0). 63 Terminal dermal cover: (0) present; (1) absent. (CI = 50; RI= 0). Zoosyst. Evol. 96 (2) 2020, 345-395 A 369 Figure 24. Coracoid bar; ventral view. A, Scyliorhinus haeckelii, UERJ 1691, male, 522 mm TL; B, Halaelurus natalensis, SAIAB 26951, male, 400 mm TL; C, Holohalaelurus regani, SATAB 25717, male, 610 mm TL; D, Cephalurus cephalus, USNM 221527, female, 285 mm TL. mpc, medial projection of the coracoid bar; Ipe, lateral processes on pectoral girdle. Soares et al. (2015) described the terminal dermal cov- er in Scyliorhinus ugoi, which consists in a membrane situated on the posterior tip of the clasper glans, lack- ing denticles and in contact with the posterior borders of the cover rhipidion and exorhipidion. This structure was illustrated by Springer (1966) in Scyliorhinus torrei (p. 588, fig. 4a) and by Compagno (1988a) in Holoha- laelurus cf. punctatus (fig. 13.14f-g), but they did not propose a name nor a definition for it. A terminal der- mal cover 1s found in most taxa examined (state 0; Fig. 25), with the exception of Apristurus longicephalus and Cephalurus sp. In these taxa, the distal tips of the dorsal and ventral terminal cartilages are evident and uncov- ered (state 1). 64 Extension of the terminal dermal cover: (0) restricted to the distal tip of the clasper glans; (1) extending up to one-third of clasper glans. (CI = 33; RI= 71). Regarding its extension, the terminal dermal cover may be restricted to the distal clasper tip or it extends up to one-third of the clasper glans, covering the posterior borders of the cover rhipidion and exorhipidion. The former condition was observed in Asymbolus rubigi- nosus, Atelomycterus fasciatus, Aulohalaelurus labio- sus, Halaelurus natalensis, Haploblepharus edwardsii, Parmaturus xaniurus, Proscyllium habereri and Schro- ederichthys saurisqualus, in which the terminal dermal cover only reaches the posterior borders of the exorhi- pidion and cover rhipidion. In scyliorhinines, Figaro boardmani, Galeus antillensis and Holohalaelurus re- gani, a more developed terminal dermal cover was ob- served (state 1; Fig. 25). 65 Configuration of the terminal dermal cover: (0) smooth; (1) rough. (CI = 33; RI = 0). Compagno (1988a) described and illustrated the presence of a ‘brush-like papillose structure’ on the distal tip of the clasper glans of Holohalaelurus cf. punctatus, this structure is here considered the terminal dermal cover (as per Soares et al. 2015). The adjective ‘rough’ is used as a substitute for ‘papillose’ by considering that the structure does not present papillae but rugosities. Additionally, we observe that, besides having a different texture, the ter- minal dermal cover projects posteriorly, corresponding to two-thirds of the clasper glans length in Holohalaelurus spp. A terminal dermal cover with similar texture was also observed in Scyliorhinus canicula and S. capensis (Soares and de Carvalho 2019; state 1, Fig. 25B). In the other taxa examined, this structure is smooth and without rugosities (state 0; Fig. 25). 66 Degree of development of the rhipidion: (0) well de- veloped and presenting a prominent posterior mar- gin; (1) reduced and consisting in a narrow strip. (CI = 25; RI=77). Some differences regarding the degree of development of the rhipidion were observed in taxa that have this struc- ture (Leigh-Sharpe 1920, 1921, 1922a, 1922b, 1924a, 1924b, 1924c, 1926a, 1926b). In Cephaloscyllium suf- flans, Scyliorhinus spp. (except S. canicula, S. duhamelii and S. torazame), Asymbolus rubiginosus, Halaelurus natalensis, Holohalaelurus regani and Schroederichthys saurisqualus, a well-developed rhipidion with a promi- nent posterior margin was observed (state 0; Fig. 26A). In zse.pensoft.net 370 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus C eee eee ee = : ‘i eat La hae oe . <= : | oF , ie: f ! a \ ‘ Wen ‘ Hens aay ‘ pata ane ee, is des , A ana I Ahi 3 fd ts ie ie iH Scan We 5 ¥ i ’ Y ane os Seer [65:1] ate Se = ae ri sau Ss ee = aS > fe, NLA Ad ot +h oe ae man ss ee = a: oe ch Soe Anat Soe i ts GEES _ eee il i sy inn i eo << ae roe ae Eee, eae se ‘ exock es a 1 Sane MELE Ke: iP * oes _ re oF ae Sites At ao ! spite Shs sos Se —— ae es oh ras aes “= Rees: Sass = oS oo Z} nd etna A ce Ke Fer ep a in Sort ae earns eae a ae [63:0] Figure 25. Clasper; external morphology. A, Scyliorhinus boa, USNM 221563, 348 mm TL; B, Scyliorhinus canicula, BMNH 1983.8.3.14, 585.7 mm TL; C, Scyliorhinus duhamelii, MCZ S-63, 338.7 mm TL; D, Scyliorhinus retifer, UF 36359, 372 mm TL; E, Scyliorhinus stellaris, BMNH 1976.7.30.10,476 mm TL; F, Scyliorhinus torazame, NSMT 50632, 427.9 mm TL. ch, clasp- er hooks; erh, cover rhipidion; dd, dermal denticles; en, envelope; erh, exorhipidion; hp, hypopyle; rh, rhipidion; tde, terminal dermal cover. Modified from Soares and de Carvalho (2019). Scyliorhinus canicula, S. duhamelii and S. torazame, the 67 Extension of the rhipidion: (0) extending throughout rhipidion is narrow and not prominent, similar to Ate/o- the clasper glans; (1) extending up to one-third of the mycterus fasciatus, Aulohalaelurus labiosus, Poroderma clasper glans. (CI = 25; RI = 73). spp., Figaro boardmani, Galeus antillensis, Haplobleph- arus edwardsii, Parmaturus xaniurus and Proscyllium The extension of the rhipidion varies, depending on the habereri (state 1; Fig. 26B). species examined. In Scyliorhinus spp. (except S. ca- zse.pensoft.net Zoosyst. Evol. 96 (2) 2020, 345-395 nicula, S. duhamelii and S. torazame), Cephaloscyllium sufflans, Asymbolus rubiginosus, Halaelurus natalensis, Holohalaelurus regani and Schroederichthys saurisqua- lus, the rhipidion extends throughout the clasper glans, reaching the posterior border of the cover rhipidion (state 0; Fig. 26A). In Scyliorhinus canicula, S. duhamelii, S. torazame, Atelomycterus fasciatus, Aulohalaelurus la- biosus, Poroderma spp., Figaro boardmani, Galeus antil- lensis, Haploblepharus edwardsii, Parmaturus xaniurus and Proscyllium habereri, the rhipidion extends up to the posterior one-third of the clasper glans, anterior to the posterior border of the cover rhipidion (state 1; Fig. 26B). 68 Cover rhipidion: (0) poorly developed; (1) well devel- oped and medially expanded. (CI = 100; RI = 100). Compagno (1988a) pointed out that the condition ‘clasper groove closed and covered’ would be a primitive charac- ter for Carcharhiniformes related to the absence or pres- ence of a poorly developed cover rhipidion on claspers. He reported the presence of a slightly differentiated and short cover rhipidion, well anterior to the clasper glans, in scyliorhinines, Ga/eus and Holohalaelurus spp. In this study, we observed some differences amongst scyliorhin- ines and the other scyliorhinids, regarding the degree of development of the cover rhipidion. In scyliorhinines, Atelomycterus fasciatus and Aulohalaelurus labiosus (Soares 2020), the cover rhipidion is medially expanded and reaches the exorhipidion and 1s sometimes covered by it anteriorly and both cover the clasper groove (state 1; Fig. 25). In the other taxa examined, the cover rhipidion is nearly straight and restricted to the dorsolateral margin of claspers, lateral to the dorsal terminal 2 cartilage and not covering it (state 0). 69 Exorhipidion: (0) medially expanded; (1) poorly de- veloped. (CI = 50; RI = 75). The exorhipidion 1s a ventromedially situated flap, cov- ering totally or partially the ventral terminal cartilage. In all species of Scyliorhinus, Cephaloscyllium sufflans, Asymbolus rubiginosus, Atelomycterus fasciatus, Aulo- halaelurus labiosus, Figaro boardmani, Galeus antil- lensis, Halaelurus natalensis, Parmaturus xaniurus and Poroderma spp., we observed a well-developed exorhi- pidion totally covering the ventral terminal cartilage and extending to the end of the glans (state 0; Fig. 25). In the other taxa examined, a poorly-developed exorhipid- ion 1s observed (state 1) corresponding to a narrow strip restricted to the posterior portion of the ventral terminal 2 cartilage and not reaching the cover rhipidion medially. 70 Envelope: (0) present; (1) absent. (CI = 14; RI = 54). The envelope is a distinct projection anterior to the ex- orhipidion, which covers the accessory terminal carti- lage, posterior border of the ventral marginal cartilage and anterior border of the ventral terminal cartilage. Ac- 371 cording to our observations, this structure is present in Scyliorhinus boa, S. cervigoni, S. haeckelii, S. retifer, S. torrei and S. ugoi (Soares and de Carvalho 2019), as well as in Apristurus longicephalus, Asymbolus rubiginosus, Figaro boardmani, Halaelurus natalensis, Parmaturus xaniurus and Schroederichthys saurisqualus (state 0; Fig. 25). In Haploblepharus edwardsii, the accessory terminal cartilage 1s covered by a thin membrane, separated from the exorhipidion and the rest of the clasper cover; this membrane is not considered a true envelope. An envelope is also absent in Atelomycterus fasciatus, Aulohalaelurus labiosus, Cephaloscyllium spp., Galeus antillensis, Ho- lohalaelurus regani, Poroderma spp., Proscyllium and other species of Scyliorhinus (state 1). 71 Degree of development of the envelope: (0) poorly de- veloped; (1) medially expanded. (CI = 100; RI = 100). A well-developed envelope, projecting medially and covering the anterior border of the cover rhipidion is observed in Scyliorhinus boa and S. retifer (state 1; Fig. 25A, D). In S. cervigoni, S. haeckelii, S. torrei and S. ugoi, we observed a discrete and poorly-developed en- velope, not covering the cover rhipidion; this condition is also present in Apristurus longicephalus, Asymbolus rubiginosus, Figaro boardmani, Halaelurus natalensis, Parmaturus xaniurus and Schroederichthys saurisqualus (state 0). 72 Accessory terminal cartilage: (0) present; (1) absent. (CI = 20; RI = 60). Jungersen (1899) described this cartilage as a structure situated between the ventral terminal cartilage and the posterior border of the ventral marginal cartilage. In taxa of Rajiformes, the accessory terminal cartilage presents a dorsal extension external to the integument, forming a protractile spine (Compagno 1988a). None of the taxa ex- amined here presented such an extension. Soares (2020) pointed out that this cartilage is usually adjacent to the medial surface of ventral terminal cartilage. In Scyliorhi- nus, only the species S. boa, S. canicula, S. capensis, S. retifer and S. torazame have an accessory terminal car- tilage (state 0; Fig. 27A, B, D, F). This cartilage is also present in Proscyllium and other scyliorhinids, except Holohalaelurus regani and Poroderma spp. (state 1). 73 Accessory dorsal marginal cartilage: (0) present; (1) absent. (CI = 33; RI = 80). Jungersen (1899) described a mobile cartilage situated in the posterior border of the dorsal marginal cartilage (and continuous with it) in Pristiurus melastomus (= Galeus melastomus), naming it the ‘accessory dorsal marginal cartilage’. According to this author, this carti- lage is absent in Scyliorhinus canicula and S. stellaris. Soares et al. (2015, 2016) misidentified the dorsal term1- nal 2 cartilage as the accessory dorsal marginal cartilage zse.pensoft.net Bia Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus tv2 t3 tv2 [66:1] | [67:1] | Figure 26. Detail of the rhipidion. A, Scyliorhinus boa, USNM 221563, 348 mm TL; B, Scyliorhinus canicula, BMNH 1983.8.3.1-4, 585.7 mm TL. Modified from Soares & de Carvalho (2019). rd, dorsal marginal cartilage; rh, rhipidion; rv, ventral marginal cartilage; t3, accessory terminal cartilage; td, dorsal terminal cartilage; td2, dorsal terminal 2 cartilage; tv, ventral terminal cartilage; tv2, ventral terminal 2 cartilage. in S. cabofriensis, S. haeckelii and S. ugoi, because it is situated between the dorsal marginal and terminal carti- lages; however, it is medial to these and is not a prolon- gation of the dorsal marginal cartilage. Thus, we verified for the present study that the accessory dorsal marginal cartilage is indeed absent in Scyliorhinus spp. (Soares and de Carvalho 2019; Soares 2020), as well as in other scyliorhinines (state 1; Fig. 27). This structure, however, is present in other taxa examined, except Galeus antil- lensis (state 0). 74 Dorsal marginal 3 cartilage: (0) absent; (1) present. (CI = 25; RI=0). The dorsal marginal 3 cartilage is situated dorsally and external to the accessory dorsal marginal cartilage or posterior to it. A dorsal marginal 3 cartilage is absent in scyliorhinines and most of the scyliorhinids examined (state 0). In Haploblepharus edwardsii, Halaelurus na- talensis and Holohalaelurus regani, this cartilage 1s pres- ent and is very slender, resembling a shell dorsal to the accessory dorsal marginal cartilage (state 1). zse.pensoft.net 75 Ventral marginal 2 cartilage: (0) present; (1) absent. (CI = 20; RI = 56). The presence of a ventral marginal 2 cartilage was re- ported for Sphyrna by Compagno (1988a), opposite to the accessory dorsal marginal cartilage. This cartilage was observed in Apristurus longicephalus, Halaelurus natal- ensis, Haploblepharus edwardsii, Parmaturus xaniurus, Poroderma spp., and Proscyllium (state 0). A ventral mar- ginal 2 cartilage is absent in Scyliorhinus spp., Cepha- loscyllium spp., Asymbolus rubiginosus, Aulohalaelurus labiosus, Atelomycterus fasciatus, Cephalurus cephalus, Figaro boardmani, Galeus spp., Holohalaelurus regani and Schroederichthys saurisqualus (state 1). 76 Position of the ventral marginal 2 cartilage: (0) con- tinuous to the posterior border of the ventral margin- al cartilage: (1) lateral to the posterior border of the ventral marginal cartilage. (CI = 100; RI = 100). Compagno (1988a) described the ventral marginal 2 car- tilage as a structure situated posteriorly to the ventral Zoosyst. Evol. 96 (2) 2020, 345-395 marginal cartilage, but continuous to it. In Apristurus lon- gicephalus, Halaelurus natalensis, Haploblepharus ed- wardsii and Parmaturus xaniurus, this cartilage present- ed the same condition described by Compagno (1988a), with a trapezoidal shape and covering the ventral terminal 2 cartilage (state 0). A different condition regarding the position of the ventral marginal 2 cartilage was found in Poroderma spp.; in these species, this cartilage is situated laterally to the ventral marginal cartilage and not contin- uous to it (state 1). 77 Dorsal terminal 2 cartilage: (0) present; (1) absent. (CI = 33; RI=0). The dorsal terminal 2 cartilage was described by Jungers- en (1899) and consists of a narrow piece situated in the lateral border of the dorsal terminal cartilage that extends posteromedially to the posterior border of the dorsal mar- ginal cartilage. White (1937) did not include the dorsal terminal 2 cartilage in her illustration of the clasper of Scyliorhinus torazame (pl. 47) and Compagno (1988a) reported its absence in Scyliorhinus. However, a dor- sal terminal 2 cartilage was observed in all species of Scyliorhinus (state 0; Fig. 27), except S. garmani, S. hes- perius and S. meadi (in these species adult males were not available for dissection). This cartilage was observed in Proscyllium and most scyliorhinids examined with the exception of Halaelurus natalensis (state 1). 78 Shape of the dorsal terminal 2 cartilage: (0) elongat- ed and rod-like; (1) rhomboidal. (CI = 50; RI = 75). Variations in the shape of the dorsal terminal 2 cartilage were observed amongst the taxa examined. In Proscylli- um and most scyliorhinids, the dorsal terminal 2 cartilage is a rod-like structure (state 0; Fig. 27), while, in some species of Scyliorhinus (S. cabofriensis, S. capensis, S. cervigoni, S. haeckelii and S. ugoi), the dorsal terminal 2 cartilage is poorly developed and has a rhomboidal shape (state 1; Fig. 27C). 79 Ventral terminal 2 cartilage: (0) present: (1) absent. (CI = 25; RI =0). Jungersen (1899) reported the presence of an elongated element connected to the anterior tip of the ventral termi- nal cartilage and situated above it in claspers of Lamna cornubica (= Lamna nasus), this element was named the ventral terminal 2 cartilage. Jungersen (1899) did not de- scribe this cartilage in Galeus melastomus, Scyliorhinus canicula and S. stellaris, the scyliorhinids he observed. However, a ventral terminal 2 cartilage is present in most species of Scyliorhinus, except S. comoroensis (Com- pagno 1988b; Soares and de Carvalho 2019). Accord- ing to our observations, this cartilage is present in most taxa examined (state 0; Fig. 27) except Cephaloscyllium sufflans, Cephalurus cephalus, Figaro boardmani, Ha- laelurus natalensis and Proscyllium habereri. 3/3 80 Position of the ventral terminal 2 cartilage: (0) ante- riorly situated and sometimes attached to the anterior tip of the ventral terminal cartilage; (1) posteriorly situated, posterior to the half-length of the ventral terminal cartilage. (CI = 50; RI = 50). A different condition from the one described by Jungers- en (1899) concerning the position of the ventral terminal 2 cartilage was observed in Aulohalaelurus labiosus and Poroderma spp. In these species, this cartilage is more posteriorly situated, posterior to the half-length of the ventral terminal 2 cartilage and not attached to the an- terior tip of the ventral terminal cartilage (state 1). In the other taxa examined, the ventral terminal 2 cartilage presents the same condition described by Jungersen (1899) (state 0). 81 Extension of the clasper siphon: (0) extending be- yond the half distance between the coracoid and cloa- ca; (1) shorter than the coracoid-cloaca half distance. (CI = 33; RI = 78). Leigh-Sharpe (1920) proposed the term ‘siphon’ for ‘a sac with extremely muscular walls, situated immediately below the corium of the ventral surface of the abdomen, close to the median line and ending blindly, having no communication with the coelom’ (1920, p. 246). Leigh- Sharpe (1924a) proposed a transformation series between Scyliorhinidae and Carcharhinidae, with 7riakis as an in- termediate link, on the basis of siphon length: Scyliorh- inidae presenting a short siphon and slightly anterior to the pelvic girdle and Carcharhinidae with extremely long siphons reaching the insertion of the pectoral fin in some taxa. Gilbert and Gordon (1972) suggested a relationship between siphon extension and reproductive mode, with Oviparous sharks presenting short siphons and viviparous sharks long siphons. However, a great variation in siphon length was observed herein amongst scyliorhinids, which are reported as oviparous with several descriptions of egg capsules in literature (Springer 1979; Compagno 1988a; Gomes and de Carvalho 1995; Flammang et al. 2007; Flammang et al. 2008; Castro 2011; Ebert and Stehman 2013; Gordon et al. 2016; Silva and Soares 2017; Soares and de Carvalho 2019). Two conditions were observed in the present study: 1) long siphons, extending beyond the coracoid-cloaca half distance; and 11) short siphons, shorter than the coracoid-cloaca half distance. In Pros- cyllium and most scyliorhinids, the longer condition was observed (state 0; Fig. 23A—C), similar to the siphons of Mustelus and Carcharhinus (Leigh-Sharpe 1924a). Short siphons were observed in scyliorhinines, Apristurus lon- gicephalus, Galeus antillensis and Holohalaelurus re- gani (state 1; Fig. 28D, E, F). Colouration 82 Colour pattern composed of saddles: (0) present; (1) absent. (CI = 25; RI = 50). zse.pensoft.net 374 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus B \y | C [72:0] [79:0] tv2 [78:0] [80:0] Figure 27. Clasper; skeleton. A, Scyliorhinus boa, USNM 221563, 348 mm TL; B, Scyliorhinus canicula, BMNH 1983.8.3.1-4, 585.7 mm TL; C, Scyliorhinus capensis, SAIAB 27577, 863 mm TL; D, Scyliorhinus retifer, UF 36359, 372 mm TL; E, Scyliorhi- nus Stellaris, BMNH 1976.7.30.10,476 mm TL; F, Scyliorhinus torazame, NSMT 50632, 427.9 mm TL. end, endstyle; rd, dorsal marginal cartilage; rv, ventral marginal cartilage; t3, accessory terminal cartilage; td, dorsal terminal cartilage; td2, dorsal terminal 2 cartilage; tv, ventral terminal cartilage; tv2, ventral terminal 2 cartilage. Modified from Soares and de Carvalho (2019). zse.pensoft.net Zoosyst. Evol. 96 (2) 2020, 345-395 The presence of transverse bands darker than the back- ground colour over most of the body, known as ‘saddles’, is widespread amongst catsharks. Gomes et al. (2006), in their re-description of Schroederichthys tenuis, proposed three types of saddles: primary saddles, secondary sad- dles and subsaddles; these latter are situated ventrally to the lateral line. Primary saddles, more prominent in re- lation to the background colour, were observed in most species of Scyliorhinus (Soares and de Carvalho 2019; Fig. 29), except in S. duhamelii and S. garmani. In S. tor- rei, these saddles are found only in juvenile specimens. In S. boa and S. retifer, spots and dark lines, slightly dark- er than the background colour, are bordering the saddles (Fig. 29B). In Asymbolus spp., Atelomycterus spp., Ceph- aloscyllium spp., Halaelurus spp., Haploblepharus spp., Proscyllium habereri and Schroederichthys spp., saddles were also observed, varying in number and position (state 0; Fig. 29E). In Apristurus spp., Cephalurus spp., Galeus spp., Holohalaelurus regani, Parmaturus spp., and Poro- derma spp., saddles are absent (state 1; Fig. 29F). 83 Dark spots: (0) present; (1) absent. (CI = 20; RI = 50). Springer (1979) pointed out the relevance of colouration for identification of Scyliorhinus species to the detriment of other features, such as morphometric data and internal morphology. In Scyliorhinus, we observed some differ- ences amongst species regarding the occurrence of dark spots. Dark spots are present in S. boa, S. cabofriensis, S. canicula (Fig. 29A), S. cervigoni, S. duhamelii, S. gar- mani, S. haeckelii, S. stellaris and S. ugoi, whereas they are absent in S. capensis, S. comoroensis, S. hesperius, S. meadi, S. torazame (Fig. 29C) and S. torrei (Fig. 29D; Soares and de Carvalho 2019). In Cephaloscyllium spp.., Poroderma pantherinum (Fig. 29H), Proscyllium habere- ri and most scyliorhinids, dark spots were observed, with the exception of Apristurus longicephalus, Cephalurus cephalus and Parmaturus spp. 84 Dark stripes: (0) absent; (1) present. (CI = 50; RI=0). A colour pattern, composed of dark stripes running in different directions, was observed in Scyliorhinus retifer (Fig. 29B) and Poroderma africanum (state 1; Fig. 29G), differing from all other scyliorhinid species examined. In Scyliorhinus retifer, stripes form polygons and are bor- dering saddles, while in Poroderma africanum, stripes are parallel to the anteroposterior axis and do not form saddles, extending throughout the body. Non-informative (autapomorphic) characters Anterior nasal flaps in Haploblepharus Bell (1993) described anterior nasal flaps as expanded and medially fused, forming a nasal curtain that covers the upper lip, in Haploblepharus. However, according to ei he our observations, nasal flaps in Haploblepharus (Fig. 3C) are not fused, but present the same point of origin, me- dially; this pattern is unique amongst carcharhiniforms. Muscle preorbitalis originating from the posterolateral wall of the nasal capsules The muscle preorbitalis is situated anteriorly to the m. adductor mandibulae and limited posteriorly by the man- dibular ramus of the nerve V (Huber et al. 2011). This muscle originates from the posteroventral wall of the nasal capsules and extends to the orbital notch in most taxa examined. In Holohalaelurus regani, a unique con- dition was observed; m. preorbitalis originates from the posterolateral surface of the nasal capsules. In all other scyliorhinid species, insertion of these muscles is on the muscle adductor mandibulae. Muscle /evator hyomandibulae with undifferentiated muscle fibres Shirai (1992b) described the muscle /evator hyomandib- ulae as united to the m. constrictor hyoideus dorsalis in Carcharhiniformes and separated from it in batoids. In Orectolobiformes and Heterodontus, the muscle /evator hyomandibulae is situated internally to the m. constrictor hyoideus dorsalis, with its ventral portion laterally exposed (Goto 2001). In most of the examined taxa, muscle fibres are differentiated in m. /evator hyomandibulae, internal- ly to the m. constrictor hyoideus dorsalis, originating on the pterotic process of the neurocranium and inserting in the distal tip of the hyomandibular cartilage. In Apristurus longicephalus, the m. levator hiomandibulae seems to be fused to the m. constrictor hyoideus dorsalis or is absent; the same condition is also observed in Squalus acanthias (Marinelli and Strenger 1959; Huber et al. 2011). Origin of the muscle coracomandibularis on the lateral borders of the coracoid bar The m. coracomandibularis is dorsally situated to the muscles intermandibularis and interhyoideus, consisting of a median bundle originating from the m. coracoarcua- lis (most taxa examined) or from the medial surface of the coracoid bar (Apristurus longicephalus, Fig. 11D). According to Shirai (1992b), the first condition is widely distributed amongst neoselachians. Association of the m. coracomandibularis directly with the coracoid bar was reported by Shirai (1992a) for Centroscyllium and Rhi- na and by Goto (2001) for Brachaelurus, Ginglymosto- ma and Stegostoma. Shirai (1996) coded in his character matrix the origin of the m. coracomandibularis on the fascia of the m. coracoarcualis for all carcharhiniforms (his character 51), differing from what we observed in Apristurus longicephalus. Shirai (1996) also coded the origin of this muscle on the coracoid bar or pericardial membrane for Heterodontus, Hexanchus, Heptranchias, Squatina, Squaliformes and some rays. zse.pensoft.net S76 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus [81:1] Figure 28. Detail of the clasper siphon. A, Apristurus longicephalus, HUMZ 170382,475 mm TL; B, Holohalaelurus regani, SAIAB 25717, 610 mm TL; C, Poroderma africanum, SAIAB 25343, 920 mm TL; D, Asymbolus rubiginosus, AMS 1.30393- 004, 527 mm TL; E, Halaelurus natalensis, SALAB 26951, 400 mm TL; F, Haploblepharus edwardsii, AMNH 40988, 480 mm TL. Scale bar: 20 mm. Clasper hooks In Scyliorhinus torazame, we observed specialised hooks in the claspers forming a row that extends from the be- ginning of the ventral marginal cartilage to the terminal dermal cover and running along the medial margin of the exorhipidion (Schimidt 1930; Soares and de Carvalho 2019; Fig. 24F). This arrangement is unique amongst the taxa examined. zse.pensoft.net Phylogenetic reconstruction The phylogenetic analysis of the data matrix (Appendix 2) including 84 morphological characters (five quanti- tative and 79 qualitative) and 35 terminal taxa and the use of implied weighting (k = 3) resulted in three equally most-parsimonious trees with 233 steps, CI = 0.37 and RI = 0.69. A strict consensus was generated and is presented in Figure 30 and its analysis is detailed below. The char- Zoosyst. Evol. 96 (2) 2020, 345-395 SAL Figure 29. Color patterns. A, Scyliorhinus canicula, MNHN 1999-1732, female, 418.5 mm TL; B, Scyliorhinus retifer, UF 36359, male, 372 mm TL; C, Scyliorhinus torazame, NSMT 50632, male, 427.9 mm TL; D, Scyliorhinus torrei, USNM 157852, male, 285 mm TL; E, Atelomycterus fasciatus, CSIRO H1298-7, male, 370 mm TL; F, Parmaturus angelae, MZUSP. 124001, female, 425 mm TL; G, Poroderma pantherinum, SAIAB 34577, male, 640 mm TL; H, Poroderma africanum, SAIAB 25343, male, 920 mm TL. Scale bar: 20 mm. zse.pensoft.net 378 acter matrix was divided into two datasets: in Appendix 2, quantitative characters with absolute and normalised values for each terminal are presented, whereas in Appen- dix 3 only qualitative characters are included. Character listings for clades numbered in Figure 30 are summarised in Appendix 4. The list of synapomor- phies, presented below, begins in Scyliorhininae and pro- gressively continues to less inclusive clades within this family. For each clade, only non-ambiguous synapomor- phies are listed. After each synapomorphy, the number of the referred character and its state changes are shown in brackets. Synapomorphies followed by an asterisk rep- resent unique transformations in the present analysis. A complete list of character transformations is presented in Appendix 5. Relative Bremer support values are shown in Figure 30 below each node. Monophyly of clade 1 The hypothesis of the monophyly of the Scyliorhininae is supported by eight synapomorphies, four of them pro- posed for the first time herein. This clade is composed of Scyliorhinus, Cephaloscyllium and Poroderma. Mono- phyly of the Scyliorhininae was previously proposed by Compagno (1998a), who listed loss of the depressor pal- pebrae nictitantis muscle and loss of the fourth ventral extrabranchial cartilage as synapomorphies for the sub- family (both corroborated herein). However, no cladistic analysis was performed by this author. Later, Iglésias et al. (2005), Human et al. (2006) and Naylor et al. (2012a, 2012b) corroborated the monophyly of this clade using molecular data. 1. Lower diplospondylous vertebral count [char. 2, 0.520—0.660 > 0.280—0.310]. 2. Muscle depressor palpebrae_ nictitantis absent [char. 29, 0 > 1]. 3. Nasal apertures at the same level in nasal capsules [char. 38, 0 > 1]. 4. Articular region of the quadratomandibular joint of Meckel’s cartilage composed by a posterior lingual condyle opposite to the facet [char. 49, 0 > 1]. 5. Three ventral extrabranchial cartilages [char. 56, 0 caauh 6. Terminal dermal cover extending up to one-third of the clasper glans [char. 64, 0 > 1]. 7. Accessory dorsal marginal cartilage absent [char. 73,0 >i, 8. Clasper siphon short and restricted to the pelvic re- gion [char. 81, 0 > 1]. Monophyly of clade 2 The monophyly of Poroderma is supported by five syn- apomorphies. 1. Anterior nasal flap divided into two portions, medi- al and lateral [char. 8, 0 > 1]. zse.pensoft.net Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus 2. Muscular nasal barbel present [char. 10, 0 > 1]. . Accessory terminal cartilage absent [char. 72, 0 > 1]. 4. Ventral terminal 2 cartilage posteriorly situated, posterior to the half-length of the ventral terminal cartilage [char. 80, 0 > 1]. 5. Colour pattern not composed of transverse saddles [char. 82,0 > 1]. Uo Poroderma africanum is characterised by the following autapomorphies: 1. Projected flap present on the upper lip margin [char. 17,0 > 1]. Independently acquired tn Scyliorhinus spp. 2. Muscle coracomandibularis inserting on the medial aspect of antimeres of Meckel’s cartilage [char. 30, O> 1]. 3. Colour pattern composed of stripes [char. 84, 0 > 1]. Poroderma pantherinum is characterised by the follow- ing autapomorphy: 1. Dermal denticles present on the medial border of the exorhipidion [char. 62, 0 > 1]. Monophyly of clade 3 The monophyly of the clade formed by Scyliorhinus and Cephaloscyllium is supported by seven synapomorphies. Compagno (1988a) already had proposed a close rela- tionship between both genera, listing the loss of upper labial furrows and pseudopera rudimentary and absent as synapomorphies. Here, absence of the upper labial fur- rows is not considered a synapomorphy for this clade and features of the pseudopera were not included in the pres- ent analysis as this structure is poorly defined. 1. Mesonarial crest prominent [char. 9, 0 > 1]. 2. One upper labial cartilage [char. 19, 0 > 1]. 3. Lateral processi rastriformis similar to dermal papillae in length [char. 53, 0 > 1]. 4. Coracoid bar with a well-developed medial projec- tion, more than twice the size of its lateral portion [char. 58, 0 > 1]. 5. Rhipidion well-developed and presenting a promi- nent posterior margin [char. 66, 1 > O]. 6. Rhipidion extending throughout the clasper glans [char 6721. >:0]: 7. Ventral marginal 2 cartilage absent [char. 75, 0 > 1]. Monophyly of clade 4 The monophyly of Cephaloscyllium is supported by the following synapomorphies: 1. Higher monospondylous vertebral count [char. 1, 0.540—0.650 > 0.730]. 2. Higher upper tooth row count [char. 3, 0.230—0.350 > 0.450—0.480]. Zoosyst. Evol. 96 (2) 2020, 345-395 2 73 4 4 5 70 99 F 6 3 100 8 65 89 7 49 10 9 11 a 100 + 12 97 |13 14 19 15 27 16 84 L17 7 > 400 Proscyllium habereri ‘Asymbolus rubiginosus Halaelurus natalensis Schroederichthys saurisqualus Holohalaelurus regani Haploblepharus edwardsii Cephalurus cephalus Apristurus longicephalus Parmaturus xaniurus Figaro boardmani Galeus antillensis Atelomycterus fasciatus Aulohalaelurus labiosus Poroderma africanum Poroderma pantherinum Cephaloscyllium umbratille Cephaloscyllium isabella Cephaloscyllium sufflans Cephaloscyllium variegatum Scyliorhinus boa Scyliorhinus hesperius Scyliorhinus retifer Scyliorhinus stellaris Scyliorhinus cabofriensis Scyliorhinus cervigoni Scyliorhinus haeckelii Scyliorhinus ugol Scyliorhinus comoroensis Scyliorhinus meadi Scyliorhinus torrei Sscyliorhinus capensis Scyliorhinus torazame Scyliorhinus garmani scyliorhinus canicula Scyliorhinus duhamelii a) 9) ~< a O 7 Ze z Zz > lm ~” >) S O By) = = Cc ~Y Figure 30. Strict consensus cladogram of the three equally most-parsimonious trees (L = 233, CI = 0.37, RI = 0.69). Number of clades shown above branches. Values of relative Bremer support shown below branches. 3. Higher lower tooth row count [char. 4, 0.230—0.370 > 0.490-0.530]. 4. Lower labial furrow absent [char. 16, 0 > 1]. 5. Postoral grooves present [char. 20, 0 > 1]. 6. Origin of second dorsal fin anterior to half-length of the anal-fin base [char. 24, 0 > 1]. 7. Muscle bundles of muscle coracohyoideus sepa- rated by at least one-half the width of each bundle [char. 32,0 > 1]. Cephaloscyllium umbratile is hypothesised as sister group of the clade formed by the species C. isabella, C. sufflans and C. variegatum (clade 5) and characterised by the following autapomorphies: 1. Higher diplospondylous vertebral count [char. 2, 0.280—0.310 > 0.670—1 .00]. 2. Higher upper tooth row count [char. 3, 0.450—0.480 > 0.570—1.00]. 3. Higher lower tooth row count [char. 4, 0.490—0.530 > 0.580—1.00]. Monophyly of clade 5 The clade, formed by species Cephaloscyllium isabella, C. sufflans and C. variegatum, 1s characterised by the fol- lowing synapomorphy: 1. Lower diplospondylous vertebral count [char. 2, 0.280-0.310 > 0.130-0.170]. zse.pensoft.net 380 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus Cephaloscyllium isabella is hypothesised as the sister group of the clade formed by the species C. sufflans and C. variegatum (clade 6) and characterised by the follow- ing autapomorphy: 1. Lower diplospondylous vertebral count [char. 2, 0.130—0.170 > 0.060—0.080]. Monophyly of clade 6 This clade is formed by Cephaloscyllium sufflans and C. variegatum and 1s characterised by the following synapo- morphy: 1. Oropharyngeal denticles large and forming rows on internal face of gill components [char. 54, 0 > 1]. No unique autapomorphies were found for Cephalos- cyllium variegatum. Cephaloscyllium sufflans is charac- terised by the following autapomorphy: 1. Higher monospondylous vertebral count [char. 1, 0.730 > 0.810]. Monophyly of clade 7 The monophyly of Scyliorhinus is supported by five syn- apomorphies in all equally most-parsimonious trees. The genus 1s divided into two main clades: S. boa, S. hesperius and S. retifer (clade 8) and a clade for all remaining species. 1. Lower values for upper tooth row count [char. 3, [0.230-0.350 > 0.210—-0.220]. 2. Lower values for lower tooth row count [char. 4, 0.230-0.370 > 0.220]. 3. Projected flap present on the upper lip margin [char. Fe OS Tilt 4. Pelvic apron present [char. 21, 1 > 0]. 5. Ephyseal notch present [char. 41, 0 > 1]. Monophyly of clade 8 This clade is formed by the species Scyliorhinus boa, S. hesperius and S. retifer and is characterised by the fol- lowing synapomorphies: 1. Envelope present on clasper [char. 70, 1 > 0]. 2. Envelope medially expanded [char. 71, 0 > 1]. No unique autapomorphies were found in the present analysis for Scyliorhinus boa. Scyliorhinus hesperius 1s characterised by the following autapomorphy: 1. Colour pattern without dark spots [char. 83, 0 > 1]. Scyliorhinus retifer is characterised by the following autapomorphy: zse.pensoft.net 1. Colour pattern composed of dark stripes [char. 84, O> 1]. Monophyly of clade 9 The monophyly of the clade, formed by S. stellaris, S. cabofriensis, S. canicula, S. capensis, S. comoroensis, S. cervigoni, S. duhamelii, S. garmani, S. haeckelii, S. mea- di, S. torazame, S. torrei and S. ugoi, 1s supported by the following synapomorphy: 1. Accessory terminal cartilage absent [char. 72, 0 > 1]. Scyliorhinus stellaris is hypothesised as the sister group of all remaining species of Scyliorhinus (cited above) and characterised by the following autapomorphy: 1. Higher intestinal valve count [char. 5, 0.420 > 0.670—0.750]. In some trees: 1. Higher monospondylous vertebral count [char. 1, 0.460-0.540 > 0.580-0.690]. Monophyly of clade 10 This clade is formed by S. cabofriensis, S. cervigoni, S. haeckelii and S. ugoi and characterised by the following synapomorphy: 1. Terminal dorsal 2 cartilage rhomboidal [char. 78, 0 eil|! In some trees: 1. Lower monospondylous vertebral count [char. 1, 0.540—0.580 > 0.420-0.460]. 2. Higher intestinal valve count [char. 5, 0.420 > 0.250]. Scyliorhinus cabofriensis is hypothesised as the sis- ter group of S. haeckelii, S. cervigoni and S. ugoi, but no unique autapomorphies were found for this species. Monophyly of clade 11 The monophyly of the clade formed by S. cervigoni, S. haeckelii and S. ugoi 1s supported by the following syn- apomorphy: 1. Envelope present on clasper [char. 70, 1 > 0]. No unique autapomorphies were found in the present analysis for S. haeckelii and S. ugoi. Scyliorhinus cervig- oni 1s characterised by the following autapomorphy: Zoosyst. Evol. 96 (2) 2020, 345-395 1. Dermal denticles restricted to the exorhipidion on dorsal surface of clasper glans [char. 61, 1 > O]. Monophyly of clade 12 The monophyly of the clade, formed by S. canicula, S. capensis, S. comoroensis, S. duhamelii, S. garmani, S. meadi, S. torazame and S. torrei, is supported by the fol- lowing synapomorphy: 1. Colour pattern without dark spots [char. 83, 0 > 1]. Scyliorhinus comoroensis and S. meadi are hypoth- esised as the sister group of the remaining species. Scyliorhinus comoroensis 1s characterised by the follow- ing synapomorphies: 1. Higher upper tooth row count [char. 3, 0.280—0.310 > 0.470]. 2. Ventral terminal 2 cartilage absent [char. 79, 0 > 1]. Scyliorhinus meadi is characterised by the following autapomorphy: 1. Higher monospondylous vertebral count [char. 1, 0.460 > 0.540—0.770]. Monophyly of clade 13 The monophyly of the clade, formed by S. canicula, S. capensis, S. duhamelii, S. garmani, S. torazame and S. torrei, 1S Supported by the following synapomorphy: 1. Pelvic inner margins almost entirely fused. [char. Zoe 2: Scyliorhinus torrei is hypothesised as sister group of clade 14 (see below) and is characterised by the following autapomorphies: 1. Lower monospondylous vertebral count [char. 1, 0.350—0.540 > 0.080—0.270]. 2. Lower upper tooth row count [char. 3, 0.170—0.220 > 0.00—0.120]. 3. Lower values for lower tooth row count [char. 4, 0.190—0.220 > 0.010—0.180]. 4. Envelope present on clasper [char. 70, 1 > 0]. Monophyly of clade 14 This clade is formed by S. canicula, S. capensis, S. du- hamelii, S. garmani and S. torazame and is characterised by the following synapomorphy: 1. Accessory terminal cartilage present [char. 72, 1 > O]. revel Scyliorhinus capensis is hypothesised as sister group of the clade 15 (see below) and is characterised by the following autapomorphies: 1. Higher monospondylous vertebral count [char. 1, 0.350—-0.540 > 0.620-0.690]. 2. Lower values for lower tooth row count [char. 4, 0.210—-0.220 > 0.260-0.770]. 3. Terminal dermal cover rough [char. 65, 0 > 1]. 4. Dorsal terminal 2 cartilage rhomboidal [char. 78, 0 eal)|, Monophyly of clade 15 The monophyly of the clade, formed by S. canicula, S. duhamelii, S. garmani and S. torazame, is supported by the following synapomorphies: 1. Rhipidion reduced and consisting of a narrow strip [char. 66, 0 > 1]. 2. Rhipidion extending up to 1/3 of the clasper glans length [char. 67, 0 > 1]. Scyliorhinus torazame is hypothesised as the sister group of the clade formed by S. canicula, S. duhamelii and S. garmani, but no unique autapomorphies were found for this species. Monophyly of clade 16 The monophyly of the clade, formed by S. canicula, S. duhamelii and S. garmani, is supported by the following synapomorphies: 1. Anterior nasal flap entirely covering excurrent ap- erture, posterior nasal flap and upper lip [char. 6, Qi 2 | 2. Colour pattern composed of dark spots [char. 83, 1 > 0]. Scyliorhinus garmani 1s hypothesised as the sister group of the clade formed by S. canicula and S. duhamelii and this species 1s characterised by the following autapomorphy: 1. Higher monospondylous vertebral count [char. 1, 0.350—0.540 > 0.770]. Monophyly of clade 17 Monophyly of the clade, formed by S. canicula and S. duhameiii, is supported by: 1. Anterior nasal flaps distant from each other by less than half of their length [char. 7, 0 > 1]. 2. Posterior nasal flap laterally situated to the excur- rent aperture [char. 13, 0 > 1]. 3. Nasoral groove present [char. 14, 0 > 1]. zse.pensoft.net 382 Scyliorhinus canicula is characterised by the following autapomorphy: 1. Terminal dermal cover rough [char. 65, 0 > 1]. Scyliorhinus duhamelii is characterised by the follow- ing autapomorphies: 1. Lower intestinal valve counts [char. 5, 0.420 > 0.250]. 2. Accessory terminal cartilage absent [char. 72, 0 > 1]. Discussion Phylogeny of Scyliorhinus species The phylogenetic relationships of species of the subfam- ily Scyliorhininae on the basis of morphological data and inferred from a numerical cladistic study including all Scyliorhinus species, are here presented for the first time. The monophyly of Scyliorhininae is supported by the ab- sence of the muscle depressor palpebrae nictitantis, nasal apertures at the same level on nasal capsules, articular region of the quadratomandibular joint of Meckel’s carti- lage, characterised by a posterior lingual condyle opposite to the facet, three ventral extrabranchial cartilages, ter- minal dermal cover extending to one-third of the clasper glans, absence of accessory dorsal marginal cartilage and clasper siphon short and restricted to the pelvic region. Compagno (1988a) listed the absence of the muscle depressor palpebrae nictitantis and the loss of the fourth ventral extrabranchial cartilage as synapomorphies for Scyliorhininae. He also mentioned the reduction of clasp- ers components and of the second dorsal fin as diagnostic characters for this subfamily. However, Compagno (1988a) did not mention which parts of the clasper are reduced or absent in the subfamily and his descriptions of carchar- hiniform genera did not present detailed information on the skeletal anatomy of the copulatory organs. Soares (2020) reported the presence of an accessory dorsal mar- ginal cartilage in many scyliorhinid taxa, but not in species of Cephaloscyllium and Poroderma. Regarding the great variability of sizes and position between dorsal fins, their relative sizes were not included in the present analysis as they are highly influenced by ontogeny and preservation, especially in specimens of Apristurus and Cephalurus. Scyliorhinus is hypothesised to be the sister group of Cephaloscyllium, sharing with it the presence of only one upper labial cartilage, lateral processi rastriformis similar in size to the dermal papillae, coracoid bar with a well-developed medial projection corresponding to more than twice the size of its lateral portion, a well-developed rhipidion presenting a prominent posterior margin and extending throughout the clasper glans and the absence of the ventral marginal 2 cartilage. A closer relationship between Scyliorhinus and Cephaloscyllium was also pro- zse.pensoft.net Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus posed by Compagno (1988a) who listed the absence of upper labial furrows and pseudopera absent or rudimen- tary in claspers as diagnostic characters. The monophyly of Scyliorhinus is supported by the presence of a projected flap on the upper lip margin, of a pelvic apron and an ephyseal notch at the posterior bor- der of the anterior fontanelle on the neurocranium. The presence of a pelvic apron is observed in Scyliorhinus, Asymbolus and Holohalaelurus, being more developed and extending to at least two thirds or almost the entire length of pelvic inner margins in Scyliorhinus species. The presence of an ephyseal notch on the neurocranium of Scyliorhinus species is unique amongst scyliorhinines. The presence of a projected flap is the main character used by many authors to identify species of Scyliorhinus (Garman 1913; Bigelow and Schroeder 1948; Springer 1966, 1979; Compagno 1988a; Compagno et al. 2005; Ebert et al. 2013), but is proposed as a synapomorphy for these species for the first time herein. This flap is also present in Poroderma africanum, but this species is hy- pothesised as the sister group of P. pantherinum, shar- ing with it the following characters: anterior nasal flap divided into two portions (medial and lateral), presence of a muscular nasal barbel, distance between internal ca- rotid foramina greater than the distance between internal carotid and stapedial foramina, absence of an accessory terminal cartilage, ventral terminal cartilage posteriorly situated (posterior to the half-length of the ventral termi- nal cartilage) and colour pattern not composed of trans- verse saddles. Additionally, Poroderma is hypothesised as sister group of the clade formed by Cephaloscyllium + Scyliorhinus. Springer (1979) pointed out that the unique configu- ration of the nasoral region of S. canicula would be suf- ficient to guarantee the allocation of all other species of the genus in a distinct taxon; we note here that the same configuration is present in S. duhamelii. Some authors also made comments on the similarities observed in the nasoral region of Scyliorhinus canicula and species of At- elomycterus and Haploblepharus, highlighting the need of a more detailed examination and the investigation of phylogenetic relationships amongst these taxa (Com- pagno 1988a; Bell 1993). According to our results, S. canicula and S. duhamelii are distinguished from species of Atelomycterus and Haploblepharus by the absence of upper labial furrows and the presence of posterior nasal flaps (vs. upper furrows present and posterior flaps absent in Atelomycterus and Haploblepharus). Additionally, S. canicula and S. duhamelii share with their congeners the presence of a flap on the upper lip that projects lateral- ly, covering the lower labial furrows and the presence of a pelvic apron in males. Similarities amongst S. canic- ula and S. duhamelii and species of Atelomycterus and Haploblepharus have been suggested as being the result of adaptative convergence to benthic habits (Bell 1993). However, these characters are not present in other dem- ersal scyliorhinids. Zoosyst. Evol. 96 (2) 2020, 345-395 Species of Cephaloscyllium, here examined, shared the following synapomorphies: absence of an upper labi- al furrow, presence of a postoral groove, origin of a sec- ond dorsal fin anterior to the half-length of the anal fin, muscle bundles of muscle coracohyoideus well separated along all their extension and higher values for monospon- dylous vertebrae, upper and lower tooth row counts. Pos- toral grooves are observed in all species of Cephaloscyl- lium with varied extensions, but no flap or labial furrow is found in any of them (17 species are considered valid; Fricke et al. 2020). Notches near the lower edge of the mouth are observed and illustrated for specimens of C. signorum (Last et al. 2008), C. variegatum and C. zebrum (Last and White 2008) and can be confused with lower labial furrows without a detailed examination. Clasper morphology contributed important charac- ters that helped elucidate the phylogenetic relationships among species of Scyliorhinus and other scyliorhinines (21 characters from the clasper were included in the present analysis). Amongst the most relevant charac- ters are the following: dermal denticles along the dorsal surface of the clasper, degree of development of the en- velope, configuration of terminal dermal cover, occur- rence of accessory terminal and ventral terminal 2 car- tilages and shape of dorsal terminal 2 cartilage. A closer relationship between S. boa and S. retifer was proposed by Goode and Bean (1896) and Garman (1913), based on colour pattern of both species, which is corroborat- ed here by clasper morphology. Both species share the presence of a well-developed and medially-expand- ed envelope that lacks dermal denticles, covering the anterior portion of the cover rhipidion; this condition is also observed in S. hesperius. The clade, formed by Scyliorhinus cabofriensis, S. cervigoni, S. haeckelii and S. ugoi, is supported by the presence of a reduced dorsal terminal 2 cartilage. These species are distributed in the Southern Atlantic Ocean, off the eastern coast of Bra- zil (S. cabofriensis, S. haeckelii and S. ugoi) and west coast of Africa (S. cervigoni), with records in similar latitudes, suggesting a common evolutionary history that may date from the formation of the Atlantic Ocean. According to Springer (1966: p. 597), species of Scyliorhinus distributed in the Western Central Atlantic may form a ‘compact infrageneric group’, as they are more similar to each other than they are to species in the Eastern Atlantic and Western Pacific. According to our results, however, this hypothesis is not corroborated be- cause of the closer phylogenetic relationships of species from the Western Central Atlantic with those from other regions. Scyliorhinus torrei is hypothesised here to being more closely related to S. capensis (Southeastern Atlan- tic), S. torazame (Western Pacific), S. canicula and S. du- hamelii (North-eastern Atlantic and Mediterranean Sea) by sharing a well-developed pelvic apron with pelvic in- ner margins almost entirely fused. Scyliorhinus boa, S. hesperius and S. retifer form a clade hypothesised as the sister group of all species of Scyliorhinus. 383 The impact of morphological characters Characters from the nasoral region, dermal denticles, claspers, vertebrae and intestinal counts were revealed to be extremely important to shed light on the phylogenetic relationships amongst scyliorhinines and may contribute to future phylogenetic analyses concerning scyliorhinids. A more detailed examination of the nasal flaps and la- bial furrows allows for the identification of differences amongst the genera Atelomycterus, Haploblepharus and Scyliorhinus and clarifies questions related to the distri- bution and variation of characters amongst species of Scyliorhinus (e.g. S. canicula and S. duhamelii). Data from tooth morphology of catsharks are scarce and the only study that reported tooth characters for Scyliorhinidae is Herman et al. (1990). Besides this study, information on sexual and ontogenetic heterodon- ty are found only for a few species (Brough 1937; Na- kaya 1975; Springer 1966; Compagno 1988a; Gomes and Tomas 1991; Litvinov 2003; Soares and de Carval- ho 2019) and for some scyliorhinid species, heterodonty seems to be absent (Weigmann et al. 2018). In this study, males and females, adults and juveniles were not avail- able for some species and, thus, it was not possible to investigate the influence of sexual dimorphism and on- togeny in tooth characters (mainly regarding the number of cusplets and degree of development of striae). As a consequence, characters of tooth morphology were not included in the present analysis. Despite the relevance of characters associated to claspers in species identification and phylogenetic anal- yses, information on the internal anatomy of these or- gans are found only for some species and mainly in classical works about clasper morphology (Jungersen 1899; Leigh-Sharpe 1920, 1921, 1922, 1924a, b; Com- pagno 1988a), being absent in most species descriptions and taxonomic reviews (Human 2006b; Séret and Last 2007; Last et al. 2008; Last and White 2008; Nakaya et al. 2013; amongst others). Soares (2020) provided detailed descriptions of clasper structures in almost all catshark genera and demonstrated the uselfulness of claspers for taxonomic and systematic purposes. We highlight here the importance of including clasper descriptions in tax- onomic studies. Morphology and molecular data The monophyly of the subfamily Scyliorhininae is cor- roborated by the present study, as well as by phylogenetic analyses, based on molecular data (Iglésias et al. 2005; Human et al. 2006; Naylor et al. 2012a, 2012b). How- ever, morphology and molecular-based studies diverge on the hypotheses of relationships amongst Scyliorhinus, Cephaloscyllium and Poroderma (Fig. 31). In this study, Cephaloscyllium is hypothesised as the sister group of Scyliorhinus, based on clasper and skeletal characters. zse.pensoft.net 384 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus Poroderma Cephaloscyllium Scyliorhinus B —_____ Cephaloscyllium Poroderma Scyliorhinus Figure 31. Phylogenetic hypotheses of the subfamily Scyliorhininae. A, morphology- based hypothesis (present work; Compagno, 1988a); B, molecular-based hypothesis (Naylor et al., 2012a, 2012b). Naylor et al. (2012a, 2012b) hypothesised a closer rela- tionship between Poroderma and Scyliorhinus, analysing the NADH2 mitochondrial gene and conducting a mod- el-Bayesian phylogenetic analysis. The monophyly of the genus Scyliorhinus is supported here and also by Human et al. (2006) and Naylor et al. (2012a, 2012b). The phylogeny of Iglésias et al. (2005) resolved Scyliorhinus as paraphyletic, with a weakly-sup- ported relationship between S. torazame and Cephalos- cyllium umbratile. These authors analysed only few spe- cies of Scyliorhinus and no species of Poroderma was included in the analysis (Iglésias et al. 2005). In the present study, we contribute to the understand- ing of the phylogenetic relationships amongst Scyliorhi- nus species. In recent molecular studies, only few species of Scyliorhinus were included and, therefore, little infor- mation on infrageneric relationships could be obtained (Iglésias et al. 2005; Naylor et al. 2012a, 2012b). Nev- ertheless, a closer relationship between S. retifer and S. stellaris was recovered by Naylor et al. (2012a), which agrees with the results presented here. Despite the contributions presented here for the phy- logeny of Scyliorhininae, there is a great need to review the taxonomy of Cephaloscyllium, including the exam- ination of clasper morphology in its species. Deeper considerations on the monophyly of Scyliorhinidae and the phylogenetic relationships amongst scyliorhinids and other taxa were not performed here, since additional taxa of other carcharhiniform families should be included in a broader phylogenetic analysis. Taxonomic reviews, de- tailed morphological studies and cladistic analyses, based on morphological and molecular data, are necessary to improve our understanding of the phylogenetic relation- ships amongst scyliorhinids and other carcharhiniforms. Conclusions ¢ The monophyly of Scyliorhininae is supported by four characters proposed by Compagno (1988a) and other four proposed for the first time; ¢ Scyliorhinus is hypothesised to be the sister group of Cephaloscyllium, sharing with it the presence of only one upper labial cartilage, lateral processi rastriformis similar in size to the dermal papillae, zse.pensoft.net coracoid bar with a well-developed medial projec- tion corresponding to more than twice the size of its lateral portion, a well-developed rhipidion pre- senting a prominent posterior margin and extending throughout the clasper glans and the absence of the ventral marginal 2 cartilage. ¢ The monophyly of Scyliorhinus is supported by the presence of a projected flap on the upper lip mar- gin, of a pelvic apron and an ephyseal notch at the posterior border of the anterior fontanelle on the neurocranium. ¢ Characters from the nasoral region, dermal denti- cles, claspers, vertebrae and intestinal counts were revealed to be extremely important to shed light on the phylogenetic relationships amongst scyliorhin- ines and may contribute to future phylogenetic anal- yses concerning scyliorhinids and carcharhiniforms. ¢ Results presented here mostly agree with those ob- tained in recent phylogenetic analyses, but further work integrating molecular and morphological data is still needed. Acknowledgements The authors wish to acknowledge Barbara Brown (AMNH), David Catania(CAS), Rob Robbins (FLMNH), Toshio Kawai (HUMZ), Karsten Hartel and Andrew Wil- liston (MCZ), Patrice Pruvost (MNHN), Marcelo Brito (MN/UFRJ), Alessio Datovo (MZUSP), Oliver Crimmen and Ralf Britz (BMNH), Sven Kullander (NRM), Masa- nori Nakae and Gento Shinohara (NSMT), Ofer Gon and Nkosinathi Mazungula (SAIAB), Albe Bosman (SAM), Ulisses Gomes and Hugo Santos (UERJ), Ricardo Rosa (UFPB), Otto Gadig (UNESP, Sao Vicente), José Ortiz (USAC), Lynne Parenti, Jeffrey Williams, Kris Murphy and Sandra Raredon (USNM), Peter Bartsch (ZMB), Ralf Thiel (ZMH), and Marcus Krag (ZMUC) for permis- sion to visit the collections and examine the specimens under their care. José Ortiz (USAC) for sending pho- tos of specimens of Scyliorhinus hesperius. Monica de Toledo-Piza (USP), Otto Gadig (UNESP, Sao Vicente), Ulisses Gomes (UERJ), Alessio Datovo (MZUSP) and Rodrigo Caires for their comments and contributions on the writing of this paper. Enio Matos and Phillip Lentit- Zoosyst. Evol. 96 (2) 2020, 345-395 akis (IBUSP) for the SEM images and Giulia Baldaconi for the neurocranium illustrations. The first author was supported by CAPES (Codigo de Financiamento 001) and FAPESP (processes 2014/20316-5, 2015/21314-9, 2016/22214-0); the second author by a grant from CNPq (304615/2011-0). References de Beer GR (1937) The Development of the Vertebrate Skull. Universi- ty of Chicago Press, Chicago. Bell MA (1993) Convergent evolution of nasal structure in sed- entary elasmobranchs. Copeia 1993(1): 144-158. https://doi. org/10.2307/1446305 Bigelow HB, Schroeder WC (1948) Fishes of the Western North At- lantic. Part I. Lancelets, Cyclostomes and Sharks. Yale University, New Haven. Cappetta H (2012) Handbook of Paleoichthyology. Vol. 3E: Chondrich- thyes. Mesozoic and Cenozoic Elasmobranchii: Teeth. Verlag Dr. Friedrich Pfeil. de Carvalho MR (1996) Higher-lever elasmobranch phylogeny, basal squaleans, and paraphyly. In: Stiassny, MLJ, Parenti LR, Johnson GD (Eds) Interrelationships of fishes. Academic Press, New York, 35-62. https://doi.org/10.1016/B978-012670950-6/50004-7 Castro JI (2011) The sharks of North America. Oxford University Press, New York. Ciena AP, Rangel BS, Bruno CEM, Miglino MA, de Amorim AF, Rici REG, Watanabe I (2016) Morphological Aspects of Oral Denti- cles in the Sharpnose Shark Rhizoprionodon lalandii (Miller & Henle, 1839) (Elasmobranchii, Carcharhinidae). Anatomia, His- tologia, Embryologia 45(2): 109-114. https://doi.org/10.1111/ ahe.12178 Compagno LJV (1988a) Sharks of the Order Carcharhiniformes. Princ- eton University Press, Princeton. Compagno LJV (1988b) Scyliorhinus comoroensis sp. n., a new cat- shark from the Comoro Islands, western Indian Ocean (Carchar- hiniformes, Scyliorhinidae). Bull. Mus. Natn. d’Histoire Naturelle 10(3): 603-625. Compagno LJV (1999) Endoskeleton. In: Hamlett WC (Ed.) Sharks, Skates, and Rays: The Biology of Elasmobranch Fishes. The Johns Hopkins University Press, Baltimore, 69-92. Compagno LJV, Stevens JD (1993a) Atelomycterus fasciatus n.sp., a new catshark (Chondrichthyes: Carcharhiniformes: Scyliorhinidae) from tropical Australia. Records of the Australian Museum 45(2): 147-169. https://doi.org/10.3853/}.0067-1975.45.1993.19 Compagno LJV, Stevens JD (1993b) Galeus gracilis n. sp., a new saw- tail catshark from Australia, with comments on the systematics of the genus Galeus Rafinesque, 1810 (Carcharhiniformes: Scyliorhi- nidae). Records of the Australian Museum 45(2): 171-194. https:// doi.org/10.3853/].0067-1975.45.1993.133 Compagno LJV, Dando M, Fowler S (2005) Sharks of the World. Princ- eton University Press, Princeton. Daniel JF (1934) The elasmobranch fishes. 3 ed. University of Califor- nia Press, Berkeley. Danois E (1913) Contribution a I’étude de systématique et biologique poisons de la Manche Occidentale. In: Joubin L (Ed.) Annales de I’ Institut océanographique. Tome V. Monaco et Cie, Paris, 214 pp. 385 Ebert DA, Fowler S, Compagno LJV, Dando M (2013) Sharks of the World: A Fully Illustrated Guide. Wild Nature Press, Plymouth. Ebert DA, Stehmann M (2013) Sharks, batoids, and chimaeras of the North Atlantic. FAO Species Catalogue for Fishery Purposes. FAO, Rome. Farris J (1969) A successive approximations approach to charac- ter weighting. Systematic Zoology 18(4): 374-385. https://doi. org/10.2307/2412182 Flammang BE, Ebert DA, Caillet GM (2007) Egg cases of the genus Apristurus (Chondrichthyes: Scyliorhinidae): Phylogenetic and ecological implications. Zoology 110(4): 308-317. https://doi. org/10.1016/j.zool.2007.03.001 Flammang BE, Ebert DA, Cailliet GM (2008) Reproductive biology of deep-sea catsharks (Chondrichthyes: Scyliorhinidae) in the east- ern North Pacific. Environmental Biology of Fishes 81(1): 35-49. https://doi.org/10.1007/s10641-006-9162-9 Fricke R, Eschmeyer WN, Van der Laan R (2020) Eschmeyer’s Catalog of Fishes: genera, species, references. Electronic version accessed 03/23/2020. http://researcharchive.calacademy.org/research/ichthy- ology/catalog/fishcatmain.asp Garman S (1913) The Plagiostomia (Sharks, skates and rays). Memoirs of the Museum of Comparative Zoology at Harvard College 36. Gilbert PW, Gordon WH (1972) The clasper-siphon sac mechanism in Squalus acanthias and Mustelus canis. Comparative Biochemistry and Physiology — Part A, Physiology 42(1): 97-119. https://doi. org/10.1016/0300-9629(72)9037 1-4 Gill TN (1862) Analytical synopsis of the order of Squali; and revision of the nomenclature of the genera. Ann. Lye. Nat. Hist. 7: 367-408 https://doi.org/10.1111/j.1749-6632.1862.tb00166.x. Gledhill DC, Last PR, White WT (2008) Resurrection of the genus Figaro Whitley (Carcharhiniformes: Scyliorhinidae) with the de- scription of a new species from northeastern Australia. CSIRO Ma- rine and Atmospheric Research Paper 22: 179-188. Goloboff PA (1993) Estimating character weights during tree search. Cla- distics 9: 83-91. https://doi.org/10.1111/j.1096-003 1.1993 tb00209.x Goloboff PA (1995) Parsimony and weighting: a reply to Turner and Zan- dee. Cladistics 11: 91-104. https://doi.org/10.1111/j.1096-0031.1995. tb00006.x Goloboff PA (1999) Analyzing large data sets in reasonable times: Solu- tions for composite optima, Cladistics, 15: 415-428. https://doi. org/10.1111/j.1096-0031.1999.tb00278.x Goloboff P, Farris J, Nixon KC (2003) T.N.T. Tree Analysis Using New Technology. http://www.zmuc.dk/public/phylogeny/tnt Goloboff PA, Mattoni CM, Quinteros AS (2006) Continuous characters analyzed as such. Cladistics 22: 589-601. https://doi.org/10.1111/ j.1096-0031.2006.00122.x Goloboff PA, Carpenter JM, Arias JS, Esquivel DRM (2008) Weighting against homoplasy improves phylogenetic analysis of morpholog- ical data sets. Cladistics 24: 1-16. https://doi.org/10.1111/j.1096- 0031.2008.00209.x Goloboff PA, Farris JS, Nixon KC (2008) TNT, a free program for phy- logenetic analysis. Cladistics 24: 774-786. https://doi.org/10.1111/ j.1096-0031.2008.00217.x Gomes UL, de Carvalho MR (1995) Egg Capsules of Schroederichthys tenuis and Scyliorhinus haeckelii (Chondrichthyes, Scyliorhinidae). Copeia 1995(1): 232—236. https://doi.org/10.2307/1446823 Gomes UL, Peters GO, Carvalho MR, Gadig OBF (2006) Anatomical investigation of the slender catshark Schroederichthys tenuis Spring- er, 1966, with notes on intrageneric relationships (Chondrichthyes: zse.pensoft.net 386 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus Carcharhiniformes: Scyliorhinidae). Zootaxa 1119: 29-58. https:// doi.org/10.11646/zootaxa.1119.1.2 Gordon CA, Hood AR, Ellis JR (2016) Descriptions and revised key to the eggcases of the skates (Rajiformes: Rajidae) and catsharks (Carcharhiniformes: Scyliorhinidae) of the British Isles. Zootaxa 4150(3): 255-280. https://doi.org/10.11646/zootaxa.4150.3.2 Goto T (2001) Comparative Anatomy, Phylogeny and Cladistic Clas- sification of the order Orectolobiformes (Chondrichthyes, Elasmo- branchii). Memoirs of the Graduate School of Fisheries Sciences Hokkaido University 48: 1-100. Hennig W (1950) Grundziige einer Theorie der phylogenetischen Sys- tematik. Deutscher Centralverlag, Berlin. Hennig W (1965) Phylogenetic systematics. Annu. Rev. Entomol. 10: 97-116. https://doi.org/10.1146/annurev.en.10.010165.000525 Hennig W (1966) Phylogenetic Systematics. University of Illinois Press, Urbana (translated by D. D. Davis & R. Zangerl). Herman J, Hovestad-Euler M, Hovestad DC (1990) Contributions to the study of the comparative morphology of teeth and other relevant ichthyodorulites in living superspecific taxa of Chondrichthyan fish- es. Part A: Selachii. No. 2b: Order: Carcharhiniformes — Familiy: Scyliorhinidae. Bulletin de I’Institut Royal des Sciences Naturelles de Belgique, Biologie 60: 181—230. Huber DR, Soares MC, de Carvalho MR (2011) Cartilaginous fishes cranial muscles. In: Farrell AP (Ed.) Encyclopedia of Fish Physiol- ogy: From Genome to Environment. Academic Press, San Diego, 449-462. https://doi.org/10.1016/B978-0-12-374553-8.00238-0 Human BA (2006a) A taxonomic revision of the catshark genus Po- roderma Smith, 1837 (Chondrichthyes: Carcharhiniformes: Scyliorhinidae). Zootaxa 1229: 1-32. https://doi.org/10.11646/zoo- taxa.1229.1.1 Human BA (2006b) A taxonomic revision of the catshark genus Ho- lohalaelurus Fowler 1934 (Chondrichthyes: Carcharhiniformes: Scyliorhinidae), with descriptions of two new species. Zootaxa 1315: 1-56. https://doi.org/10.11646/zootaxa.1315.1.1 Human BA (2007) A taxonomic revision of the catshark genus Hap- loblepharus Garman 1913 (Chondrichthyes: Carcharhiniformes: Scyliorhinidae). Zootaxa 1451: 1-40. https://doi.org/10.11646/zoo- taxa.1451.1.1 Human BA, Owen EP, Compagno LJV, Harley EH (2006) Testing mor- phologically based phylogenetic theories within the cartilaginous fishes with molecular data, with special reference to the catshark family (Chondrichthyes; Scyliorhinidae) and the interrelationships within them. Molecular Phylogenetics and Evolution 39(2006): 384-391. https://doi.org/10.1016/j.ympev.2005.09.009 Iglésias SP, Lecointre G, Sellos DY (2005) Extensive paraphylies within sharks of the order Carcharhiniformes inferred from nuclear and mitochondrial genes. Molecular Phylogenetics and Evolution 34(2005): 569-583. https://doi.org/10.1016/j.ympev.2004. 10.022 Jordan DS, Evermann BW (1896) The fishes of North and Middle America. Bull. U.S. Natl. Mus. 47: 1-1240. Jungersen HFE (1899) On the apendices genitales in the greenland shark Somniosus microcephalus (Bl. Schn.) and another selachians. Danish Ingolf Expedition. Vol. II. Bianco luno, Copenhagen. https:// doi.org/10.5962/bh1 title. 14334 Last PR, Gomon MF, Gledhill DC (1999) Australian spotted catsharks of the genus Asymbolus (Carcharhiniformes: Scyliorhinidae). Part 2: descriptions of three new, dark-spotted species. CSIRO Marine Laboratories 239: 19-35. zse.pensoft.net Last PR, Séret B, White WL (2008) New swellsharks (Cephaloscyllium: Scyliorhinidae) from the Indo—Australian region. CSIRO Marine and Atmospheric Research Paper 22: 129-146. Last PR, White WL (2008) Two new saddled swellsharks (Cephalos- cyllium: Scyliorhinidae) from eastern Australia. CSIRO Marine and Atmospheric Research Paper 22: 159-170. Leible M, Dittus DM, Belmar CGG (1982) Atlas Anatomico de Pintar- roja Schroederichthys chilensis (Guichenot, 1848) (Chondrichthyes: Scyliorhinidae). Vol. II. Sistemas: Muscular, Esqueletico, Respirato- rio y Digestivo. Pontificia Universidad Catolica de Chile, Talcahuano. Leigh-Sharpe WH (1920) The comparative morphology of the sec- ondary sexual characters of elasmobranch fishes. The claspers, clasper siphons, and clasper glands. Memoir I. Journal of Mor- phology and Physiology 34(2): 245-265. https://doi.org/10.1002/ jmor. 1050340202 Leigh-Sharpe WH (1921) The comparative morphology of the sec- ondary sexual characters of elasmobranch fishes. The claspers, clasper siphons, and clasper glands. Memoir II. Journal of Mor- phology and Physiology 35(2): 359-380. https://doi.org/10.1002/ jmor. 1050350203 Leigh-Sharpe WH (1922a) The comparative morphology of the sec- ondary sexual characters of elasmobranch fishes. The claspers, clasper siphons, and clasper glands. Memoir HI. Journal of Mor- phology and Physiology 36 (2): 191-198. https://doi.org/10.1002/ jmor. 1050360203 Leigh-Sharpe WH (1922b) The comparative morphology of the sec- ondary sexual characters of holocephali and elasmobranch fishes. The claspers, clasper siphons, and clasper glands. Memoir V. Jour- nal of Morphology and Physiology 36(2): 221-243. https://doi. org/10.1002/jmor. 1050360205 Leigh-Sharpe WH (1924a) The comparative morphology of the sec- ondary sexual characters of elasmobranch fishes. The claspers, clasper siphons, and clasper glands. Memoir VI. Journal of Mor- phology and Physiology 39(2): 553-566. https://doi.org/10.1002/ jmor. 1050390210 Leigh-Sharpe WH (1924b) The comparative morphology of the sec- ondary sexual characters of elasmobranch fishes. The claspers, clasper siphons, and clasper glands. Memoir VI. Journal of Mor- phology and Physiology 39(2): 567-577. https://doi.org/10.1002/ jmor.1050390211 Leigh-Sharpe WH (1926a) The comparative morphology of the sec- ondary sexual characters of elasmobranch fishes. The claspers, clasper siphons, and clasper glands. Memoir VIII. Journal of Mor- phology and Physiology 42(1): 307-320. https://doi.org/10.1002/ jmor. 1050420109 Leigh-Sharpe WH (1926b) The comparative morphology of the second- ary sexual characters of elasmobranch fishes. The claspers, clasper siphons, and clasper glands. Memoir IX. Journal of Morphology and Physiology 42(1): 321-334. https://doi.org/10.1002/jmor. 1050420110 Leigh-Sharpe WH (1926c) The comparative morphology of the sec- ondary sexual characters of elasmobranch fishes. The claspers, clasper siphons, and clasper glands. Memoir X. Journal of Mor- phology and Physiology 42(1): 335-348. https://doi.org/10.1002/ jmor.1050420111 Linnaeus C (1758) Systema Naturae, 10 ed. Maisey JG (1984) Higher elasmobranch phylogeny: a look at the evi- dence. J. Vert. Paleontol. 4(3): 359-371. https://doi.org/10.1080/02 724634. 1984.10012015 Zoosyst. Evol. 96 (2) 2020, 345-395 Marinelli W, Strenger A (1959) Vergleichende Anatomie und Morphol- ogie der Wirbeltiere HI Lieferung (Squalus acanthias). Franz Deu- tick, Vienna. Moss SA (1972) The feeding mechanism of sharks of the fami- ly Carcharhinidae. J. Zool. London 167: 423-436. https://doi. org/10.1111/j.1469-7998.1972.tb01734.x Motta PJ, Wilga CD (1995) Anatomy of the feeding apparatus of the lemon shark, Negaprion brevirostris. Journal of Morphology 226(3): 309-329. https://doi.org/10.1002/jmor. 1052260307 Mufioz CR (1985) Sobre la clasificacion tipoldgica del esqueleto dér- mico de escualos (Chondrichthyes). Miscellania Zoologica 17: 283-285. Nakaya K (1975) Taxonomy, comparative anatomy and phylogeny of Japanese catsharks, Scyliorhinidae. Memoirs of the Faculty of Fish- eries, Hokkaido University 23: 1-94. Nakaya K, Inoue S, Ho H-C (2013) A review of the genus Cephaloscylli- um (Chondrichthyes: Carcharhiniformes: Scyliorhinidae) from Tai- wanese waters. Zootaxa 3752: 101-129. https://doi.org/10.11646/ zootaxa.3752.1.8 Nakaya K, Séret B (2000) Re-description and taxonomy of Pentan- chus profundicolus Smith & Radcliffe, based on a second speci- men from the Philippines (Chondrichthyes, Carcharhiniformes, Scyliorhinidae). Ichthyological Research 47(4): 373-378. https:// doi.org/10.1007/BF02674265 Naylor GJP, Caira JN, Jensen K, Rosana KAM, Straube N, Lakner C (2012) Elasmobranch phylogeny: A mitochondrial estimate based on 595 species. In: Carrier JC, Musick JA, Heithaus MR (Eds) Biology of Sharks and Their Relatives, 2"! ed. CRC Press, Boca Raton, 31—56. Naylor GJP, Caira JN, Jensen K, Rosana KAM, White WT, Last PR (2012) A sequence-based approach to the identifcation of shark and ray species and its implications for global elasmobranch diversity and parasitology. Bull. Am. Mus. Nat. Hist. 367, 1-262. https://doi. org/10.1206/754.1 Nelson GJ (1970) Pharyngeal denticles (placoid scales) of sharks, with notes on the dermal skeleton of vertebrates. American Museum No- vitates 2415: 1-53. Nelson GJ, Platnick NI (1981) Systematics and Biogeography: Cladis- tics and Vicariance. Columbia University Press. Nelson JS, Grande TC, Wilson MVH (2016) Fishes of the World. 5 ed. John Wiley & Sons, New Jersey. https://doi. org/10.1002/9781119174844 Nixon KC, Carpenter JM (1993) On outgroups. Cladistics 9(4): 423— 426. https://doi.org/10.1111/).1096-0031.1993.tb00234.x Rangel BS, Wosnick N, Hammerschlag N, Ciena AP, Kfoury JR, Rici REG (2017) A preliminary investigation into the morphology of oral papillae and denticles of blue sharks (Prionace glauca) with infer- ences about its functional significance across life stages. Journal of Anatomy 203(3): 389-397. https://doi.org/10.1111/joa.12574 Regan CT (1908) A synopsis of the sharks of the family Scyliorh- inidae. Ann. Mag. Nat. Hist. 8(1): 453-465. https://doi. org/10.1080/00222930808692434 Reif WE (1982) Morphogenesis and function of the squamation in sharks. 1. Comparative functional morphology of shark scales, and ecology of sharks. Neues Jahr. Geol. Palaontol. 164, 172-183. Reif WE (1985) Squamation and ecology of sharks. Cour. Forsch. Inst. Senckenberg 78: 1-101. Sabaj MH (2016) Standard symbolic codes for institutional resource collections in herpetology and ichthyology: an Online Reference. S67 Version 6.5 (16 August 2016). American Society of Ichthyologists and Herpetologists, Washington, DC. http://www.asih.org/ Sato K, Nakaya K, Yorozu M (2008) Apristurus australis sp. nov., anew long-snout catshark (Chondrichthyes: Carcharhiniformes: Scyliorh- inidae) from Australia. CSIRO Marine and Atmospheric Research Paper 22: 113-122. Schmidt PJ (1930) A Selachian clasper with a hundred hooks. Copeia 1930: 48-50. https://do1.org/10.2307/1435684 Séret, B, Last PR (2007) Four new species of deepwater catsharks of the genus Parmaturus (Carcharhiniformes: Scyliorhinidae) from New Caledonia, Indonesia and Australia. Zootaxa 1657: 23-39. https:// doi.org/10.11646/zootaxa.1657.1.2 Shirai S (1992a) Identity of extra branchial arches of Hexanchiformes (Pisces, Elasmobranchii). Bulletin of the Faculty of Fishery of the Hokkaido University 43 (1): 24-32. Shirai S (1992b) Squalean Phylogeny: a new framework of ‘Squaloid’ sharks and related taxa. Sapporo: Hokkaido University Press. Shirai S (1996) Phylogenetic interrelationships of Neoselachians (Chon- drichthyes: Euselachii). In: Stiassny MLJ, Parenti LR, Johnson GD (Eds) Interrelationships of fishes. Academic Press, New York, 9-34. https://doi.org/10.1016/B978-012670950-6/50003-5 Silva JPCB, de Carvalho MR (2015) Morphology and phylogenetic sig- nificance of the pectoral articular region in elasmobranchs. Zoologi- cal Journal of the Linnean Society 2015: 1-44. Silva JPCB, Soares KDA (2017) Morphology and phylogenetic signifi- cance of the siphon sac in Scyliorhinidae (Carcharhiniformes: Elas- mobranchii: Chondrichthyes). Il International Symposium on Phy- logeny and Classification of Neotropical Fishes. Londrina, Brazil. Soares KDA (2020) Comparative anatomy of the clasper of catsharks and its phylogenetic implications (Chondrichthyes: Carcharhini- formes: Scyliorhinidae). Journal of Morphology 2020, 1-17. https:// doi.org/10.1002/jmor.21123 Soares KDA, Gadig OBF, Gomes UL (2015) Scyliorhinus ugoi, a new species of catshark from Brazil (Chondrichthyes: Carcharhini- formes: Scyliorhinidae). Zootaxa 3937(2): 347-361. https://doi. org/10.11646/zootaxa.3937.2.6 Soares KDA, Gomes UL, de Carvalho MR (2016) Taxonomic review of catsharks of the Scyliorhinus haeckelii group, with the description of a new species (Chondrichthyes: Carcharhiniformes: Scyliorhi- nidae). Zootaxa 4066(5): 501-534. https://doi.org/10.11646/zoot- axa.4066.5.1 Soares KDA, de Carvalho MR (2019) The catshark genus Scyliorhinus (Chondrichthyes: Carcharhiniformes: Scyliorhinidae): taxonomy, morphology and distribution. Zootaxa 4601(1): 1-147. https://doi. org/10.11646/zootaxa.4601.1.1 Springer S (1966) A review of Western Atlantic cat sharks, Scyliorhin- idae, with descriptions of a new genus and five new species. Fish. Bull. US. Fish Wildl. Serv. 65(3): 581-624. Springer S (1979) A revision of the catsharks, Family Scyliorhini- dae. NOAA Tech. Rep. NMFS 422. https://doi.org/10.5962/bhl. title.63029 Springer VG, Garrick JAF (1964) A survey of vertebral numbers in sharks. Proceedings U.S. National Museum 116: 73-96. https://doi. org/10.5479/s1.00963801.116-3496.73 Weigmann S (2016) Annotated checklist of the living sharks, batoids and chimaeras (Chondrichthyes) of the world, with a focus on bio- geographical diversity. Journal of Fish Biology 88(3): 837-1037. https://doi.org/10.1111/)fb.12874 zse.pensoft.net 388 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus Weigmann S, Kaschner CJ, Thiel R (2018) A new microendemic species of the deep-water catshark genus Bythaelurus (Carcharhiniformes, Pentanchidae) from the northwestern Indian Ocean, with inves- tigations of its feeding ecology, generic review and identification key. PLoS ONE 13(12): e0207887. https://doi.org/10.1371/journal. pone.0207887 Appendix 1 List of examined material Apristurus longicephalus. HUMZ 170382, male, 475 mm TL (Japan). Asymbolus rubiginosus. AMS 1.30393-004, male, 527 mm TL (Australia), AMS 1.34899-002, female, 390 mm TL (Australia). Atelomycterus fasciatus. CSIRO H1298-7, male, 370 mm TL (Australia), MZUSP 118095, female, 363 mm TL (western Australia). Aulohalaelurus labiosus. ZMH 2-1989, female, 480 mm TL, male, 572 mm TL (32°S, 115°30'E). Cephaloscyllium isabella. AMNH 59832, male, 370 mm TL (Sagami, Japan), USNM 320594, female, 390 mm TL (New Zealand); C. sufflans. SAIAB 6242, male, 800 mm TL (South Africa); C. umbratile. USP uncat- alogued, male, 409 mm TL (Taiwan); C. variegatum. AMS 1.43762-001, female, 670 mm TL (Australia), AMS I.24039-007, male, 235 mm TL (Australia). Cephalurus cephalus. USNM 221527, female, 285 mm TL (15°4.30'S, 75°45'W); HUMZ 174830, male, 230 mm TL (no locality data). Figaro boardmani. CSIRO H989-5, female, 465 mm TL (northern Australia), MZUSP 118096, male, 486 mm FL.G5°55.6'S,. 150°1.9'B). Galeus antillensis. UF 77853, female, 370 mm TL, male, 385 mm TL (18°33'N, 65°25'W). Halaelurus natalensis. SALAB 26951, male, 400 mm TL (South Africa). Haploblepharus edwardsii. AMNH 40988, male, 480 mm TL (6°S, 12°40'E); BMNH 1953.5.10.3, male, 720 mm TL (Cape of Good Hope, South Africa). Holohalaelurus regani. SALAB 25717, male, 610 mm TL (South Africa). Parmaturus xaniurus. CAS 232152, female, 450 mm TL (California, United States). Proscyllium habereri. CAS 57189, male, 410 mm TL (27°30'N, 121°30'E). Poroderma africanum. SAJAB 25343, male, 920 mm TL (Cape Town, South Africa); P pantherinum. SATIAB 34577, male, 640 mm TL (Cape Town, South Africa). Schroederichthys saurisqualus. UERJ uncatalogued, fe- male, 564mm TL, male, 580 mm TL (no locality data), ZMH 106492, female, 577 mm TL (30°7'S, 47°58'W, 520 m depth). Scyliorhinus boa. NSMT 30514, male, 179.9 mm TL (7°36'N, 52°26'W, 33 m depth); NSMT 30516A, male, 210.1 mm (7°33'N, 54°10'W); USNM 221532, male, zse.pensoft.net White EG (1936) A classification and phylogeny of the elasmobranch fishes. American Museum Novitates 837: 1-16. White EG (1937) Interrelationships of the elasmobranchs with a key to the order Galea. Bulletin of the American Museum of Natural History 74(2): 25-138. 500 mm TL (11°09'N, 74°26'W); USNM 221562, female, 185 mm TL (15°36'N, 61°13'W); USNM 221563, male, 348 mm TL (12°17'N, 72°40'W); USNM 221564, male, 488 mm TL (17°17'N, 62°23'W). Scyliorhinus cabofriensis. VERJ 1425, female, 319 mm TL (Cabo Frio, Rio de Janeiro, south-eastern Brazil); UERJ 1427, female, 446 mm TL (Cabo Frio, Rio de Janeiro, south-eastern Brazil), UERJ 1582, female, 415 mm TL (Cabo Frio, Rio de Janeiro, south-eastern Brazil); UERJ 1694, male, 412 mm TL (Cabo Frio, Rio de Janeiro, south-eastern Brazil); UERJ 1702, male, 468 mm TL (Cabo Frio, Rio de Janeiro, south-eastern Brazil). Scyliorhinus canicula. BMNH_ 1860.4.22.36-37, male, 512 mm TL, male, 374.7 mm TL (Lisbon, Portugal); BMNH 1888.5.23.48, female, 673.9 mm TL (Kerre- ra, United Kingdom); BMNH 1983.8.3.1-4, male, 585.7 mm TL (Dale Roads, Pembrokeshire, Unit- ed Kingdom); CAS 20612, female, 354 mm TL, fe- male, 361 mm TL, male, 439 mm TL (Naples, Italy); MNHN 1997-0450, female, 575 mm TL (Pas-de-Cal- ais, France, 50°1'N, 1°6'E); NRM 7550, female, 227.2 mm TL (37°57.2'N, 21°4.13'E), NRM 7551, male, 248.7 mm TL (37°57.2'N, 21°4.13'E); NRM 7552, male, 286.3 mm TL (37°57.2'N, 21°4.13'E); NRM 7553, female, 342.4 mm TL (37°57.2'N, 21°4.13'E); NRM 21745, female, 591.3 mm TL (Bohtslan, Swe- den); NRM 46988, female, 662 mm TL (Skagerrak,B Sweden); NRM 49164, male, 677.4 mm TL (Skag- errak, Sweden); NRM 50183, male, 582.4 mm TL (Southern Bohtislan, Sweden); NRM 50450, female, 468.6 mm TL (Skagerrak, Sweden); USNM 221218, female, 334 mm TL, female, 399 mm TL, male, 388 mm TL, male, 383 mm TL (35°41'N, 5°13'W); USNM 221464, male, 345 mm TL, male, 403 mm TL (37°17'N, 10°29'E); USNM 221470, female, 438 mm TL (38°8'N, 10°23.30'E); USNM 221509, female, 317 mm TL, female, 687 mm TL, male, 434 mm TL, male, 419 mm TL (25°28'N, 06°29'E); USNM 221600, male, 459 mm TL, female, 254 mm TL, female, 189 mm TL (35°28'N, 6°31'E); USNM 221601, female, 229 mm TL, male, 247 mm TL (35°12.50'N, 6°33.30'E); USM 221602, female, 258 mm TL (35°09'N, 6°32.20'W): USNM 221603, male, 207 mm TL (35°12.50'N, 6°33.30'W); USNM 221604, male, 306 mm TL (35°12.50'N, 6°33.30'W); USNM 221605, female, 350 mm TL, female, 308 mm TL, male, 346 mm TL (35°41'N, 12°30'W); USNM 221615, female, 354 mm Zoosyst. Evol. 96 (2) 2020, 345-395 TL (37°21.30'N, 10°43'E); USNM 221616, female, 367 mm TL (37°7.30'N, 10°41'E). Scyliorhinus capensis. BMNH_ 1900.11.6.18, female, 898 mm TL (Cape Town, South Africa); BMNH 1935.5.2.54, female, 558.8 mm TL (Cape Town, South Africa); CAS 31455, female, 382 mm TL (Cape of Good Hope, South Africa); SAIAB 12159, female, 224 mm TL (East London, 33°S, 27°9'E); SAIAB 26440, female, 843 mm TL (35°23'6"S, 22°4'E); SATAB 27577, male, 863 mm TL (35°37'12"S, 15°23'42"E); SAM 38774, female, 670 mm TL, male, 686 mm TL (Eastern Cape, South Africa); SAM 38775, male, (Eastern Cape, South Africa). Scyliorhinus cervigoni. USNM 221596, female, 251 mm (10°36'S, 13°12'E); USNM 221598, male, 333 mm TL (11°25'N, 17°21"W); USNM 221599, male, 283 mm TL (10°44'N, 17°06'W); USNM 221617, female, 337 mm TL (6°31'N, 11°29'W). Scyliorhinus comoroensis. MNHN_ 1984—0701, male, 457.2 mm TL (Moroni, Comoro Islands, 400 m depth); MNHN 1991-0420, male, 175 mm TL, female, 181.5 mm TL (Madagascar, 13°45.8'S, 47°38.5'E, 430-700 m depth). Scyliorhinus duhamelii. USNM 203744, male, 399.6 mm TL (36°57'N, 10°28'E, 64-75 m depth); USNM 221645, male, 400 mm TL (42°42.48'N, 17°58.50'E). Scyliorhinus garmani. USNM 43749, female, 267.2 mm TL (“East Indies”, probably Philippines). Scyliorhinus haeckelii. AC.UERJ 1420, male, 365 mm TL (Cabo Frio, Rio de Janeiro, southeastern Brazil); AC.UERJ 1421, female, 412 mm TL (Cabo Frio, Rio de Janeiro, southeastern Brazil); AC.UERJ 1422, fe- male, 379 mm TL (Cabo Frio, Rio de Janeiro, south- eastern Brazil); AC.UERJ 1423, male, 478 mm TL (no locality data); UERJ 1496.1, female, 361 mm TL (Ita- jai, Santa Catarina, Southern Brasil); UERJ 1496.2, female, 367 mm TL (Itajai, Santa Catarina, Southern Brazil); UERJ 1573, female, 297 mm TL (Parana, Southern Brazil); UERJ 1574, female, 371 mm TL (Parana, Southern Brazil); UERJ 1689, male, 566 mm TL (Southern Brazil); UERJ 1690, female, 467 mm TL (Southern Brazil); UERJ 1691, male, 522 mm TL (Rio de Janeiro, Southern Brazil); UERJ 1695, female, 494 mm TL (Southern Brazil); UERJ 1696, female, 451 mm TL (Southern Brazil); UERJ 1697, male, 491 mm TL (Southern Brazil); UERJ 1698, male, 454 mm TL (Southern Brazil); UERJ 1704, male, 425 mm TL (Southern Brazil); UERJ 2202, male, 444 mm TL (Southern Brazil). Scyliorhinus hesperius. CAS 65844, male, 354 mm TL (12°35'N, 82°21'W); USNM 187688, female, 288 mm TL, female, 316 mm TL (16°45'N, 81°27'W); USNM_ 187728, female, 338 mm TL (14°10'N, 81°55'W); USNM 187731, male, 305 mm TL (9°N, 81°23'W); USNM_ 188732, female, 425 mm TL (9°03'N, 81°22'W); USNM 402344, male, 290 mm TL (12°16'N, 72°40'W); USNM 405705, male, 348 mm TL (9°N, 81°23'W). 389 Scyliorhinus meadi. USNM 188049, male, 267 mm TL (28°21'N, 79°51'W); USNM 188050, female, 239 mm TL, male, 175 mm TL (28°31'N, 79°51'W); USNM 188051, male, 190 mm TL (29°44'N, 80°12'W); USNM 221570, male, 180 mm TL (29°1.5'N, 79°56.5'W); USNM 221571, male, 204 mm TL (29°23'N, 79°56.5'W); USNM 221594, male, 271 mm TL (24°48'N, 79°17'W). Scyliorhinus retifer. AMNH 19453, female, 361 mm TL (United States); MCZ 125401, female, 381 mm TL (39°58'N, 70°54'W); UF 28525, female, 500 mm TL (20°43'N, 92°25.8'W); UF 36359, male, 372 mm TL (36°30'N, 74°45'W);, UF 41734, female, 500 mm TL (Gulf of Mexico, 26°N); USNM 26745, male, 340 mm TL (37°26'N, 74°19'W); USNM 84501, male, 449 mm TL, male, 318 mm TL, female, 355 mm TL, female, 291 mm TL (37°03'N, 74°31.40'W); USNM_ 121954, male, 443 mm TL, male, 448 mm TL (Cape Henry, Vir- ginia); USNM 157865, female, 475 mm TL (29°10'N, 88°13'W); USNM 158480, male, 428 mm TL, male, 319 mm TL, female, 297 mm TL (34°39'N, 75°05'W); USNM 187725, male, 465 mm TL, male, 415 mm TL (38°43'N, 73°08'W); USNM 188067, female, 537 mm TL, fe- male, 374 mm TL, female, 241 mm TL (29°03.30'N, 88°28 .30'W): USNM 188069, female, 451 mm TL, male, 374 mm TL (28°57.30'N, 88°39.30'W); USNM 188073, female, 298 mm TL, male, 261 mm TL, female, 287 mm TL, male, 242 mm TL (29°11'N, 88°07'W); USNM 188074, female, 513 mm TL (29°15'N, 87°45.30'W),; USNM 188075, male, 440 mm TL, male, 437 mm TL (28°54'N, 88°51'W); USNM 221469, male, 410 mm TL, male, 420 mm TL (36°54'N, 74°39'W); USNM 221579, female, 200 mm TL (29°02'N, 88°34.5'W); USNM 221580, male, 252 mm TL, male, 298 mm TL (29°30'N, 87°10'W); 221593, male, 338 mm TL, male, 308 mm TL (29°58'N, 80°8.30'W); USNM 221606, female, 476 mm TL (29°25'N, 87°22'W), USNM 221644, female, 389 mm TL, female, 277 mm TL (24°23'N, 82°42'W),; USNM_ 371557, female, 365 mm TL (27°53'N, 85°13'W); USNM 387819, male, 412 mm TL, male, 405 mm TL (38°24.10'N, 73°26.13'W). Scyliorhinus stellaris. NRM 8989, female, 528.3 mm TL (Sicilia, Italy); NRM 8993, female, 236.3 mm TL, fe- male, 174.6 mm TL (Nice, France); NRM 8995, male, 342.4 mm TL (Sicilia, Italy); USNM 28461, female, 317 mm TL (Livorno, Italy); USNM 34352, male, 358 mm TL (Venice, Italy); USNM 221693, female, 483 mm TL, female, 430 mm TL (45°30'N, 13°32'E). Scyliorhinus torazame. HUMZ 117496, male, 401.6 mm TL (Shimoda, Shizuoka Prefecture, Japan); HUMZ 39459, male, 494.4 mm TL (Hakodate, Hokkaido, Ja- pan); HUMZ 40047, male, 381.3 mm TL, male, 382 mm TL (no locality data); HUMZ 113575, female, 352.8 mm TL (Shimoda, Shizuoka Prefecture, Japan); MCZ 35309, male, 470 mm TL, male, 480 mm TL (Ja- pan); MCZ 61163, female, 357 mm TL, male, 324 mm TL (South Korea); NSMT 66232, male, 246.1 mm TL (Sagami-nada, Tateyama, Japan); NSMT 66387, fe- zse.pensoft.net 390 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scy/liorhinus male, 297.3 mm TL, female, 289.7 mm TL (36°30.9'N, 140°59.6'E, 250 m depth); USNM 161525, female, 423 mm TL, female, 413 mm TL (Hakodate, Japan). Scyliorhinus torrei. CAS 65845, female, 255 mm TL (23°34'N, 79°07'W); USNM_ 157845, male, 283 mm TL, female, 259 mm TL (22°59'N, 79°17'W): USNM 157852, male, 285 mm TL, female, 252 mm TL, male, 289 mm TL, male, 283 mm TL, male, 285 mm TL (22°55'N, 79°27'W); USNM 187685, female, 114 mm TL, male, 145 mm TL, female, 257 mm TL (23°05'N, 78°49'W); USNM 187713, female, 215 mm TL (23°34'N, 79°07'W); USNM 187940, female, 294 Appendix 2 mm TL (28°08'N, 77°52'W); USNM 372729, male, 277 mm TL (Playa Santa, Porto Rico). Scyliorhinus ugoi. UERJ 1426, female, 513 mm TL (Ba- hia, Brazil); UERJ 1722, female, 600 mm TL (Salva- dor, Bahia, Brazil); UERJ 1723, female, 427 mm TL (Brazil, between Pernambuco and Northern Rio de Ja- neiro); UERJ 1725, male, 530 mm TL (Brazil, between Southern Bahia and Northern Rio de Janeiro); UERJ 1726, female, 597 mm TL (Brazil, between Pernambu- co and Northern Rio de Janeiro); UERJ 2179, male, 415 mm TL (Southern Bahia, north-eastern Brazil); USNM 221611, male, 432 mm TL (15°22'N, 61°27'W). Matrix summarising quantitative characters used in the phylogenetic study. 1, monospondylous vertebral counts; 2, diplospondylous vertebral counts; 3, upper tooth row counts; 4, lower tooth row counts; 5, intestinal valve counts. Absolute values and normalised ones (in parentheses) are given. Terminal Proscyllium habereri Apristurus longicephalus Asymbolus rubiginosus Atelomycterus fasciatus Aulohalaelurus labiosus Cephaloscyllium isabella C. sufflans C. umbratile C. variegatum Cephalurus cephalus Figaro boardmani Galeus antillensis Halaelurus natalensis Haploblepharus edwardsii Holohalaelurus regani Parmaturus xaniurus Poroderma africanum P. pantherinum Schroederichthys saurisqualus Scyliorhinus boa S. cabofriensis S. canicula S. capensis S. cervigoni S. comoroensis S. duhamelii S. garmani S. haeckelii S. hesperius S. meadi S. retifer S. stellaris S. torazame S. torrei S. ugol zse.pensoft.net 1 38 (0.38) 30-33(0.08-0.19) ? 39-46(0.42-0.69) 45-46(0.65-0.69) 45-48(0.65-0.77) 49(0.77) 47-54(0.73-1) 44-47(0.61-0.73) 28-35(0-0.27) 35-38(0.27-0.38) 33-39(0.19-0.42) 31-33(0.11-0.19) 33-40(0.19-0.46) 28-33(0-0.19) 38-39(0.38-0.42) 42-45(0.54-0.65) 32-46(0.15-0.69) 35-40(0.27-0.46) 39-42(0.42-0.54) 37-39(0.35-0.42) 35-40(0.27-0.46) 44-46(0.61-0.69) 40-45(0.46-0.65) 40(0.46) 35-37(0.27-0.35) 48(0.77) 36-40(0.31-0.46) 39-42(0.42-0.54) 46-48(0.69-0.77) 38-42(0.38-0.54) 43-46(0.58-0.69) 32-37(0.15-0.35) 30-35(0.08-0.27) 38-39(0.38-0.42) 110-115(0.67-0.75) 103-109(0.56-0.65) 71-72(0.06-0.08) 75-91(0.12-0.37) 110-131(0.67-1) 72-77(0.08-0.16) 67-7 1(0-0.06) 105-111(0.59-0.68) ? 92-100(0.39-0.51) 88-100(0.32-0.51) 78-103(0.17-0.56) 71-108(0.06-0.64) 73-92(0.07-0.39) 70-78(0.05-0.17) 80-129(0.20-0.96) 82-95(0.23-0.43) 83-85(0.25-0.28) 83-95(0.25-0.43) 78-88(0.17-0.32) 80-91(0.20-0.37) 97(0.46) 83-88(0.25-0.32) 83(0.25) 81-87(0.21-0.31) 85-96(0.28-0.45) 84-90(0.26-0.35) 84-93(0.26-0.40) 87-89(0.31-0.34) 73-89(0.09-0.34) 81-83(0.21-0.25) 81-96(0.21-0.45) 3 47-62 (0.18-0.38) 35-45(0.03-0.16) 61-62(0.35-0.37) 56-73(0.30-0.52) 50-70(0.22-0.48) 50-70(0.22-0.48) 67-84(0.44-0.66) 77-110(0.57-1.00) 68-82(0.45-0.63) 54-66(0.27-0.43) 54-57(0.27-0.31) 56(0.30) 56(0.30) 66-75(0.43-0.54) 60-70(0.35-0.48) 67-71(0.44-0.49) 45-55(0.16-0.29) 45-51(0.16-0.23) 53-65(0.26-0.42) 39-49(0.08-0.21) 45-58(0.16-0.32) 40-6 1(0.09-0.36) 46-76(0.17-0.56) 44-58(0.14-0.32) 50(0.22) 42-48(0.04-0.19) 46(0.17) 48-54(0.19-0.27) 39-49(0.08-0.21) 46-52(0.17-0.25) 36-55(0.04-0.29) 40-52(0.09-0.25) 50-76(0.22-0.56) 33-42(0-0.04) 47-56(0.18-0.30) 4 49-59 (0.27-0.41) 29-40(0-0.15) 56-62(0.37-0.45) 50-59(0.29-0.41) 45-59(0.22-0.41) 45-65(0.22-0.49) 67-87(0.52-0.79) 71-102(0.57-1.00) 68-82(0.53-0.73) 54-68(0.34-0.53) 54-62(0.34-0.45) 52(0.31) 47(0.25) 58-72(0.40-0.59) 62-67(0.45-0.52) 78-82(0.67-0.73) 42-49(0.18-0.27) 40-46(0.15-0.23) 41-56(0.16-0.37) 40-45(0.15-0.22) 44-50(0.20-0.29) 36-50(0.09-0.29) 48-85(0.26-0.77) 42-52(0.18-0.31) 43(0.19) 36-4.4(0.09-0.20) 44(0.20) 42-53(0.18-0.33) 38-46(0.12-0.23) 43-50(0.19-0.29) 34-50(0.07-0.29) 33-50(0.05-0.29) 45-81(0.22-0.71) 31-42(0.03-0.18) 45-53(0.22-0.33) 5 10 (0.42) 15-17(0.83-1) ? 11-13(0.5-0.67) 16(0.92) ? 10(0.42) ? ? 5-6(0-0.08) 2 6-8(0.08-0.25) ? 6-7(0.08-0.17) 7(0.17) 7-8(0.17-0.25) 9-13(0.33-0.67) 8-13(0.25-0.67) ? ? 6-8(0.08-0.25) 10-11(0.42-0.5) 10-11(0.42-0.5) ? ? 8(0.25) ? 6-8(0.08-0.25) ? ? 10-11(0.42-0.5) 13-14(0.67-0.75) 10-11(0.42 -0.5) 6-8(0.08-0.25) 6-8(0.08-0.25) Zoosyst. Evol. 96 (2) 2020, 345-395 Appendix 3 391 Discrete morphological characters utilised in the phylogenetic analysis. Characters numbers correspond with those in the character descriptions (see text). Terminal taxon 6-10 11-20 21-30 31-40 41-50 51-60 61-70 71-80 81-84 Proscyllium habereri 00000 0000000000 0711072000 0000000001 0000010001 0010000011 2100011100 200000001? 1000 Apristurus longicephalus 10000 1770000000 0700001000 1000101001 0001111121 1011000011 ?717717010 0001000000 1110 Asymbolus rubiginosus 00000 0000000000 1000000000 0000000000 0000000000 0000000010 0000000000 0000000000 0000 Atelomycterus fasciatus 21000 1771000000 0700010000 0001000010 1120000000 0010000010 0000011101 7001170000 0000 Aulohalaelurus labiosus 00000 1770000000 0700070027? 227271000010 012107272? 2222202270 0000011101 2000011701 0000 Cephaloscyllium isabella 00010 0000110711 0701011010 0101000110 0110010010 0010110102? 2272222222 222222222? 2000 C. sufflans 00010 0000110711 0701011010 0101000110 0110010010 0011110100 1001000101 7201017001? 1000 C. umbratile 00010 0000110711 0701011010 0101000110 0110010010 0010110102 2222222222 2222222200 2000 C. variegatum 00010 0000110711 0701011010 0101000110 0110010010 0011110102 2222222222 222222222? 2000 Cephalurus cephalus 10100 0100000000 02711111000 0110010000 0021001111 72010000111 2771721701? 200017001? 2110 Figaro boardmani 10000 0100000000 02700002100 0000100000 0000017100 0070000010 0101011000 0000170000 0000 Galeus antillensis 10000 0100000000 02700001100 0000101001 0000001101 0010000070 0101011001 2010171700 1100 Halaelurus natalensis 00000 0000000000 0700010000 0110000000 1020000000 0101000001 2100000000 0001001700 0000 Haploblepharus 21000 1771000000 0700010001 0010010000 0000001011 00100012701 ?700011011 72010000000 0000 edwardsii Holohalaelurus regani 10000 0100110770 1000001000 1010010000 1010011010 0110000010 1001100011 2101170000 1100 Parmaturus xaniurus 00000 0100000100 0711107100 0000100000 0020001111 1011000011 2700011000 0000000000 0110 Poroderma africanum 00101 00001012700 0700017011 0001000110 0100010010 0000110000 0001011101 ?110010001 11?1 P. pantherinum 00101 0000010000 0700017010 0001000110 0100010010 0000110000 0101011101 27110010001 1100 Schroederichthys 10110 0100000110 02700111000 0010000110 1110110010 0100001700 0001000100 0000000000 0000 saurisqualus Scyliorhinus boa 00010 0000101710 1100011010 0001000110 1110010010 0010010100 1001000101 1010170000 1000 S. cabofriensis 00010 00001012710 1100011010 0001000110 1110010010 0010010100 1001111101 ?110170100 1000 S. canicula 21010 0011101710 1200011010 0001000110 1110010010 0010010100 1001100101 72010170000 1000 S. capensis 00010 0000101710 1200011010 0001000110 1110010010 0010010100 0001000100 7010170100 1010 S. cervigoni 00010 0000101710 1100011010 0001000110 1110010010 0010010100 1001000101 0110170100 1000 S. comoroensis 00010 0000101710 1100011022? 2727702??? 2227227272??? 27222227270 %1001000101 ?11017?0000 2010 S. duhamelii 21010 0011101710 1200011010 0001000110 1110010010 0010010100 1001011101 211017001? 1000 S. garmani 20010 0000101710 20007107? 2222270272? 2122222222 2222222222 227227222222 2272727272?? + 2000 S. haeckelii 00010 0000101710 1100011010 0001000110 1110010010 0010010100 1001000100 0110170100 1000 S. hesperius 00010 0000101710 1100011010 0001000110 0010010 0010010100 0077700100 1272272272??? 2010 S. meadi 00010 0000101710 1100011010 0001000110 0010010 0010010100 1001000101 22222222?? 2010 S. retifer 00010 0000101710 1100011010 0001000110 0010010 0010010100 0001000100 1010170000 1001 S. stellaris 00010 0000101710 1100011010 0001000110 0010010 0010010100 1001000101 2110170000 1000 S. torazame 00010 0000101710 1200011010 0001000110 0010010 0010010100 1001011101 2010170000 1010 S. torrei 00010 0000101710 1200011010 0001000110 0010010 0010010100 1001000100 0110170000 1010 S. ugoi 00010 0000101210 1100011010 0001000110 0010010 0010010100 1001000100 0110170100 1000 Appendix 4 List of non-ambiguous synapomorphies of clades and terminal taxa based on the three most-parsimonious cladograms obtained. Synapomorphies followed by “!” appear only in some trees. Clade 1 Clade 3 Char. 2: 0.52—0.66 > 0.28-0.31. Char. 9:0> 1. Char. 29: 0 > 1. Char. 19: 0> 1. Char. 38: 0 > 1. Char. 53: 0 > 1. Chat 560-1. Char. 58: 0 > 1. Char. 59: 0 > 1. Char. 66: 1 > 0. Char. 64: 0 > 1. Char. 67: 1 > 0. Char. 73: 0> 1. Char.75: 01. Char. 81:0> 1. Clade 4 Clade 2 Char. 1: 0.54—0.65 > 0.73. Char. 8:0 > 1. Char. 3: 0.23-0.35 > 0.45-0.48. Char. 10: 0> 1. Char. 4: 0.23-0.37 > 0.49-0.53. Char. 72: 0 > 1. Char. 16: 0> 1. Char. 80: 0 > 1. Char. 20: 0 > 1. Char. 82: 0> 1. Char. 24: 0 > 1. zse.pensoft.net 392 Char. 32: 0 > 1. Clade 5 Char. 2: 0.28—0.31 > 0.13-0.17. Clade 6 Char. 54: 0 > 1. Clade 7 Char. 3: 0.23-0.35 > 0.21-0.22. Char. 4: 0.23-0.37 > 0.22. Char. 17: 0> 1. Char. 21: 1 > 0. Char. 41: 0> 1. Clade 8 Char. 70: 1 > 0. Char. 71-0: = 1, Clade 9 Char. 72: 0 > 1. Clade 10 Char. 1: 0.54—0.58 > 0.42-0.46! Chat. 3--0.423-0.05! Char. 78: 0 > 1. Clade 11 Char. 70: 1 > 0. Clade 12 Char. 83: 0> 1. Clade 13 Char 22-1 2, Clade 14 Char. 72: 1 > 0. Clade 15 Char. 66: 0 > 1. Char. 67: 0 > 1. Clade 16 Char. 6: 0 > 2. Char. 83: 1 > 0. Clade 17 Char 2-01 Char. 13: 0> 1. Char. 14: 0> 1. C. isabella Char. 2: 0.13—0.17 > 0.06—0.08. C. sufflans Char. 1: 0.73 > 0.81. C. umbratile Char. 2: 0.28—0.31 > 0.67—1.00. Char. 3: 0.45—0.48 > 0.57—1.00. Char. 4: 0.49-0.53 > 0.58—1.00. C. variegatum No autapomorphies. Appendix 5 Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus P_ africanum Chat 17: O> 1, Char. 30: 0 > 1. Char. 84:0 > 1. P. pantherinum Char. 62:0 > 1. S. boa No autapomorphies. S. cabofriensis No autapomorphies. S. canicula Char. 65:0 > 1. S. capensis Char. 1: 0.35—0.54 > 0.62-0.69. Char. 4: 0.21—0.22 > 0.26-0.77. Char. 65:0 > 1. Char. 78:0 > 1. S. cervigoni Char. 61: 1 > 0. S. comoroensis Char. 3: 0.28-0.31 > 0.47. Char. 79:0 > 1. S. duhamelii Char. 5: 0.42 > 0.25. Char. 72:0 > 1. S. garmani Char. 1: 0.35—0.54 > 0.77. S. haeckelii No autapomorphies. S. hesperius Char. 83:0 > 1. S. meadi Char. 1: 0.46—0.54 > 0.69-0.77. S. retifer Char. 84:0 > 1. S. stellaris Char. 1: 0.46—0.54 > 0.58—0.69! Char. 5: 0.42 > 0.67-0.75. S. torazame No autapomorphies. S. torrei Char. 1: 0.35—0.54 > 0.08-0.27. Char. 3: 0.17—0.22 > 0.00-0.12. Char. 4: 0.19-0.22 > 0.01-0.18. Char. 5: 0.42 > 0.08-0.25. Char: 70? 10. S. ugoi No autapomorphies. List of character transformation, based on the three most-parsimonious cladograms obtained. Char. 1 (L = 2,345) Clade 4: 0.54—0.65 > 0.73. Clade 7: 0.54—0.65 > 0.46-0.58. P. pantherinum: 0.54—0.65 > 0.15—0.69. zse.pensoft.net Clade 8: 0.46—0.58 > 0.46—0.54. C. umbratile: 0.73 > 0.73—-1.00. Clade 10: 0.46—0.58 > 0.42-0.46. C. isabella: 0.73 > 0.65—0.77. Zoosyst. Evol. 96 (2) 2020, 345-395 S. boa: 0.46—-0.54 > 0.42-0.54. S. cabofriensis: 0.42—0.46 > 0.35-0.42. S. hesperius: 0.46—0.54 > 0.42-0.54. S. retifer: 0.46—0.54 > 0.38-0.54. S. stellaris: 0.46—0.58 > 0.58-0.69. Clade 13: 0.46—-0.58 > 0.35-0.54. C. sufflans: 0.73 > 0.81. C. variegatum: 0.73 > 0.62-0.73. S. capensis: 0.35—0.54 > 0.15-0.35. S. cervigoni: 0.42—0.46 > 0.46—0.65. S. comoroensis: 0.46—0.58 > 0.46. S. garmani. 0.35—0.54 > 0.77. S. haeckelii: 0.42—0.46 > 0.31-0.46. S. meadi: 0.46—-0.58 > 0.69-0.77. S. torazame: 0.35—0.54 > 0.15-0.35. S. torrei: 0.35—0.54 > 0.08—0.27. S. ugoi:. 0.42—0.46 > 0.38—0.42. Clade 17: 0.35—0.54 > 0.35-0.46. S. canicula: 0.35—0.46 > 0.27-0.46. S. duhamelii: 0.35—0.46 > 0.27-0.35. Char. 2 (L = 1,862) Clade 1: 0.52—0.66 > 0.28—0.31. Clade 2: 0.28—0.31 > 0.17—0.31. P. africanum: 0.17—0.31 > 0.09-0.39. P. pantherinum: 0.17—0.31 > 0.05—0.17. Clade 5: 0.28—0.31 > 0.13—0.17. C. isabella. 0.12—0.16 > 0.23-0.43. C. umbratile: 0.28—0.31 > 0.67—1.00. S. boa: 0.28-0.31 > 0.23-0.44. S. cabofriensis: 0.28—0.31 > 0.25-0.28. S. hesperius: 0.28—0.31 > 0.28-0.45. S. retifer: 0.28—0.31 > 0.27-0.4. S. stellaris: 0.28—0.31 > 0.31-0.34. Clade 13: 0.28—0.31 > 0.25-0.31. C. sufflans: 0.13—0.17 > 0.13-0.38. C. variegatum: 0.13—0.17 > 0.08—0.17. S. canicula: 0.25 > 0.25-0.43. S. capensis: 0.25 > 0.17-0.32. S. cervigoni: 0.28—0.31 > 0.20-0.37. S. comoroensis: 0.28—0.31 > 0.46. S. duhamelii: 0.25—0.31 > 0.25-0.33. S. haeckelii: 0.28—0.31 > 0.22-0.31. S. meadi: 0.28—0.31 > 0.27-0.36. S. torazame: 0.25—0.31 > 0.09-0.34. S. torrei: 0.25 > 0.21-0.25. S. ugoi. 0.28—0.31 > 0.22-0.45. Char. 3 (L=1,205) Clade 1: 0.30-0.35 > 0.23—0.35. Clade 2: 0.23-0.35 > 0.23-0.29. Clade 4: 0.23-0.35 > 0.45—0.48. P. africanum: 0.23—0.29 > 0.16-0.23. P. pantherinum: 0.23-0.29 > 0.16—0.29. Clade 7: 0.23-0.35 > 0.21-0.22. Clade 8: 0.21-0.22 > 0.21. C. isabella: 0.455—0.48 > 0.22-0.48. C. umbratile: 0.45—0.48 > 0.57—1.00. S. boa: 0.21 > 0.08—0.21. S. cabofriensis: 0.21—0.22 > 0.16—0.32. S. hesperius: 0.21 > 0.08-0.21. S. retifer: 0.21 > 0.04—0.29. S. stellaris: 0.21-0.22 > 0.09-0.25. Clade 13: 0.21—0.22 > 0.17-0.22. C. sufflans: 0.45—0.48 > 0.44—0.66. C. variegatum: 0.45—0.48 > 0.45—0.64. S. capensis: 0.17—-0.22 > 0.17—0.56. S. cervigoni: 0.21—0.22 > 0.14—0.32. S. comoroensis: 0.21—-0.22 > 0.22. S. haeckelii: 0.21—0.22 > 0.19-0.27. S. meadi: 0.21—0.22 > 0.17-0.25. S. torazame: 0.17—-0.22 > 0.22-0.56. S. torrei. 0.17-0.22 > 0.22-0.56. S. ugoi: 0.22—0.25 > 0.18-0.30. Clade 16: 0.17—0.22 > 0.17-0.19. S. canicula: 0.17 — 0.19 > 0.09-0.36. S. duhamelii: 0.17—0.19 > 0.12-0.19. S. garmani. 0.17-0.19 > 0.17. Char. 4 (L=1,464) Clade 1: 0.29-0.37 > 0.23-0.37. Clade 2: 0.23-0.37 > 0.23-0.27. Clade 7: 0.23-0.37 > 0.22. Clade 4: 0.23-0.37 > 0.49-0.53. P. africanum: 0.23—0.27 < 0.18-0.27. P. pantherinum: 0.23-0.27 > 0.15—0.23. Clade 6: 0.49-0.53 > 0.52-0.53. C. isabella: 0.49-0.53 > 0.22-0.49. C. sufflans: 0.52—0.53 > 0.52-0.79. C. variegatum: 0.52—-0.53 > 0.53-0.73. S. cabofriensis: 0.22—0.25 > 0.21-0.29. S. cervigoni: 0.22 > 0.18—0.32. S. comoroensis: 0.19-0.22 > 0.19. S. haeckelii: 0.22 > 0.18—0.33. S. meadi: 0.19-0.22 > 0.19-0.29. S. torrei. 0.19-0.22 > 0.10-0.18. S. ugot. 0.22 > 0.22-0.33. Clade 12: 0.22-0.25 > 0.21. S. boa: 0.21 > 0.08-0.21. S. hesperius: 0.21 > 0.08-0.21. S. retifer: 0.21 > 0.04—0.29. S. stellaris: 0.22—0.25 > 0.09-0.25. Clade 14: 0.19-0.22 > 0.21-0.22. S. capensis: 0.20-0.22 > 0.26—0.77. S. torazame: 0.20-0.22 > 0.22-0.71. Clade 16: 0.20-0.22 > 0.21. S. canicula: 0.21 > 0.10-0.29. S. duhamelii: 0.20 > 0.10-0.21. Char. 5 (L=2,251) P. africanum: 0.42 > 0.33-0.67. P. pantherinum: 0.42 > 0.25—0.67. C. umbratile: 0.42 >? Clade 10: 0.42 > 0.25. Clade 12: 0.42 > 0.25—0.42. S. boa: 0.42 >? S. hesperius: 0.42 >? S. retifer: 0.42 > 0.42-0.50. S. stellaris: 0.42 >0.67-0.75. C. isabella: 0.42 >? 398 zse.pensoft.net 394 S. cabofriensis: 0.25 > 0.8—0.25. S. meadi: 0.25—0.42 >? S. comoroensis: 0.25—0.42 >? C. variegatum: 0.42 >? S. ugoi: 0.25 > 0.08—0.25. S. haeckelii: 0.25 > 0.08—0.25. S. cervigoni: 0.25 >? S. torrei. 0.25—0.42 > 0.08-0.25. Clade 14: 0.25-0.42 > 0.42. S. canicula: 0.42 > 0.42-0.50. S. capensis: 0.42 > 0.42—0.50. S. duhamelii 0.42 > 0.25. S. garmani. 0.42 >? S. torazame: 0.42 > 0.42-0.50. Char. 6 (L = 8) Clade 16: 0 > 2. Char. 7 (L = 3) Clade 17: 0> 1. Char. 8 (L = 3) Clade 2: 0> 1. Char. 9 (L= 2) Clade 3: 0 > 1. Char. 10 (L= 1) Clade 2: 0 > 1. Char. 11 (L= 3) No transformation in Scyliorhininae. Char. 12 (L=2) No transformation in Scyliorhininae. Char. 13 (L=1) Clade 17: 1 > 0. Char. 14 (L=3). Clade 17:0> 1. Char. 15 (L= 3) Clade 1:0>01. Clade 3: 01> 1. P. africanum: 01 > 1. P. pantherinum: 01 > 0. Char. 16 (L =2) Clade 4: 0 > 1. Char. 17 (L= 2) Clade'7: Or 1. P. africanum: 0 > 1. Char. 18 (L= 3) Clade 1:0>01. P. pantherinum: 01 > 1. Char. 19 (L=2) Clade 3:0> 1. Char. 20 (L= 1) Clade 4: 0> 1. Char. 21 (L=3) Clade 7: 1 >0. Char. 22 (L= 2) Clade 7: 01 > 1. Clade 13: 1 >2. Char. 23 (L= 1) No transformation in Scyliorhininae. Char. 24 (L= 2) Clade 4:0> 1. zse.pensoft.net Soares, K.D.A. & de Carvalho, M. R.: Phylogeny of the genus Scyliorhinus Char. 25 (L=2 No transformation in Scyliorhininae. Char. 26 (L=3 No transformation in Scyliorhininae. Char. 27 (L=1) Clade 1:0>01. Clade3: O1->>1. Char. 28 (L=2) No transformation in Scyliorhininae. Char. 29 (L=2 Clade 1:0> 1. Char. 30 (L=2 P. africanum: 0 > 1. Char. 31 (L=2 No transformation in Scyliorhininae. Char. 32 (L=3 Clade 4: 0> 1. Char. 33 (L= 4) No transformation in Scyliorhininae. Char. 34 (L=1 No transformation in Scyliorhininae. Char. 35 (L=2 No transformation tn Scyliorhininae. Char, 36.(ly=.3 No transformation in Scyliorhininae. Char. 37 (L=2 No transformation in Scyliorhininae. Char. 38 (L=2 Clade 1: 01 > 1. Char. 39 (L= 1) No transformation in Scyliorhininae. Char. 40 (L = 2) No transformation in Scyliorhininae. Char. 41 (L=5 Clade 7: 01 > 1. Char. 42 (L= 1 No transformation in Scyliorhininae. Char. 43 (L=7) Clade 2: 01 > 0. Clade 3: 01 > 1. Char. 44 (L = 3) No transformation in Scyliorhininae. Char. 45 (L=2) No transformation in Scyliorhininae. Char. 46 (L= 5) Clade 1: 01 > 1. Char. 47 (L=1) No transformation in Scyliorhininae. Char. 48 (L= 1) No transformation in Scyliorhininae. Char. 49 (L = 4) Clade Ik: 0'> 1. Char. 50 (L=2) No transformation in Scyliorhininae. Char. 51 (L= 1) No transformation in Scyliorhininae. Char. 52 (L=3 No transformation in Scyliorhininae. Zoosyst. Evol. 96 (2) 2020, 345-395 Char. 53 (L=2) Clade 3:0> 1. Char. 54 (L=4) Clade 6: 0> 1. Char. 55 (L=2) Clade 1:0> 01. Clade 2: 01 > 1. Clade 4: 01 > 1. Clade 7: 01 > 0. Char. 56 (L=1) Clade 1:0> 1. Char. 57 (L=2) No transformation in Scyliorhininae. Char. 58 (L=2) Clade 3:0 > 1. Char. 59 (L=4) Clade 1: 01 > 1. Char. 60 (L = 3) No transformation in Scyliorhininae. Char. 61 (L=4) Clade 3: 0 > 01. Clade 8: 01 > 0. Clade 9: 01 > 1. Char. 62 (L=3) P. pantherinum: 0 > 1. Char. 63 (L= 2) No transformation in Scyliorhininae. Char. 64 (L = 3) Clade 1: 0> 1. Char. 65 (L = 3) S. canicula: 0 > 1. S. capensis: 0 > 1. Char. 66 (L=4) Clade 2: 01 > 1. Clade 3: 01 > 0. Clade 15: 0> 1. Char. 67 (L=4) Clade 3: 01 > 0. Clade 15:0 > 1. Char. 68 (L= 1) No transformation in Scyliorhininae. Char. 69 (L= 4) No transformation in Scyliorhininae. Char. 70 (L=7 Clade 8: 1 > 0. Clade 11: 1 >0. S. torret: 1 >0. Char. 71 (L=1 Clade 8:0> 1. Char. 72 (L= 5) Clade 2:0 > 1. Clade 9:0 > 1. Clade 14: 1 > 0. S. duhamelii: 0 > 1. Char. 73 (L=3 Clade 1:0> 1. Char. 74 (L= 4) No transformation in Scyliorhininae. Char. 75 (L=5 Clade 3:0> 1. Char. 76 (L=1 No transformation in Scyliorhininae. Char. 77 (L=3 No transformation in Scyliorhininae. Char. 78 (L=2 Clade 10: 0> 1. S. capensis: 0 > 1. Char. 79 (L = 3) Clade 4: 0 > 01. S. comoroensis: 0 > 1. C. sufflans: 01 > 1. Char. 80 (L=2 Clade 2:0> 1. Char. 81 (L=4) Clade 1:0> 1. Char. 82 (L=4) Clade 2:0> 1. Char. 83 (L= 4) Clade 12:0> 1. Clade 16: 1 > 0. Char. 84 (L = 2) P. africanum: 0 > 1. S. retifer: 0 > 1. 325 zse.pensoft.net