Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-27T14:30:32.389Z Has data issue: false hasContentIssue false

Scaphanocephalus spp. (Trematoda: Opisthorchiidae) in intermediate and definitive hosts of the Yucatán Peninsula, Mexico, with a re-description of Scaphanocephalus expansus

Published online by Cambridge University Press:  14 December 2023

M. T. González-García
Affiliation:
Departamento de Zoología, Instituto de Biología, Universidad Nacional Autónoma de México, Avenida Universidad 3000, Ciudad Universitaria, C.P. 04510, Ciudad de México, México Posgrado en Ciencias Biológicas, Universidad Nacional Autónoma de México, Avenida Universidad 3000, Ciudad Universitaria, C. P. 04510, México
M. García-Varela
Affiliation:
Departamento de Zoología, Instituto de Biología, Universidad Nacional Autónoma de México, Avenida Universidad 3000, Ciudad Universitaria, C.P. 04510, Ciudad de México, México
A. López-Jiménez
Affiliation:
Departamento de Zoología, Instituto de Biología, Universidad Nacional Autónoma de México, Avenida Universidad 3000, Ciudad Universitaria, C.P. 04510, Ciudad de México, México
M. P. Ortega-Olivares
Affiliation:
Departamento de Zoología, Instituto de Biología, Universidad Nacional Autónoma de México, Avenida Universidad 3000, Ciudad Universitaria, C.P. 04510, Ciudad de México, México
G. Pérez-Ponce de León
Affiliation:
Departamento de Sistemas y Procesos Naturales, Escuela Nacional de Estudios Superiores Unidad Mérida, Km 4.5, Carretera Mérida-Tetiz, Ucú, Yucatán, C.P. 97357, México
L. Andrade-Gómez*
Affiliation:
Departamento de Sistemas y Procesos Naturales, Escuela Nacional de Estudios Superiores Unidad Mérida, Km 4.5, Carretera Mérida-Tetiz, Ucú, Yucatán, C.P. 97357, México
*
Corresponding author: L. Andrade-Gómez; Email: leoango23@gmail.com
Rights & Permissions [Opens in a new window]

Summary

Scaphanocephalus is a small trematode genus belonging to the family Opistorchiidae. The genus currently contains only three species associated with marine fish as intermediate hosts and fish-eating birds as definitive hosts. Here, specimens of Scaphanocephalus were collected from the Osprey, Pandion haliaetus, and the White mullet, Mugil curema in the Yucatán Peninsula, Mexico. We report for the first-time DNA sequences of adult specimens of Scaphanocephalus, particularly S. expansus, as well as a sequence of a different species sampled as metacercaria. Morphological comparisons of Scaphanocephalus expansus confirmed the identity of the adult specimens, with minor morphological variations; Scanning electron photomicrographs were included, and the species was re-described. Phylogenetic analysis based on 28S rDNA sequences showed that Scaphanocephalus is monophyletic within Opisthorchiidae and consists of three independent lineages. Sequences of adults are identical to those of S. expansus. Instead, the sequence of the metacercaria sampled from the mesentery of Mugil curema nested with specimens reported as Scaphanocephalus sp. from a labrid fish in the Mediterranean Sea, herein named it as Scaphanocephalus sp. 2.

Type
Research Paper
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press

Introduction

The opisthorchiid trematode genus Scaphanocephalus Jägerskiöld, Reference Jägerskiöld1904 comprises only three species which exhibit a complex life cycle involving marine snails as their first intermediate hosts, marine fishes as second intermediate hosts, and fish-eating birds from the order Accipitriformes as definitive hosts (Kohl et al. Reference Kohl, Calhoun, Elmer, Peachey, Leslie, Tkach and Johnson2019; Sokolov et al. Reference Sokolov, Frolov, Novokreshchennykh and Atopkin2021). They are characterized by a distinctive T-shaped form of the body resembling a cross-section of a mushroom (Pearson, Reference Pearson, Gibson, Jones and Bray2008; Katahira et al. Reference Katahira, Shimose and Kanaiwa2021).

Scaphanocephalus expansus (Creplin Reference Creplin1842) [type-species] was originally described as Monostomum expansum from the intestine of the Osprey, Pandion haliaetus (L.) (Pandionidae) in Europe. Subsequently, Jägerskiöld (Reference Jägerskiöld1904) reassigned those specimens and new material collected from the same host in the Red Sea, to the genus Scaphanocephalus. Adult specimens of this species seem to be highly host-specific since they have been reported in the Red Sea, Mediterranean Sea, Gulf of California, Gulf of Mexico, and Canary Archipelago, only parasitizing the Osprey, P. haliaetus (Creplin, Reference Creplin1842; Jägerskiöld, Reference Jägerskiöld1904; Hoffman et al. Reference Hoffman, Wu and Kingscote1953; Dubois, Reference Dubois1960; Schmidt & Huber, Reference Schmidt and Huber1985; Kinsella et al. Reference Kinsella, Cole, Forrester and Roderick1996; Foronda et al. Reference Foronda, Santana-Morales, Feliu and Valladares2009). In contrast, larval stages of S. expansus have been observed on fins and scales of 17 species of reef fish allocated in 12 families in the Western Pacific, Caribbean Sea, and Gulf of Mexico (Yamaguti, Reference Yamaguti1942; Kohl et al. Reference Kohl, Calhoun, Elmer, Peachey, Leslie, Tkach and Johnson2019; Montoya-Mendoza et al. Reference Montoya-Mendoza, Morales-Sánchez, Arenas-Fuentes and González-Solís2021). The metacercariae of Scaphanocephalus causes the black-spot syndrome (Elmer et al. Reference Elmer, Kohl, Johnson and Peachey2019; Kohl et al. Reference Kohl, Calhoun, Elmer, Peachey, Leslie, Tkach and Johnson2019; Cohen-Sánchez et al. Reference Cohen-Sánchez, Valencia, Box, Solomando, Tejada, Pinya, Catanese and Sureda2023).

The other two congeneric species are S. australis Johnston, Reference Johnston1917 reported from the White-bellied Sea Eagle, Haliaeetus leucogaster (Gmelin) (Accipitridae) in Australia (WoRMS, 2023), and S. adamsi Tubangui, Reference Tubangui1933 which was described solely from the metacercarial stage obtained from the fins and under the scales of the Splitlevel hogfish, Bodianus mesothorax (Bloch & Schneider) (Labridae) in the Philippines (Yamaguti, Reference Yamaguti1958). Even though Yamaguti (Reference Yamaguti1942) suggested that S. adamsi represented the same species as S. expansus, it has been also reported from marine fishes of the families Mullidae and Scaridae in Japan (Shimose et al. Reference Shimose, Katahira and Kanaiwa2020; Katahira et al. Reference Katahira, Shimose and Kanaiwa2021). Additionally, the metacercariae of Scaphanocephalus sp. (unidentified) has been recorded as parasite of marine fishes belonging to six families, i.e., Tetraodontidae, Serranidae, Hemiramphidae, Pomacentidae, Siganidae, and Labridae, in localities of the Northeast Pacific Ocean, Gulf of Mexico, Gulf of California, Arabian Gulf and Mediterranean Sea (Pérez-Urbiola et al. Reference Pérez-Urbiola, Flores Herrera, Nieth Kmoth and Inohuye- Rivera1994; Bullard & Overstreet, Reference Bullard, Overstreet, Eiras, Segner, Wahil and Kapoor2008; Al-Salem et al. Reference Al-Salem, Baghdadi, Mahmoud, Ibrahim and Bayoumy2021; Cohen-Sánchez et al. Reference Cohen-Sánchez, Valencia, Box, Solomando, Tejada, Pinya, Catanese and Sureda2023).

Until now, 28S rDNA sequences have been obtained only for the metacercarial stages of Scaphanocephalus spp. We collected metacercariae of Scaphanocephalus from the body cavity of one out of 67 specimens of White mullet, Mugil curema (Mugilidae) sampled in three coastal lagoons of Yucatán. Additionally, adult specimens of Scaphanocephalus spp. were collected from the intestine of one specimen of the Osprey, Pandion haliaetus from Champotón, Campeche, also in the Yucatán Peninsula. Adults were morphologically identified as Scaphanocephalus expansus. Thus, the main objective of this study was two-fold: to corroborate molecularly the identification of the adults, and to establish a molecular link between the metacercariae and adults from the same geographical region.

Materials and methods

Specimen collection

Adult specimens of Scaphanocephalus sp. were recovered from the intestine of one specimen of Pandion haliaetus collected in Champotón, Campeche. Additionally, we collected 67 specimens of the White mullet, Mugil curema (Valenciennes) in four coastal lagoons of northern Yucatán, namely La Carbonera, Dzilam de Bravo, Ría Lagartos, and Celestún (Fig. 1). Only one fish was found to be infected in the mesentery with an individual of Scaphanocephalus sp. from Celestún. All Scaphanocephalus specimens were fixed in nearly boiling distilled water and preserved in 100% ethanol for DNA analyses. Some adult specimens were fixed in hot 4% formalin solution for morphological studies and Scanning Electron Microscopy (SEM) studies.

Figure 1. Sampling collection in Yucatán Peninsula, Mexico. Pandion haliaetus, 1. Champotón, Campeche (19º21’40”N, 90º43’5”W). Mugil curema, 2. La Carbonera (21º8’1”N, 90º7’55”W) 3. Celestún (20º50’53N, 90º24’22”W), 4. Dzilam (21º23’40”N, 88º53’13”W) 5. Ría Lagartos (21º35’45”N, 88º8’48”W), Yucatán. Localities where Scaphanocephalus spp. were recovered are in red. In blue, localities where the parasite was not found.

Morphological analyses

Specimens of Scaphanocephalus were stained with Mayer’s paracarmine, dehydrated in a graded ethanol series, cleared with methyl salicylate, and mounted on permanent slides with Canada balsam. Mounted specimens were examined and measured using a Leica DM 1000 LED compound microscope (Leica Microsystems CMS GmbH, Wetzlar, Germany). Measures are reported in micrometers (μm). Illustrations of internal morphological features were generated using a drawing tube attached to a Leica MC120HD. Drawings were processed in Adobe Illustrator 27.9 (Adobe, Inc). Voucher specimens, including hologenophores (Pleijel et al. Reference Pleijel, Jondelius, Norlinder, Nygren, Oxelman, Schander, Sundberg and Thollesson2008) (Fig. 2), were deposited in the Colección Nacional de Helmintos (CNHE), Instituto de Biología, Universidad Nacional Autónoma de México (UNAM), Mexico City. Additionally, three adult specimens preserved in 4% formalin, dehydrated in a graded ethanol series, critical point dried, sputter-coated with gold and examined with a Hitachi Stereoscan Model S-2469N scanning electron microscope at 15 kV at the LaNABIO, Instituto de Biología, Universidad Nacional Autónoma de México (UNAM).

Figure 2. A–C) Hologenophore of Scaphanocephalus expansus from Pandion haliaetus of Champotón, Campeche. D) Hologenophore of Scaphanocephalus sp. from Mugil curema in Celestún, Yucatán. CNHE and GenBank accession numbers are indicated. Scale bars= (A–C) 1 mm, D 300 μm.

DNA isolation, amplification, and sequencing

Four hologenophores, including three adults and one metacercaria, were placed individually in tubes, digested, and DNA extracted following the protocol by González-García et al. (Reference González-García, Ortega-Olivares, Andrade-Gómez and García-Varela2020). Domains D1-D3 from LSU were amplified using forward primer 391, 5′-AGCGGAGGAAAAGAAACTAA-3′ (Nadler et al. Reference Nadler, D’Amelio, Fagerholm, Berland and Paggi2000), and reverse primer 536, 5′-CAGCTATCCTGAGGGAAAC-3′ (Garcia-Varela & Nadler, Reference García-Varela and Nadler2005). The amplification and sequencing protocols followed those used in Andrade-Gómez et al. (Reference Andrade-Gómez and García-Varela2021). Contigs were assembled, and base-calling differences resolved using Codoncode Aligner version 9.0.1 (Codoncode Corporation, Dedham, Massachusetts) and deposited in GenBank.

Alignments and phylogenetic analyses

The newly generated sequences of Scaphanocephalus were aligned with other Opisthorchiidae species downloaded from GenBank, along with 14 members of Heterophyidae used as outgroups. Alignment was performed using the software SeaView version 4 (Gouy et al. Reference Gouy, Guindon and Gascuel2010) and adjusted with the Mesquite program (Maddison & Maddison, Reference Maddison and Maddison2011). A nucleotide substitution model was selected using jModelTest v2.1.7 (Posada, Reference Posada2008) and applying the Akaike information criterion, with the best nucleotide substitution model for data being GTR + G + I.

Phylogenetic analyses were conducted using Bayesian inference (BI) and maximum likelihood (ML) methods through the online interface Cyberinfrastructure for Phylogenetic Research (CIPRES) Science Gateway v3.3 (Miller et al. Reference Miller, Pfeiffer and Schwartz2010). The BI analysis was inferred with MrBayes v.3.2.7 (Ronquist et al. Reference Ronquist, Teslenko, Van der Mark, Ayres, Darling, Höhna, Larget, Liu, Suchard and Huelsenbeck2012), with two simultaneous runs of the Markov Chain Monte Carlo (MCMC) for 10 million generations, sampled every 1000 generations, using a heating parameter value of 0.2 and the first 25% of the sampled trees were discarded. The ML analysis was carried out with RAxML v.7.0.4 (Silvestro & Michalak, Reference Silvestro and Michalak2011), and 1000 bootstrap replicates were run to assess nodal support. Phylogenetic trees were drawn in FigTree v.1.3.1 (Rambaut, Reference Rambaut2012). Genetic divergence among taxa was estimated using uncorrected ‘p’ distances with MEGA version 6 (Tamura et al. Reference Tamura, Stecher, Peterson, Filipski and Kumar2013).

RESULTS

Class Trematoda Rudolphi, 1808

Order Plagiochiida La Rue, 1957

Family Opisthorchiidae, Looss 1899

Subfamily Cryptocotylinae Lühe, 1909

Genus Scaphanocephalus Jägerskiöld, Reference Jägerskiöld 1904

Scaphanocephalus expansus (Creplin, Reference Creplin 1842 )

Records: Adult specimens, 1. Creplin (Reference Creplin1842); 2. Jägerskiöld (Reference Jägerskiöld1904); 3. Gohar (Reference Gohar1935); 4. Smogorzhevskaya (Reference Smogorzhevskaya1956); 5. Dubois (Reference Dubois1960); 6. Lee (Reference Lee1966); 7. Schmidt & Huber (Reference Schmidt and Huber1985); 8. Kinsella et al. (Reference Kinsella, Cole, Forrester and Roderick1996); 9. Hoffman (1999); 10. Foronda et al. (Reference Foronda, Santana-Morales, Feliu and Valladares2009); 11. This study.

Metacercariae specimens, 12. Yamaguti (Reference Yamaguti1942); 13. Hutton (Reference Hutton1964); 14. Skinner (Reference Skinner1978); 15. Overstreet & Hawkins (Reference Overstreet and Hawkins2017); 16. Kohl et al. (Reference Kohl, Calhoun, Elmer, Peachey, Leslie, Tkach and Johnson2019); 17. Elmer et al. (Reference Elmer, Kohl, Johnson and Peachey2019); 18. Montoya-Mendoza et al. (Reference Montoya-Mendoza, Morales-Sánchez, Arenas-Fuentes and González-Solís2021).

Type and only definitive host: (1–11) Pandion haliaetus (L.) (Pandionidae)

Intermediate hosts: Acanthuridae: (16, 17) Acanthurus tractus Poey; (16), A. chirurgus Bloch; (16) A. coeruleus Bloch & Schneider; Carangidae: (16) Caranx ruber Bloch; Gerreidae: (14) Eucinostomus argenteus Baird & Girard; Labridae: (18) Halichoeres radiatus (L.); Lutjanidae: (15) Lutjanus griseus (L.); Monacanthidae: (16) Cantherhines pullus Ranzani; Mullidae: (16) Mulloidichthys martinicus Cuvier; (12) Parupeneus multifasciatus Quoy & Gaimard; Pristidae: (13) Pristis pectinata Latham; Scaridae: (16) Sparisoma aurofrenatum Valenciennes, (16) S. chrysopterum Bloch & Schneider; Sciaenidae: (14) Micropogonias undulatus (L.); Serranidae: (16) Cephalopholis cruentata Lacepède; (16) Paranthias furcifer Valenciennes; Sparidae: (14) Archosargus rhomboidalis (L.).

Type locality: (1, 4) ‘Europe’ (unspecified locality)

Other localities: Adult specimen records: (2, 3) Red Sea; (5) Mediterranean Sea; (6) Malaysia; (7) Gulf of California; (8, 11) Gulf of Mexico; (9) USA; and (10) Canary Archipelago. Metacercariae specimen records: (12) Japan; (13, 14, 15, 18) Gulf of Mexico; and (16, 17) Caribbean Sea.

Site of infection: Adults in intestine, Metacercaria on fins and underneath scales

Intensity: Adults specimens: (7) 35; (8) 361; (9) 30; (10) 43; (11) 123.

Specimens deposited: 16 vouchers (CNHE12838); 3 hologenophores (CNHE12839–12841)

Representative DNA sequences: OR794151–OR794153 (28S)

Redescription based on 19 specimens, including three hologenophores plus three individuals analyzed for SEM (Fig. 2A–C & 3). Measurements are provided in Table 1.

Table 1. Comparative morphometric data for Scaphanocephalus expansus

Body elongated and flared anteriorly (Fig. 3A, 3B). Tegument completely armed with two types of spines, pectinate on large wing-like anterior expansions, and arrow-like on posterior body end (Fig. 3D, 3E). Anterior of the body exhibiting prominent wing-like anterior expansions; anterior border crenulated, striated (3B, 3C). Oral sucker subterminal, bearing dome-like papillae (Fig. 3C). Prepharynx inconspicuous; pharynx oval, oesophagus relatively large. Caeca narrow extending to posterior body end reaching the inter-testicular level. Ventrogenital complex located in first third of body. Ventral sucker greatly reduced embedded in body parenchyma, opening into the ventrogenital complex.

Figure 3. A) Scaphanocephalus expansus from Pandion haliaetus. Schematic whole worm, ventral view. Scanning electron photomicrographs, B) whole worm; C) Oral sucker; D) Tegumental spines at level of wing-like expansions; E) Spines at the level of the third part of the body. Scale bars = (A, B) 1 mm; (C), 50 μm; (D, E) 5 μm.

Testes two, deeply lobated, in tandem, in posterior third of body. Posterior testis larger. Seminal vesicle tubular, winding, dorsally to uterus. Ovary deeply lobated, post-ecuatorial, pretesticular. Mehli´s gland sinistral, at ovary level. Seminal receptacle postovarian, pretesticular. Laurer’s canal between ovary and Mehli’s gland. Vitelline glands in the lateral fields, from caecal bifurcation to posterior body end; confluent in post-testicular area. Vitelloducts at ovary level. Uterus pretesticular, deeply coiled, between ventrogenital complex region and ovary, opening to genital pore in ventrogenital complex. Eggs small, oval to round shaped. Excretory vesicle Y-shaped, extending sinuously to reach anterior of body.

Remarks

Our specimens were identified as Scaphanocephalus expansus by having features consistent with the diagnosis of the original description and recent re-descriptions (Creplin, Reference Creplin1842; Jägerskiöld, Reference Jägerskiöld1904; Dubois, Reference Dubois1960; Pearson, Reference Pearson, Gibson, Jones and Bray2008; Foronda et al. Reference Foronda, Santana-Morales, Feliu and Valladares2009). For instance, our specimens possess an anterior body with wing-like expansions, testes deeply lobated located in tandem in the posterior third of body, ovary lobated, pretesticular, and tegument covered with spines. In addition, metrical data of our specimens are like those reported in previous studies (Table 1). We noticed, however, some slight morphological differences, such as the variation in body length; the newly sampled specimens are, on average, smaller than those reported in previous studies (2256−5063 vs 2812−5000). The ovary and testes are also smaller (Ovary: 105−263 vs 160−380; anterior testis: 221−465 vs 400−690; posterior testis: 276−495 vs 548−840) (Table 1). In addition, SEM images allowed to show that S. expansus presented two types of spines, i.e., pectinate and arrow-like (Figs. 3D, 3E). Dome-like papillae surrounding the aperture of the oral sucker are described for the first time (Fig. 3C).

We re-examined the specimens deposited of the metacercariae of Scaphanocephalus expansus by Montoya-Mendoza et al. (Reference Montoya-Mendoza, Morales-Sánchez, Arenas-Fuentes and González-Solís2021) in the Colección Nacional de Helmintos, México City (CNHE No. 11508) reported encysted on the fins of the labrid Halichoeres radiatus in the Veracruz coral reef, Mexico; and noticed that they are morphologically very similar to our adult specimens, both with multilobulated testes. However, DNA sequences are required of these larvae stages to corroborate the status and complete the life cycle of the species.

Phylogenetic analyses

The LSU dataset comprised 36 sequences with 1,247 characters. The alignment was trimmed to the shortest and included the four newly generated sequences of Scaphanocephalus, and six sequences previously identified either as Scaphanocephalus sp., or S. expansus. In addition, the alignment included 11 sequences of members of Opisthorchiidae, plus 14 sequences of Heterophyidae. Phylogenetic analyses conducted through BI and ML methods recovered similar topologies (Fig. 4). Analyses showed Opisthorchiidae as monophyletic with strong nodal support (1/100); Cryptocotylinae was also recovered as monophyletic, including Cryptocotyle spp. and Scaphanocephalus spp. The genus Scaphanocephalus was divided in three highly supported clades (1/100). The first clade herein referred as Scaphanocephalus sp. 1 (MN160569 and MT461356) included two sequences from a siganid and acanthurid from the Arabian Gulf and the Caribbean Sea, respectively. This clade was resolved as the sister group of the other two clades of Scaphanocephalus. The second clade was formed by three of the newly generated sequences of S. expansus (from the definitive host, P. haliaetus), plus one sequence previously identified as S. expansus (MK680936) sampled from a scarid fish, and one sequence identified as Scaphanocephalus sp. (MN160570) from an acanthurid fish, both from the Caribbean Sea. Finally, the third clade was yielded as the sister group of S. expansus and is formed by four sequences, the specimen from Mugil curema (Fig. 2D; CNHE 12842), and those previously identified as Scaphanocephalus sp. from a labrid in the Mediterranean Sea, herein referred as Scaphanocephalus sp. 2.

Figure 4. Consensus Bayesian inference (BI) tree and Maximum likelihood (ML) tree inferred from the large subunit from nuclear ribosomal DNA. Numbers on internal nodes show posterior probabilities (BI) and ML bootstrap clade frequencies. Sequences generated in this study in bold. Grey: specimens from Arabian Gulf. Blue: specimens from the Caribbean Sea. Green: specimens from Yucatán Peninsula, Mexico. Red: specimens from the Mediterranean Sea. Scale bar shows the number substitutions per site.

The genetic divergence of 28S estimated among the three clades of Scaphanocephalus ranged from 1.06%−1.73%., whereas the divergence between Scaphanocephalus sp. 1 (MN160569 + MT461356) and S. expansus + Scaphanocephalus sp. 2 was 1.39−1.73%. The species pair S. expansus and Scaphanocephalus sp. 2 varied 1.06−1.12%. The genetic divergence within each clade of Scaphanocephalus was null.

Discussion

The genus Scaphanocephalus was originally considered a member of Heterophyidae (Pearson, Reference Pearson, Gibson, Jones and Bray2008; Dennis et al. Reference Dennis, Izquierdo, Conan, Johnson, Giardi, Frye and Freeman2019). However, Sokolov et al. (Reference Sokolov, Frolov, Novokreshchennykh and Atopkin2021) recently transferred Scaphanocephalus to the subfamily Cryptocotylinae within Opisthorchiidae using molecular and morphological data. Our LSU phylogenetic analyses unequivocally corroborated that Scaphanocephalus is monophyletic and belongs to Opisthorchiidae.

Dennis et al. (Reference Dennis, Izquierdo, Conan, Johnson, Giardi, Frye and Freeman2019) performed the first molecular phylogenetic analysis including two lineages of Scaphanocephalus, recognizing to separate species as Scaphanocephalus sp. 1 and Scaphanocephalus sp. 2. The latter was nested with S. expansus reported by Kohl et al. (Reference Kohl, Calhoun, Elmer, Peachey, Leslie, Tkach and Johnson2019). Al-Salem et al. (Reference Al-Salem, Baghdadi, Mahmoud, Ibrahim and Bayoumy2021) and Cohen-Sánchez et al. (Reference Cohen-Sánchez, Valencia, Box, Solomando, Tejada, Pinya, Catanese and Sureda2023) obtained sequences of Scaphanocephalus although without a molecular phylogenetic analysis. Unexpectedly, our phylogenetic analysis showed that the sequence of Scaphanocephalus from the Arabian Gulf (MT461356) reported by Al-Salem et al. (Reference Al-Salem, Baghdadi, Mahmoud, Ibrahim and Bayoumy2021) nested with Scaphanocephalus sp. 1 from the Caribbean Sea reported by Dennis et al. (Reference Dennis, Izquierdo, Conan, Johnson, Giardi, Frye and Freeman2019).

Moreover, the sequence reported as Scaphanocephalus sp. 2 (MN160570) by Dennis et al. (Reference Dennis, Izquierdo, Conan, Johnson, Giardi, Frye and Freeman2019) from the acanthurid A. chirurgus in the Caribbean Sea nested as sister taxon of S. expansus from the scarid S. chrysopterum in the same geographical region as reported by Kohl et al. (Reference Kohl, Calhoun, Elmer, Peachey, Leslie, Tkach and Johnson2019). Our sequences from adult specimens of S. expansus from the definitive host were also nested within this clade. Based on this evidence, we consider that Scaphanocephalus sp. 2 of Dennis et al. (Reference Dennis, Izquierdo, Conan, Johnson, Giardi, Frye and Freeman2019) as S. expansus. Furthermore, Cohen-Sánchez et al. (Reference Cohen-Sánchez, Valencia, Box, Solomando, Tejada, Pinya, Catanese and Sureda2023) recently reported sequences of Scaphanocephalus sp. from the labrid X. novacula, in the Mediterranean Sea. These authors made no further taxonomic consideration because their study was focused on determining the effect of the black spot disease produced by Scaphanocephalus on the oxidative stress of the host. We included these sequences in our analyses and found that they nested with the newly sequenced individual from M. curema from Yucatán, with null genetic variation, suggesting this lineage is widely distributed across the Atlantic Ocean parasitizing marine fishes. This clade is herein considered as Scaphanocephalus sp. 2 (Fig. 4). In addition, M. curema represent a new species of intermediate hosts for Scaphanocephalus. Interestingly, only one specimen of Scaphanocephalus was recovered from the mesentery of 1 out of 67 individuals of M. curema studied for parasites in four localities of northern Yucatán. The infection site (mesentery) and the low prevalence of infection suggest that the presence of the parasite in M. curema may represent an accidental infection. The metacercariae of Scaphanocephalus have been reported as ectoparasites encysted on the fins and under the scales of their hosts (Dennis et al. Reference Dennis, Izquierdo, Conan, Johnson, Giardi, Frye and Freeman2019; Elmer et al. Reference Elmer, Kohl, Johnson and Peachey2019; Kohl et al. Reference Kohl, Calhoun, Elmer, Peachey, Leslie, Tkach and Johnson2019; Shimose et al. Reference Shimose, Katahira and Kanaiwa2020; Katahira et al. Reference Katahira, Shimose and Kanaiwa2021; Montoya-Mendoza et al. Reference Montoya-Mendoza, Morales-Sánchez, Arenas-Fuentes and González-Solís2021; Cohen-Sánchez et al. Reference Cohen-Sánchez, Valencia, Box, Solomando, Tejada, Pinya, Catanese and Sureda2023).

The LSU genetic divergence observed among the three clades of Scaphanocephalus was relatively low (1.06 to 1.73%). However, these genetic divergence values are like those reported for other members of the Cryptocotylinae. For instance, Tatonova & Besprozvannykh (Reference Tatonova and Besprozvannykh2019) reported a genetic divergence of 2% between Cryptocotyle lingua (Creplin, 1825) and C. lata Tatonova and Besprozvannykh, Reference Tatonova and Besprozvannykh2019. Interestingly, in our study we found no genetic divergence within each lineage of Scaphanocephalus. These data suggest that definitive and intermediate hosts are playing an important role for the distribution of each lineage of Scaphanocephalus.

Up to the present, only three species have been included in the genus Scaphanocephalus, namely S. expansus, S. australis and S. adamsi. The first two were described from their definitive hosts, S. expansus from P. haliaetus, and S. australis from H. leucogaster. In contrast, the description of S. adamsi was based only on the metacercarial stage from the labrid B. mesothorax (Tubangui Reference Tubangui1933). In the present study, two unidentified Scaphanocephalus (sp.1 and sp. 2) represent larval stages. Whether or not they represent the two species of Scaphanocephalus not yet sequenced needs to be tested once molecular studies are conducted on these species. Interestingly, Scaphanocephalus sp. 2 contain specimens sampled from M. curema in Mexico and, more importantly, from the labrid X. novacula from Spain (Cohen-Sánchez et al. Reference Cohen-Sánchez, Valencia, Box, Solomando, Tejada, Pinya, Catanese and Sureda2023). The latter represents the same host-type (Labridae) of S. adamsi, although the species was described from B. mesothorax in the Philippines. Two pieces of information are missing to test the hypothesis that Scaphanocephalus sp. 2 represents in fact S. adamsi, sampling and sequencing metacercariae from the type locality and, ideally, sampling adult forms from Accipitriformes. Furthermore, the only published record of S. australis was from adults collected from the accipitrid H. leucogaster in Australia (Johnston Reference Johnston1917), and their larval forms have not been reported from marine fishes in the area. The White bellied sea eagle, H. leucogaster has an extensive geographic distribution, extending from the Indian west coast, China, to all over South-East Asia, including Indonesia, Papua New Guinea, and Australia (Shephard et al. Reference Shephard, Catterall and Hughes2005). It is imperative to collect new material of S. australis to determine whether or not Scaphanocephalus sp. 1 or 2 belong in that species, or if they represent a separate species whose adults have not been found in Accipitriformes.

The osprey, Pandion haliaetus has a cosmopolitan distribution which probably has a key role in the distribution of S. expansus across the world. In the Caribbean two subspecies have been reported, the North American osprey, P. haliaetus carolinensis, which is migratory, and the non-migratory osprey, P. haliaetus ridgwayi (Wiley et al. Reference Wiley, Poole and Clum2014) This could explain the presence of the two lineages of Scaphanocephalus in the Caribbean, S. expansus and Scaphanocephalus sp. 1. Still, gathering new samples of adults and metacercariae from different host species is necessary to test the hypothesis on the potential link between larval forms and adults described from accipitriformes, and to understand the evolution and biogeographic history, as well as the interrelationships and host-parasite interactions of this enigmatic group of digeneans.

Autor contribution

MTGG and LAG conceptualization, genetic analysis, writing – original draft, review, editing. ALJ and MPOO sampling, investigation, genetic data curation, morphology analysis. GPPL and MGV funding acquisition, review, and editing.

Acknowledgements

LAG thanks to the Dirección General de Asuntos de Personal Académico (DGAPA-UNAM) Mexico for the Postdoctoral Fellowship granted. MTGG thank the support of the Programa de Posgrado en Ciencias Biológicas, UNAM, and CONAHCYT (CVU 956064) for granting a scholarship to complete his Master program. We also thank Berenit Mendoza for her help with the use of the SEM unit and Laura Márquez and Nelly López Ortiz from LaNabio for their help during the sequencing of the DNA fragments. We thank Luis García Prieto for the loan of material of CNHE and for providing bibliography. We thank Norberto Colín for the use of facilities. We sincerely thank Maribel Badillo Alemán and Alfredo Gallardo Torres, Laboratorio de Biología de la Conservación, Facultad de Ciencias UNAM for lab facilities and the identification of the fish host.

Financial support

This research was supported by grants from the Programa de Apoyo a Proyectos de 375 Investigación e Innovación Tecnológica (PAPIIT-UNAM) IN212621 to GPPL and IN201122 to MGV, and the Consejo Nacional de Humanidades, Ciencias y Tecnologías (CONAHCYT A1-S-21694) to GPPL.

Conflict of interest

None

Ethical standard

Specimens were collected under the Cartilla Nacional de Colector Científico (FAUT 0202) issued by the Secretaría del Medio Ambiente y Recursos Naturales (SEMARNAT), to MGV. Specimens were collected under the sampling permit granted to the Laboratorio de Biología de la Conservación (BioCon) by the Comisión Nacional de Acuacultura y Pesca (CONAPESCA), No. PPF/DGOPA-001/20. Fish were humanely euthanized following the protocols described by the 2020 edition of the AVMA Guidelines for the Euthanasia of Animals.

References

Al-Salem, AAM, Baghdadi, HB, Mahmoud, M, Ibrahim, MM and Bayoumy, EM (2021) Morphomolecular and pathological study of Scaphanocephalus sp. in new host Siganus argenteus in the Arabian Gulf. Diseases of Aquatic Organisms 144, 221230.Google Scholar
Andrade-Gómez, L and García-Varela, M (2021) Unexpected morphological and molecular diversity of trematode (Haploporidae: Forticulcitinae) parasites of mullets from the ocean Pacific coasts in Middle America. Parasitology Research 120(1), 5572.Google Scholar
Bullard, SA and Overstreet, RM (2008) Digeneans as enemies of fishes. In: Eiras, J., Segner, H., Wahil, T., Kapoor, B.G. (eds) Fish Diseases. Science Publishers, 817976Google Scholar
Cohen-Sánchez, A, Valencia, JM, Box, A, Solomando, A, Tejada, S, Pinya, S, Catanese, G and Sureda, A (2023) Black spot disease related to a trematode ectoparasite causes oxidative stress in Xyrichtys novacula. Journal of Experimental Marine Biology and Ecology 560, 151854. doi:10.1016/j.jembe.2022.151854.Google Scholar
Creplin, JCH (1842) Beschreibung der Psorospermien des Kaulbarsches nebst einigen Bemerkungen uber die der Plotze und andere., Vol. 8, Archiv für Naturkunde.Google Scholar
Dennis, MM, Izquierdo, A, Conan, A, Johnson, K, Giardi, S, Frye, P and Freeman, MA (2019) Scaphanocephalus-associated dermatitis as the basis for black spot disease in Acanthuridae of St. Kitts, West Indies. Diseases of Aquatic Organisms 137(1), 5363.Google Scholar
Dubois, G (1960) Redecouverte de Scaphanocephalus expansus (Creplin 1842) (Trematoda: Heterophyidae). Annales De Parasitologie Humaine Et Comparée 35, 426427.Google Scholar
Elmer, F, Kohl, ZF, Johnson, PTJ and Peachey, RBJ (2019) Black spot syndrome in reef fishes: using archival imagery and field surveys to characterize spatial and temporal distribution in the Caribbean. Coral Reefs 38(6), 13031315.Google Scholar
Foronda, P, Santana-Morales, MA, Feliu, C and Valladares, B (2009) New record of Scaphanocephalus expansus from the Canary Islands (Spain). Helminthologia 46(3), 198200.Google Scholar
García-Varela, M and Nadler, SA (2005) Phylogenetic relationships of Palaeacanthocephala (Acanthocephala) inferred from SSU and LSU rRNA gene sequences. Journal of Parasitology 91(6), 14011409.Google Scholar
Gohar, N (1935) Liste des trématodes parasites et de leurs hôtes vertébrés signalés dans la vallée du Nil. Annales De Parasitologie Humaine Et Comparée 13(1), 8090.Google Scholar
González-García, MT, Ortega-Olivares, MP, Andrade-Gómez, L and García-Varela, M (2020) Morphological and molecular evidence reveals a new species of Lyperosomum Looss, 1899 (Digenea: Dicrocoeliidae) from Melanerpes aurifrons (Wagler, 1829) from northern Mexico. Journal of Helminthology 94, E156. doi:10.1017/S0022149X20000425Google Scholar
Gouy, M, Guindon, S and Gascuel, O (2010) SeaView Version 4: A multiplatform graphical user interface for sequence alignment and phylogenetic tree building. Molecular Biology and Evolution 27(2), 221224.Google Scholar
Hoffman, GL, Wu, LY and Kingscote, AA (1953) Scaphanocephalus expansus (Crepl.), a Trematode of the Osprey, in North America. Journal of Parasitology 39(5), 568.Google Scholar
Hutton, RF (1964) A Second List of Parasites from Marine and Coastal Animals of Florida. Transactions of the American Microscopical Society 83(4), 439.Google Scholar
Jägerskiöld, LA (1904) Scaphanocephalus expansus (Crepl.), eine genital-napftragende Distomide. In Results of the Swedish Zoological Expedition to Egypt and the White Nile 1901, Part I.Google Scholar
Johnston, SJ (1917) On the Trematodes of Australian birds. Journal and Proceedings of the Royal Society of New South Wales 50, 187226.Google Scholar
Katahira, H, Shimose, T and Kanaiwa, M (2021) New host records for Scaphanocephalus adamsi (Platyhelminthes: Trematoda: Heterophyidae) from commercial fishes off Yaeyama Islands. Fauna Ryukyuana 59, 2126.Google Scholar
Kinsella, JM, Cole, RA, Forrester, DJ and Roderick, CL (1996) Helminth parasites of the osprey, Pandion haliaetus, in North America. Journal of the Helminthological Society of Washington 63(2), 262265.Google Scholar
Kohl, ZF, Calhoun, DM, Elmer, F, Peachey, RBJ, Leslie, KL, Tkach, Kinsella JM VV and Johnson, PTJ (2019) Black-spot syndrome in Caribbean fishes linked to trematode parasite infection (Scaphanocephalus expansus). Coral Reefs 38(5), 917930.Google Scholar
Lee, O (1966) Some helminths from Malayan wild birds with descriptions of two new species. Bulletin of the Natural History Museum of Singapore 33, 7781.Google Scholar
Maddison, WP and Maddison, DR (2011) Mesquite: a modular system for evolutionary analysis. Version 3.6.1.Google Scholar
Miller, MA, Pfeiffer, W and Schwartz, T (2010) Creating the CIPRES Science Gateway for inference of large phylogenetic trees. Gateway Computing Environments Workshop (GCE), http://www.phylo.org.Google Scholar
Montoya-Mendoza, J, Morales-Sánchez, G, Arenas-Fuentes, V and González-Solís, D (2021) Parasitic Helminth Community of the Puddingwife Wrasse Halichoeres radiatus (Linnaeus) in Reefs from Veracruz, Mexico. Journal of Parasitology 107(6).Google Scholar
Nadler, SA, D’Amelio, S, Fagerholm, H-P, Berland, B and Paggi, L (2000) Phylogenetic relationships among species of Contracaecum Railliet & Henry, 1912 and Phocascaris Høst, 1932 (Nematoda:Ascaridoidea) based on nuclear rDNA sequence data. Parasitology 121(4), 455463.Google Scholar
Overstreet, RM and Hawkins, WE (2017) Diseases and Mortalities of Fishes and Other Animals in The Gulf of Mexico. In Springer New York, New York, 1589–1738Google Scholar
Pearson, J (2008) Family heterophyidae Leiper, 1909. In Gibson, DI, Jones, A, Bray, RA (eds), Keys to the Trematoda, Volume 3, pp. 113141. Wallingford: CAB International.Google Scholar
Pérez-Urbiola, JC, Flores Herrera, AR, Nieth Kmoth, AR and Inohuye- Rivera, RB (1994) Registro de una metacercaria del género Scaphanocephalus Jagerskiold, 1903 (Trematoda: Heterophyidae), en peces del Golfo de California, B.C.S., México. Revista De Investigación Científica - Area De Ciencias Del Mar 5, 6568.Google Scholar
Pleijel, F, Jondelius, U, Norlinder, E, Nygren, A, Oxelman, B, Schander, C, Sundberg, P and Thollesson, M (2008) Phylogenies without roots? A plea for the use of vouchers in molecular phylogenetic studies. Molecular Phylogenetics and Evolution 48(1), 369371.Google Scholar
Posada, D (2008) JModelTest: Phylogenetic Model Averaging. Molecular Biology and Evolution 25(7), 12531256.Google Scholar
Rambaut, A (2012) FigTree v1.4.0., UK, Institute of Evolutionary Biology, University of Edinburgh.Google Scholar
Ronquist, F, Teslenko, M, Van der Mark, P, Ayres, DL, Darling, A, Höhna, S, Larget, B, Liu, L, Suchard, MA and Huelsenbeck, JP (2012) MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Systematic Biology 61, 539542.Google Scholar
Schmidt, GD and Huber, PM (1985) Two rare helminths in an osprey, Pandion haliaetus, in Mexico. Proceedings of the Helminthological Society of Washington 52(1), 139140.Google Scholar
Shephard, JM, Catterall, C and Hughes, J (2005) Long-term variation in the distribution of the White-bellied Sea-Eagle (Haliaeetus leucogaster) across Australia. Austral Ecology 30, 131145.Google Scholar
Shimose, T, Katahira, H and Kanaiwa, M (2020) Interspecific variation of prevalence by Scaphanocephalus (Platyhelminthes: Trematoda: Heterophyidae) metacercariae in parrotfishes (Labridae: Scarini) from an Okinawan coral reef. International Journal for Parasitology. Parasites and Wildlife 12, 99104.Google Scholar
Silvestro, D and Michalak, I (2011) raxmlGUI: a graphical front-end for RAxML. Organisms Diversity & Evolution 12(4), 335337.Google Scholar
Skinner, RH (1978) Some external parasites of Florida fishes. Bulletin of Marine Science 28, 590595Google Scholar
Smogorzhevskaya, LA (1956) Trematodes of fish-eating birds of the River Dnieper valley. Parazitologicheskii Sbornik, 16, 244263.Google Scholar
Sokolov, S, Frolov, E, Novokreshchennykh, S and Atopkin, DM (2021) An opisthorchiid concept of the genus Liliatrema (Trematoda: Plagiorchiida: Opisthorchioidea): an unexpected systematic position. Zoological Journal of the Linnean Society 192(1), 2442.Google Scholar
Tamura, K, Stecher, G, Peterson, DS, Filipski, A and Kumar, S (2013) MEGA6: Molecular Evolutionary Genetics Analysis Version 6.0. Molecular Biology and Evolution 30(12), 27252729.Google Scholar
Tatonova, YV and Besprozvannykh, VV (2019) Description of a new species, Cryptocotyle lata sp. nov., and discussion of the phylogenetic relationships in Opisthorchioidea. Parasitology International 72, 101939. doi:10.1016/j.parint.2019.101939.Google Scholar
Tubangui, MA (1933) Trematode parasites of Philippine vertebrates, VI. Descriptions of new species and classification. The Philippine Journal of Science 52(2), 167197.Google Scholar
Wiley, JW, Poole, A and Clum, NJ (2014) Distribution and Natural History of the Caribbean Osprey (Pandion haliaetus ridgwayi). Journal of Raptor Research 48(4), 396407.Google Scholar
WoRMS World Register of Marine Species (2023) Scaphanocephalus australis Johnston, 1917. Accessed at: https://www.marinespecies.org/aphia.php?p=taxdetails&id=727966Google Scholar
Yamaguti, S (1942) Studies on the helminth fauna of Japan. Part 38. Larval trematodes of fishes. Japanese Journal of Medical Sciences. VI. Bacteriology and Parasitology 2(3), 131160.Google Scholar
Yamaguti, S (1958) Systema helminihum. Volume I. The digenetic trematodes of vertebrates., New York: Interscience Publishers Inc.Google Scholar
Figure 0

Figure 1. Sampling collection in Yucatán Peninsula, Mexico. Pandion haliaetus, 1. Champotón, Campeche (19º21’40”N, 90º43’5”W). Mugil curema, 2. La Carbonera (21º8’1”N, 90º7’55”W) 3. Celestún (20º50’53N, 90º24’22”W), 4. Dzilam (21º23’40”N, 88º53’13”W) 5. Ría Lagartos (21º35’45”N, 88º8’48”W), Yucatán. Localities where Scaphanocephalus spp. were recovered are in red. In blue, localities where the parasite was not found.

Figure 1

Figure 2. A–C) Hologenophore of Scaphanocephalus expansus from Pandion haliaetus of Champotón, Campeche. D) Hologenophore of Scaphanocephalus sp. from Mugil curema in Celestún, Yucatán. CNHE and GenBank accession numbers are indicated. Scale bars= (A–C) 1 mm, D 300 μm.

Figure 2

Table 1. Comparative morphometric data for Scaphanocephalus expansus

Figure 3

Figure 3. A) Scaphanocephalus expansus from Pandion haliaetus. Schematic whole worm, ventral view. Scanning electron photomicrographs, B) whole worm; C) Oral sucker; D) Tegumental spines at level of wing-like expansions; E) Spines at the level of the third part of the body. Scale bars = (A, B) 1 mm; (C), 50 μm; (D, E) 5 μm.

Figure 4

Figure 4. Consensus Bayesian inference (BI) tree and Maximum likelihood (ML) tree inferred from the large subunit from nuclear ribosomal DNA. Numbers on internal nodes show posterior probabilities (BI) and ML bootstrap clade frequencies. Sequences generated in this study in bold. Grey: specimens from Arabian Gulf. Blue: specimens from the Caribbean Sea. Green: specimens from Yucatán Peninsula, Mexico. Red: specimens from the Mediterranean Sea. Scale bar shows the number substitutions per site.