<<

El Colegio de la Frontera Sur

Neréididos (Polychaeta: ) del Gran Caribe

TESIS Presentada como requisito parcial para optar al grado de Maestría en Ciencias en Recursos Naturales y Desarrollo Rural

por

Víctor Manuel Conde Vela

2015

Para las familias Conde Vela y Te Gómez

En especial para ti, Astrid AGRADECIMIENTOS

Esta contribución es resultado del esfuerzo sumado de muchas personas, quienes aportaron de diversas maneras a su culminación. Dedicar el tiempo necesario a este trabajo no sólo costó al autor, sino también a quienes lo apoyaron y prestaron su valioso tiempo y esfuerzo. Gracias por todo, y es para ustedes.

Quiero expresar mi más sincero agradecimiento a mi director de tesis Sergio I.

Salazar Vallejo, ya que sin su apoyo y su vasta experiencia en el terreno de los poliquetos, esta contribución no hubiera tenido los mismos resultados. Gracias por la insistencia en hacer cada vez más y mejor, y por los valiosos comentarios y revisiones que me guiaron durante toda la maestría.

A Luis F. Carrera Parra y a Jesús Á. de León González, por aceptar ser parte del comité evaluador, y proporcionar valiosos comentarios que permitieron mejorar el escrito final.

A mis sinodales adicionales Eduardo Suárez Morales, Víctor H. Delgado Blas y

David González Solís, por aceptar evaluar el escrito en tan apretadas situaciones, y por sus importantes sugerencias que proporcionaron al mismo. A Patricia Bardales

Pastrana por las innumerables facilidades administrativas proporcionadas.

A Nancy Voss (UMML) por facilitar el material base de esta contribución. Gracias a su confianza y el préstamo del material de manera incondicional, yo y varios colegas hemos podido realizar nuestras metas en maestría y doctorado. A Karen Osborn y

Geoff Keel (USNM) quienes facilitaron mi estancia de investigación en el

Smithsonian Institution (Washington, EUA), y a Leslie Harris por mi estancia en el Natural History Museus of Los Angeles County (LACM-AHF). A todos ellos gracias por la confianza y el acceso ilimitado a las colecciones a su cargo. A Jesús Á. de

León Gónzález (UANL) y a Leslie Harris por prestar material valioso para esta contribución.

A Leslie Harris y David Ocker por la cálida hospitalidad y diversas enseñanzas durante mi estancia en Los Ángeles.

Agradezco enormemente a mi esposa Astrid Estefanía Te Gómez, ya que sin su completo apoyo este trabajo no hubiera sido posible. Muchas gracias por estar ahí cuando más lo necesitaba.

A las familias Conde Vela y Te Gómez, quienes siempre me han apoyado de innumerables maneras, y por dar parte de su tiempo y recursos para la realización de este trabajo, sin los cuales las metas de este trabajo no se hubieran alcanzado.

A los colegas del laboratorio, Emilia, Reyna, Isabel, Jani, Stephanie, Tulio y Víctor

Hugo, que hicieron más amena la estadía en ECOSUR, además de compartir sus experiencias y conocimientos.

De manera especial, agradezco a todos mis compañeros de la maestría, que tuvieron muchas oportunidades de convivir conmigo y hacer esta maestría la mejor que pude haber tenido.

Agradezco a CONACyT por la beca que me otorgó, sin la cual no se hubiera podido dedicarme a mis estudios de maestría. Además a Sergio I. Salazar Vallejo y Luis F.

Carrera Parra, quienes aportaron recursos importantes para la realización de las estancias en el extranjero. CONTENIDO

Introducción …………………………………………………………………………………… 1

Capítulo 1. Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi

González-Escalante & Salazar-Vallejo, 2003 and description of N. confusa sp. n.

(Annelida, Nereididae) ……………………………..………………………….……….……… 7

Capítulo 2. Nereididae (Polychaeta: ) from the Grand Caribbean and other related regions ………..…………………..………………………………………………….. 43

Conclusiones ………………………………………………………………….…………… 282

Referencias ………………………………………………………………….……………… 285

Anexos ………………………………………………………………………………………. 292

La presente tesis no debe ser considerada como una publicación en el sentido del

Código Internacional de Nomenclatura Zoológica, y los nombres científicos mencionados en ésta no deben ser citados en ninguna forma, para evitar ser considerados nomina nuda.

This thesis is not a publication in the sense of the International Code of Zoological

Nomenclature, and scientific names mentioned in it must not be cited in any way to avoid them to become nomina nuda. INTRODUCCIÓN

La familia Nereididae de Blainville, 1818 es un grupo amplio de poliquetos que se distingue por tener una faringe tubular eversible con papilas y/o paragnatos en un arreglo regular, aunque pueden carecer de ambos, así como un par de grandes mandíbulas en su extremo, además de presentar una morfología parapodial compleja, con lóbulos y lígulas. Los neréididos representan una de las familias más comunes en diversos sustratos; tienen una amplia distribución, debido a que son frecuentes han sido extensamente estudiados (Wilson, 2000).

Actualmente, la familia tiene 44 géneros y unas 460 especies (de León-González, 2009), comprendidos en tres subfamilias: Nereidinae de Blainville, 1818, Namanereidinae

Hartman, 1959, y Gymnonereidinae Banse, 1977 (Fitzhugh, 1987; Glasby, 1991). Bakken y Wilson (2005) y Santos et al. (2005) notaron algunos problemas en la clasificación actual a distintos niveles, como el bajo soporte filogenético de las subfamilias actuales, el posible restablecimiento de dos subfamilias sinonimizadas, Dendronereidinae Pillai,

1961 y Notophycinae Knox & Cameron, 1970.

En América tropical se han citado 25 géneros y 104 especies de neréididos (de León-

González, 2009). Sin embargo persisten problemas taxonómicos en el grupo, principalmente la supuesta existencia de un gran número de especies consideradas cosmopolitas o de amplia distribución; de especial interés son las que se consideran como anfiamericanas, es decir, que viven en las costas del Pacífico y Atlántico americano, donde al menos 18 están en esa categoría (de León-González, 2009).

1

En el Gran Caribe, que comprende el Golfo de México, el Mar Caribe, Bermudas y las costas nortes de Brasil (Salazar-Vallejo, 2000), el estudio exhaustivo de la poliquetofauna del Gran Caribe tuvo su origen en las campañas oceanográficas puntuales o de alcance mundial, como las expediciones del Challenger por el escocés Charles W. Thomson

(McIntosh, 1885), las realizadas por el estadounidense Alexander Agassiz en el Mar

Caribe y el Golfo de México (Ehlers, 1887; Augener, 1906), y por los daneses Anders S.

Ørsted y Henrik Kröyer en el Mar Caribe, incluyendo América Central y Sudamérica

(Grube, 1857, 1858, 1859).

Otras campañas importantes han sido las Expediciones del Atlantis a las Indias

Occidentales (Hartman, 1942), la Expedición de Barbados-Antigua de la Universidad de

Iowa en 1918 (Treadwell, 1924), las Expediciones Smithsonian-Bredin en el Mar Caribe

(de León-González, Solís-Weiss y Ochoa-Rivera, 1999), y las Expediciones del Mar profundo de la Universidad de Miami, realizadas por Gilbert L. Voss, Richard Robins y C.

Staiger en los años 1962 a 1975. Esta última es la principal fuente de especímenes para esta investigación. Además, otras expediciones danesas sirvieron para conocer y describir varios neréididos de agua dulce y de depósitos de agua continentales, principalmente las dirigidas por P. W. Hummelinck en gran parte de las Antillas

(Wesenberg-Lund, 1958; Hummelinck, 1981).

A pesar de que nuevas especies fueron descritas a partir del material colectado en aquellas expediciones y en otros trabajos puntuales, muchas de éstas fueron sinonimizadas con especies de otras partes del mundo. La falta de descripciones actualizadas en varias especies, basadas en la revisión de materiales originales o tipo de las especies involucradas, ocasionó la introducción de nombres foráneos en listados

2 regionales. Debido a la tendencia contemporánea a la conservación del nombre más antiguo, especialmente después del esfuerzo de Olga Hartman (Hartman, 1959a, b) en su catálogo mundial de poliquetos, los nombres regionales del Gran Caribe fueron considerados como variantes o subespecies, esto aunado a que muchas descripciones originales son breves, con la consecuente pérdida de caracteres diagnósticos. Además algunas especies fueron sinonimizadas con otras, a pesar de que fueron parte de una revisión más profunda (e.g. Pettibone, 1971).

Diversos autores han abordado el problema en varios grupos para el Gran Caribe con estudios morfológicos finos, herramientas moleculares o una combinación de ambas, y han demostrado que la conclusión de la supuesta amplia distribución se trata de una mala taxonomía o de estudios sistemáticos conservadores, es decir, con poca introducción de caracteres novedosos que permitieran separar a las especies (e.g. Klautau et al., 1999;

Carrera-Parra y Salazar-Vallejo, 2011; Yáñez-Rivera y Carrera-Parra, 2012; Yáñez-

Rivera y Salazar-Vallejo, 2011).

En el caso de los neréididos, se han realizado varios esfuerzos para describir las especies de neréididos presentes a nivel local, principalmente de las costas caribeñas de México

(de León-González y Solís-Weiss, 1997, 1998; de León-González, Solís-Weiss y Ochoa-

Rivera, 1999; de León-González, Solís-Weiss y Valadez-Rocha, 2001; Salazar-Vallejo y

Jiménez-Cueto, 1997), Venezuela (e.g. Liñero-Arana y Reyes-Vásquez, 1979; Liñero-

Arana y Díaz-Díaz, 2007; Vanegas-Espinosa, Díaz-Díaz y Liñero-Arana, 2007) y Brasil

(Santos y Lana, 2000, 2001, 2003; Santos, 2007). Amaral et al. (2013) realizaron una compilación de las especies de poliquetos registrados para Brasil, junto con la gran cantidad de literatura generada. Sin embargo, son pocos los trabajos con una amplia

3 cobertura geográfica; de León-González, Solís-Weiss y Ochoa-Rivera (1999) registraron

23 especies pertenecientes a 7 géneros para el Mar Caribe. Glasby (1999) registró 13 especies de namaneréidinos para la región; de las cuales, 8 se trataron de nuevas especies, reflejando la falta de conocimiento de estos neréididos en el Gran Caribe.

Además, Williams (2004) revisó el estado actual del género Namanereis en el Caribe.

No se han realizado esfuerzos de revisión para aclarar el estatus de las especies o revertir las sinonimias prevalentes, a pesar de los numerosos esfuerzos de compilación de listas locales y regionales de las especies de poliquetos para Gran Caribe (e.g. Perkins y

Savage, 1975; Salazar-Vallejo, 1996; Báez y Ardila, 2003; Dean, 2012), que sintetizan los nombres registrados y la bibliografía correspondiente con el fin de impulsar y facilitar las revisiones taxonómicas.

Basado en dichas listas, y en revisiones adicionales de literatura, hay unas 108 especies nominales de neréididos registradas en el Gran Caribe, pertenecientes a 21 géneros. De

éstas, 38 corresponden a registros cuestionables, ya que son nombres de especies ajenas a la región; 16 son especies regionales, pero que ameritan ser redescritas ya que actualmente son consideradas sinónimas de otras especies (principalmente de N. riisei

Grube & Ørsted in Grube, 1858), y/o las descripciones originales son breves y no aseguran la validez de la especie (Dean, 2012; Salazar-Vallejo, 1996; de León-González,

2009). Al menos 5 especies ameritan revisiones para su posible restablecimiento y otras

5 especies requieren definir sus afinidades genéricas; y solamente 44 son consideradas especies válidas. Así, el 41% de las especies son válidas, mientras que el 59% restante tiene algún estatus cuestionable.

4

Ciertas especies con estatus cosmopolita son ampliamente reportadas, como Alitta succinea (Leuckart, 1847), Nereis pelagica Linnaeus, 1758 y

(Audouin & Milne-Edwards, 1834) (Salazar-Vallejo, 1996; Dean, 2012). Por otro lado, especies como Ceratonereis mirabilis Kinberg, 1865 y Nereis riisei son consideradas de amplia distribución y para esta última, varias especies del Gran Caribe han sido sinonimizadas con ella. Finalmente, especies anfiamericanas comunes en los registros son Pseudonereis gallapagensis Kinberg, 1865 y P. variegata (Grube & Kröyer en Grube,

1858), siendo el único género donde los registros pertenecen a especies del Pacífico americano, sin nombres disponibles para el Gran Caribe. Por el contrario, el único género restringido al Gran Caribe es Stenoninereis Wesenberg-Lund, 1959, con dos especies válidas y una en sinonimia (Wesenberg-Lund, 1958; de León-González y Solís-Weiss,

1997).

Sin embargo, se han realizado algunos estudios que demuestran la posibilidad de separar neréididos considerados anfiamericanos o cosmopolitas con caracteres morfológicos y con herramientas moleculares (Conde-Vela y Salazar-Vallejo, 2015; Villalobos-Guerrero y Carrera-Parra, 2015).

Dado el panorama anterior de confusión a nivel regional, el objetivo central de este estudio fue realizar un estudio exhaustivo de las 108 especies descritas y registradas en el Gran Caribe, con base en la revisión de material tipo y no tipo disponible de diferentes colecciones, principalmente de los especímenes provenientes de las Expediciones del

Mar Profundo de la Universidad de Miami, que tiene una amplia cobertura geográfica de la región. De las que no se pudo obtener material, se incluyen comentarios basados en la literatura. Esto con el objetivo de delimitar la distribución de especies consideradas

5 como cosmopolitas y anfiamericanas en el Gran Caribe, evaluar el restablecimiento de especies regionales, y la posible descripción de nuevas especies para el Gran Caribe.

Adicionalmente, en el presente trabajo se muestran algunos problemas puntuales en la terminología de parápodos actualmente aceptada, por lo que se introducen nuevos términos para describir las estructuras parapodiales, tratando de hacer más precisas las descripciones por un lado, y por otro esclarecer algunos puntos problemáticos en la interpretación de la morfología del grupo.

6

Capítulo 1

Redescriptions of Nereis oligohalina (Rioja, 1956) and N. garwoodi González-Escalante

& Salazar-Vallejo, 2003 and description of N. confusa n. sp. (Annelida, Nereididae)

7

A peer-reviewed open-access journal ZooKeys 518: 15–49Redescriptions (2015) of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 15 doi: 10.3897/zookeys.518.9564 RESEARCH ARTICLE http://zookeys.pensoft.net Launched to accelerate biodiversity research

Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi González-Escalante & Salazar-Vallejo, 2003 and description of N. confusa sp. n. (Annelida, Nereididae)

Víctor M. Conde-Vela1, Sergio I. Salazar-Vallejo1

1 El Colegio de la Frontera Sur, CONACYT, Departamento de Sistemática y Ecología Acuática, Chetumal, México

Corresponding author: Víctor M. Conde-Vela ([email protected])

Academic editor: Chris Glasby | Received 12 March 2015 | Accepted 9 August 2015 | Published 25 August 2015

http://zoobank.org/4A0139A5-AE60-4988-ACFA-A16E8EEF35DB

Citation: Conde-Vela VM, Salazar-Vallejo SI (2015) Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi González-Escalante & Salazar-Vallejo, 2003 and description of N. confusa sp. n. (Annelida, Nereididae). ZooKeys 518: 15–49. doi: 10.3897/zookeys.518.9564

Abstract Type material of several described by Enrique Rioja from Mexican coasts are lost, and the current status of some species is doubtful. Nereis oligohalina (Rioja, 1946) was described from the Gulf of Mexico, but it has been considered a junior synonym of N. occidentalis Hartman, 1945, or regarded as a distinct species with an amphiamerican distribution. On the other hand, N. garwoodi González-Escalante & Salazar-Vallejo, 2003, described from Chetumal Bay, Caribbean coasts, could be confused with N. oligohalina. In order to clarify these uncertainties, N. oligohalina is redescribed based on specimens from the Mexican Gulf of Mexico, including a proposed neotype; further, N. garwoodi is redescribed including the selection of lectotype and paralectotypes, and N. confusa sp. n. is described with material from the Gulf of California. A key for the identification of similar species and some comments about speciation in nereidid are also included.

Keywords Amphiamerican, , estuarine nereidids, cryptic species, Polychaeta

Copyright V.M. Conde-Vela, S.I. Salazar-Vallejo. 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. 16 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015)

Introduction

Among the non-marine polychaetes, the family Nereididae de Blainville, 1818 has the largest number of brackish and freshwater species (61), and 31 out of these species oc- cur in estuaries and coastal lagoons (Glasby et al. 2009). Of the 40 species of nereidid species recorded from the Gulf of Mexico (Fauchald and Solís-Weiss 2009), seven are reported in brackish or freshwater areas, and among the 10 species belonging to Nereis, only N. oligohalina (Rioja, 1946) is reported from estuaries (Glasby et al. 2009). Enrique Rioja documented extensively the Mexican polychaetes from Pacific or Atlantic coasts in a series of papers; unfortunately, his material is lost, and most spe- cies require designation of neotypes (Salazar-Vallejo 1989). Rioja (1946) dealt with three estuarine nereidids from Veracruz, Mexico; he regarded one as a known spe- cies, succinea (Leuckart, 1847), and the two others were described as new: Neanthes oligohalina and Lycastopsis tecolutlensis. The former species is now regarded as belonging in Alitta, but it differs from the North Sea species (T.F. Villalobos-Guer- rero, pers. comm.); L. tecolutlensis was regarded as a junior synonym of Namanereis amboinensis (Pflugfelder, 1933), nowadays a widespread species (Glasby 1999); and N. oligohalina has been regarded as amphiamerican (Dean 2001), or restricted to At- lantic coasts (Santos and Lana 2003, Liñero-Arana and Díaz-Díaz 2007). However, other amphiamerican species have been shown to be restricted to one coast or the other, often resulting in description of new taxa (e.g. Carrera-Parra and Salazar-Valle- jo 2011; Yáñez-Rivera and Carrera-Parra 2012). On the other hand, the Caribbean species N. garwoodi González-Escalante & Salazar-Vallejo, 2003 could be confused with N. oligohalina, and without an updated description of the latter, a synonymy can be anticipated. However, Nereis oligohalina and N. garwoodi have morphological differences that separate them. In this contribution our objectives were first, to redescribe N. oligohali- na based upon material collected from Veracruz, including topotypes, and to propose a neotype. Second, to redescribe N. garwoodi to clarify some doubtful features in the original description, and to select lectotype and paralectotypes specimens from the syntype series. Third, to recognize what has been regarded as N. oligohalina from the Mexican Pacific as a distinct species and describe it asN. confusa sp. n., based upon material from the Gulf of California. Further, a key to identify similar Nereis species and comments about species delimitation are also included.

Material and methods

Specimens studied are deposited in the Reference Collection of El Colegio de la Frontera Sur, Chetumal (ECOSUR) including ethanol-fixed specimens (ECOSUR-OH), and in the Polychaetological Collection of the Universidad Autónoma de Nuevo León (UANL). Topotypes of N. oligohalina from Estero Casitas, Nautla were examined, including additional specimens that were recently collected along the coast of Veracruz, Mexico, Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 17

Gulf of Mexico. Specimens of N. confusa sp. n. from Bahía La Paz and Bahía de Los Ángeles, Gulf of California were found in unidentified material in ECOSUR, now formally deposited. To assess variation in paragnath morphology, specimens of N. pelagica Linnaeus from England (ECOSUR P2840), and Pseudonereis sp. from the Caribbean Sea (ECOSUR P1170), were also examined. The best preserved specimens were used for designation of type material. Some specimens were fixed and preserved in 96% ethanol directly, otherwise the specimens were fixed with formalin and later preserved in 70% ethanol. For analysis of variation, type materials and a number of non-type specimens for each species were measured. Total body length (TL), length up to chaetiger 3 (L3) or 10 (L10), width at same chaetigers (W3 and W10), number of chaetigers (nC), and length of longest tentacular cirrus (rTC) were measured with a millimeter rule under the stereomicroscope. TL was measured from palp tips to the end of the pygidium,

W3 and W10 were measured excluding parapodia. Also, paragnath numbers in all areas were counted; if the pharynx was not everted, a ventral dissection was made, and areas VII-VIII were described as if the pharynx was exposed. With these results simple de- scriptive statistics (mean, range and standard deviation) were performed. As different fixation methods were used, a Mann-Whitney U-test was used for evaluating if there were significant differences in body measures as a result of fixation method. For the microscopical observation of parapodial features and chaetae, right-side parapodia along body were removed and mounted in semi-permanent slides; the pho- tographs were made with a digital camera, and distal-view drawings of parapodia were included to depict spatial disposition of ligules and chaetae. Descriptions of pigmen- tation patterns were included, because they are consistent and useful for recognizing the three species; other authors have noted their utility for identifying cryptic nereidid species (Read 2007, Glasby et al. 2013). Bakken and Wilson’s (2005) terminology was followed for describing parapodia, and Bakken et al. (2009) for paragnaths. Parapodia from both atokes and epitokes (if available) were illustrated to show parapodial changes along the body. For determining dorsal cirri length and position of attachment, we considered the beginning of the dor- sal or notopodial ligule to be approximately at the same vertical position of attachment as the ventral cirri; therefore, the relative length of dorsal cirrus was measured from that position toward the distal end of dorsal or notopodial ligule. The dorsal cirrus was considered basally attached if placed at, or near such a position, or medially attached if it was displaced more distally from that position. Also, the reach of the dorsal cirrus and its relative length in respect of the dorsal or notopodial ligules were considered as separate attributes. For determining the reach of the dorsal cirrus, the tips of the both dorsal cirrus and dorsal or notopodial ligules were taken into account; if the dorsal cirrus tip extended beyond the tip of the dorsal or notopodial ligule, then we report “dorsal cirrus extended beyond dorsal/notopodial ligule” rather than “dorsal cirrus longer than dorsal/notopodial ligule”. On the con- trary, if dorsal cirrus is shorter or not exceeding the dorsal or notopodial ligules, then we report “dorsal cirrus not extended beyond dorsal/notopodial ligule”. In the species 18 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015) herein treated, length of dorsal cirri and length of dorsal or notopodial ligules were generally subequal, and the dorsal cirri change their attachment along the body, but not their length necessarily. For the designation of lectotype and neotype, the International Code of Zoologi- cal Nomenclature (ICZN 1999) was followed. The designation of a neotype forN . oligohalina follows Article 75, and the designation of a lectotype for N. garwoodi fol- lows Article 74 (ICZN 1999). The non-formal term ‘paraneotypes’ is used for figured topotypic specimens (Evenhuis 2008), and their utility has been pointed out elsewhere (Salazar-Vallejo 2011, Sendall and Salazar-Vallejo 2013).

Results Family Nereididae de Blainville, 1818

Genus Nereis Linnaeus, 1758

Type species. Nereis pelagica Linnaeus, 1758, by subsequent designation (Hartman 1948:63). Remarks. Linnaeus (1758:654) listed five species under Nereis: N. lacustris (now Stylaria lacustris, an oligochaete), N. caerulea (questionable after Hartman 1959:254), N. gigantea (after Hartman 1959:259, same as Hermodice carunculata (Pallas, 1766), rendering it a nomen oblitum because it would have priority over Pallas’ name), N. pelagica, and N. noctiluca. Further, it was Hartman (1948:63) who fixed the type species, and therefore this should be regarded as a subsequent designation (ICZN 1999, Art. 69.1), in contrast to Bakken and Wilson (2005) who regarded it as an original designation. Although the species described here have more attributes than those included in the current diagnosis of the (Bakken and Wilson 2005), the generic diagnosis was not modified because it first requires a redescription of the type-species, and a phylogenetic analysis with subsequent delimitation of the genus. Among the tradi- tionally used features for descriptions and delimitation of Nereis species are some that are highly variable, especially paragnath number. Bakken et al. (2009) made a useful revision of paragnath morphology and introduced new terminology to standardize de- scriptions. In addition to conical paragnaths, the species described here present other types of paragnaths that are not currently included in the diagnosis of the genus. Coni- cal paragnaths are pointed to various degrees, being more acute in the maxillary ring, especially on area II. The pyramidal paragnaths inN . oligohalina and N. garwoodi have quadrilateral bases but they can also be polygonal, having more defined surfaces in the latter species. Further, N. confusa sp. n. apparently has smooth bars on area IV, but a closer inspection confirms that this is an artifact because the bars are formed by lateral and basal fusions of some small conical paragnaths (Fig. 6H); these modified struc- tures were regarded as melted paragnaths (Bakken et al. 2009). However, Glasby et al. (2011) suggested limiting use of the term for conical paragnaths mounted on a plate- like basement as occurs in Neanthes pachychaeta (Fauvel, 1918), and Villalobos-Guer- Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 19 rero and Carrera-Parra (2015) found paragnaths on a soft basement in A. acutifolia (Ehlers, 1901). Because neither a basement is present in N. confusa sp. n., we suggest the term ‘merged’ for paragnaths fused at the base but without formation of a plate. In his revision of Pseudonereis Kinberg, 1865, Bakken (2007) introduced the term ‘P-bar’, which was later defined by Bakken et al. (2009) as “small bars having a pro- truding apex in one end of the bar”; and they can appear in areas II, III, IV and VII- VIII, often accompanied by conical paragnaths. The monophyly ofPseudonereis was supported by, among other characters, the presence of both P-bars and paragnaths in comb-like rows (Bakken 2007). The Nereis species studied herein have two main rows, each one with other two sub-rows; the anterior-most sub-rows are often aligned horizontally, while the posterior-most ones form a jagged line. The anterior-most rows have P-bars alternating with conical or pyramidal paragnaths in a similar way as in Pseudonereis (Fig. 6D–F), which has been also reported for Alitta (Villalobos-Guerrero and Carrera-Parra 2015). Therefore, P-bars are not an exclusive feature ofPseudonereis as Bakken et al. (2009) concluded. The neuropodial postchaetal lobe has been considered absent for Nereis species in recent phylogenetic analyses (Bakken and Wilson 2005, Santos et al. 2005). Neverthe- less some authors indicate its presence; Read (1980) for N. ovarius (Read), Santos and Lana (2003) for N. pseudomoniliformis (Santos and Lana), Chambers and Garwood (1992) for N. pelagica (and corroborated by us), and Darbyshire (2014) for N. eugeniae (Kinberg, 1865). Also, the three species herein described have postchaetal lobes shorter or subequal than neuroacicular ligules, and in epitoke specimens these lobes carry nata- tory lamellae; therefore, we considered postchaetal lobes as present in Nereis. Reproductive nereidids or epitokes can have two or three different regions; parapo- dial cirri and the pygidium are transformed is especially relevant for chemorreception, parapodial lobes or ligules are expanded and chaetae replaced for swimming (Herpin 1925, Boilly-Marer 1972). Charrier (1920) studied muscular tissue transformation associated with epitoky in the commensal species Nereis fucata (Savigny in Lamarck, 1818) (currently belonging to Neanthes Kinberg). He observed that parapodial cirri vary in some features as the length relative to corresponding ligules, the modifications of attachment and the displacement along the body; in fact, the attachment site of dorsal cirrus often show a distal displacement along notopodial ligules.

Nereis oligohalina (Rioja, 1946) Figures 1, 2, 6A, D, J–L, O, Q

Neanthes oligohalina Rioja 1946: 207–210, pl. 1, figs 3–6, pl. 2, figs 13–19; 1947: 529, 531; 1960: 295. Nereis oligohalina Hartman 1951: 46; 1954: 414.

Type material. Veracruz, Mexico. Neotype ECOSUR 0172 and paraneotypes ECO- SUR 0173 (5), mouth of Actopan River (19°25'2.95"N, 96°19'32.28"W), Chacha- 20 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015) lacas Sandbar, Gulf of Mexico, 1 m depth, on Crassostrea virginica reef, fine sediment, April 10 2012, Coll. V.M. Conde-Vela, A.E. Te-Gómez. Additional material. Veracruz, Mexico. ECOSUR P2826 (15), and ECOSUR P2827 (37), Mouth of Actopan River (19°25'2.95"N, 96°19'32.28"W), Chachalacas Sandbar, Gulf of Mexico, 1 m depth, on Crassostrea virginica reef, fine sediment, April 10 2012, Coll. C. Licona-Rosado, V.M. Conde-Vela, A.E. Te-Gómez. ECOSUR P2828 (1) Laguna de Alvarado, St. 8 (18°45'20.34"N, 95°46'29.04"W), December 6 2012, in rocks, Coll. J.M. Aguilar-Camacho. ECOSUR-OH-P0760 (23), Las Bar- rillas, St. 4 (18°11'20.15"N, 94°35'56.97"W), December 5 2012, on C. virginica culture, Coll. J. Cruz-Terrón. ECOSUR-OH-P0761 (15), Laguna Grande, Mand- inga, St. 28 (19°1'54.96"N, 96°4'8.10"W), December 10 2012, on oyster, 11.41‰, 27.24 °C, Coll. T.F. Villalobos-Guerrero, MA. Tovar-Hernández, J.M. Aguilar-Ca- macho. ECOSUR-OH-P0762 (3), Laguna Grande, Mandinga, St. 26 (19°2'20.64’’ N 96°4'24.24"W), December 10 2012, on mangrove and oyster, 10.74 ‰, 26.63 °C, Coll. J.M. Aguilar-Camacho, T.F.Villalobos-Guerrero. UANL-3918 (6), Estero Casitas, Nautla, March 25 1990, Coll. A. Contreras-Arquieta. Neotype locality. Mouth of Actopan River (19°25'2.95"N, 96°19'32.28"W), Gulf of Mexico, in C. virginica (Gmelin) reef, in muddy sediment, 1 m depth. Description. Neotype complete (ECOSUR 0172), atokous female. Body taper- ing, 38 mm long, 2.1 mm wide, 74 chaetigers, filled with oocytes. Body yellowish, red- dish brown pigmentation present dorsally on first quarter of body, discoloring towards midbody chaetigers; lateral oblique pale lines along chaetigers 1–9 (Fig. 1A), replaced by fingerprint-like lines from chaetiger 10 (Fig. 6L), anterior margin of segments with thin transverse band (Fig. 1A). Prostomium with pigmentation reddish brown along inner half of palps and around eyes; two broad hourglass-shaped lines extending from antennae towards eyes, separated by a thin pale line and two oval patches, one on each side of darker areas (Fig. 1A). Peristomium pigmented, pale lines present (Fig. 1A). Prostomium 1.5 times longer than wide; antennae cirriform, slightly passing palps; eyes subequal, black, in trapezoidal arrangement (Fig. 1A). Peristomium three times longer than first chaetiger; tentacular cirri with short ceratophores; dorsal cirri longer than ventral ones, longest posterodorsal cirri reaching chaetiger 6 (Fig. 1A). Pharynx dissected; jaws light brown with 11 rounded teeth, extending to base (Fig. 1G). Maxillary ring: I = 12 cones in triangle, II = 32–31 cones in arc, III = 50 cones in an ellipse, IV = 34–32 cones in arc. Oral ring: V = 1 cone, VI = 4–4 pyramids in diamond, VII-VIII = 42 in two irregular rows, P-bars alternating with small pyra- mids in anterior-most row, similar-sized pyramids alternating in posterior-most row (in everted pharynx). Parapodial cirri pattern: Dorsal cirri longer than upper dorsal ligules throughout body; basally inserted on anterior region, displaced medially on midbody and poste- rior chaetigers. Ventral cirri shorter than neuropodial ligules throughout body, longer in few anterior chaetigers, basally inserted on anterior region, progressively distant throughout body. Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 21

Figure 1. Nereis oligohalina. Neotype female A–G (ECOSUR 0172); paraneotypes H–O (ECOSUR 0173). A Anterior end, dorsal view B Posterior end, dorsal view C Parapodium 2, anterior view D Para- podium 10, anterior view E Parapodium 46, anterior view F Parapodium 64, anterior view G Left jaw, dorsal view H Supra-acicular homogomph spiniger, parapodium 40 I Sub-acicular heterogomph spiniger, from same J Supra-acicular homogomph spiniger, from same K Notopodial homogomph falciger, from same L Supra-acicular heterogomph falciger, parapodium 28 M Sub-acicular heterogomph falciger, from same N Notopodial homogomph spiniger, parapodium 40 O Sub-acicular heterogomph spiniger, from same. Scale bars: 1 mm (A); 0.3 mm (B, G); 50 μm (C); 0.1 mm (D–F); 10 μm (H–M); 30 μm (N, O). 22 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015)

First two chaetigers uniramous, remaining ones biramous. Uniramous parapo- dia (Fig. 1C) with dorsal cirri basal, slightly longer than dorsal ligules. Dorsal ligules digitate; neuroacicular ligules subconical, subequal to dorsal ones; neuropodial ventral ligules digitate, slightly longer and basally twice as broad as dorsal ones. Ventral cirri slightly shorter than neuropodial ventral ligules; both dorsal and ventral cirri with similar width. In anterior parapodia (Fig. 1D), dorsal cirri medial, slightly longer than notopo- dial dorsal ligules, extending beyond their tips. Notopodial dorsal ligules subconical; subequal to ventral ones; notopodial ventral ligules globose, notoacicular papillae con- spicuous. Neuroacicular ligules globose, postchaetal lobes rounded, slightly shorter than neuroacicular ligules; neuropodial ventral ligules digitate, slightly shorter than neuroacicular ones. Ventral cirri shorter than neuropodial ligules; both dorsal and ven- tral cirri with similar width. In middle and posterior parapodia (Fig. 1E, F), dorsal cirri medial, slightly shorter than notopodial dorsal ligules. Notopodial dorsal and ventral ligules subequal, sub- conical, longer than wide, notoacicular papillae conspicuous in middle parapodia only. Neuroacicular ligules subconical, wider than long, postchaetal lobes rounded, about half as long as neuroacicular ligules; neuropodial ventral ligules digitate, 3–4 times longer than wide, medially attached to neuroacicular ligules, 2–3 times longer than them. Ventral cirri half as long or one-third as long as neuropodial ventral ligules; dorsal and ventral cirri with similar width. Glandular masses slightly conspicuous on ligules in posterior parapodia (Figs 1F, 6Q). In anterior and midbody parapodia notochaetae homogomph spinigers; neuro- chaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles. In posterior parapodia, notochaetae homogomph spinigers and falcigers; neurochaetae as in anterior parapo- dia. Chaetae decreasing rapidly in number toward posterior end. Notopodial homogomph spinigers pectinate (i.e. blade narrow with parallel teeth), teeth decreasing distally (Fig. 1N). Notopodial homogomph falcigers with sigmoidal blade, pectinate, distal tooth recurved, fused to blade (Fig. 1K). Neuropodial homogo- mph spinigers pectinate (Fig. 1H) or basally serrated (i.e. blade small with coarse teeth) (Fig. 1J), heterogomph spinigers pectinate (Fig. 1I); spinigers of similar size, teeth decreasing in size distally (Fig. 1N, O). Neuropodial heterogomph falcigers pectinate, distal tooth recurved, fused to blade, supra-acicular falcigers slightly broader than sub- acicular ones (Fig. 1L, M). Pygidium not modified; anal cirri cirriform, as long as last 5–6 chaetigers (Fig. 1B). Epitokes. Male fully transformed (ECOSUR-OH-P0761) complete; body taper- ing, 9 mm long, 0.9 mm wide, 57 chaetigers. Partially transformed male (ECOSUR P2827) complete; body tapering, 38 mm long, 2.1 mm wide, 57 chaetigers. Partially transformed female (ECOSUR P2827) complete; body tapering, 20 mm long, 1.6 mm width, 55 chaetigers. All with body yellowish with brown pigmentation present dorsally on first quarter of body, discoloring towards midbody chaetigers; faint lateral lines (Fig. 2A, D, E). Prostomium with pigmentation as in atokes, but less intense (Fig. 2B). Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 23

Prostomium longer than wide; antennae cirriform, slightly wider than those present in atokous female, as long as palps; eyes subequal, two (Fig. 2D) or three times (Fig. 2B) larger than antennal basal width, black, in trapezoidal arrangement (Fig. 2B, D). Peristomium twice as long as first chaetiger, slightly pigmented; tentacular cirri present; dorsal tentacular cirri longer than ventral ones, posterodorsal ones reaching to chaetiger 10 (Fig. 2B, D). Fully transformed male with pharynx everted, jaws amber with 10 teeth, inner edge toothed throughout. Maxillary ring: I = 8 cones in triangle, II = 30–32 cones in arc, III: 40 cones in rectangle, IV: 28–28 pointed cones in arc. Oral ring: V = 1 cone, VI: 4–4 pyramids in diamond, VII-VIII: 46 in two irregular rows, pyramids alternat- ing with small cones in most-anterior row, pyramids with similar size alternating in most-posterior row. Male body divided into two regions (Fig. 2A). Pre-natatory region includes chaetigers 1–16, natatory region from chaetiger 17 to end of body. Partially trans- formed female divided into two inconspicuous regions, parapodial lamellae visible from chaetiger 24. Parapodial cirri pattern: Anterior parapodia with dorsal cirri modified in chaetigers 1–7 in males, 1–5 in females; ventral cirri modified in chaetigers 1–5 in males, 1–4 in females; modification attenuated in females. Dorsal cirri subequal to upper dor- sal ligules in anterior chaetigers, slightly longer throughout body; basally inserted in anterior-most region, displaced medially toward end of body. Ventral cirri shorter than neuropodial ventral ligules in unmodified chaetigers, subequal in modified region; -ba sally inserted in anterior region, barely displacing ventrally throughout body. Chaetigers 1–2 uniramous (Fig. 2F); modified dorsal cirri basal, subpyriform (i.e. basally broad, medially broader, distally narrow), slightly longer than dorsal ligules. Dorsal and neuropodial ventral ligule subequal, subconical. Neuroacicular ligule sub- conical, much shorter than ventral one; postchaetal lobes rounded. Modified ventral cirri subpyriform, subequal to neuropodial ventral ligule; both dorsal and ventral cirri with similar width. Chaetigers 3–7 in males (Fig. 2G) and 3–5 in females, with slightly modified bi- ramous parapodia. Modified dorsal cirri medial, cattail-like (i.e. basal section broad, long; cirrostyle markedly narrower), slightly longer than notopodial dorsal ligules, ex- tending beyond them; basal section 2–3 longer than distal one. Notopodial dorsal ligules subconical, subequal to notopodial ventral one; notopodial ventral ligule sub- conical, notoacicular papilla conspicuous. Neuroacicular ligule subconical, postchaetal lobe rounded, shorter than neuroacicular ligule; neuropodial ventral ligule digitate, as long as neuroacicular one (Fig. 2G). Modified ventral cirri cattail-like, subulate in chaetigers 6–7 in males and 5 in females, shorter than neuropodial ventral ligules; modified dorsal and ventral cirrus with similar width. Chaetigers 8–16 in males and 6–23 in females with parapodial proportions as in atokes, but with more acute ligules (Fig. 2H). Remaining parapodia biramous, modified (Fig. 2I, J, L, M). Dorsal cirri medial, subulate, ventral margins sinuate in males only, longer than notopodial ligules; ba- 24 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015)

Figure 2. Nereis oligohalina. Non-type male A–C, F–J, N (ECOSUR–OH–P0761); non-type partially transformed male D, L–M (ECOSUR P2827); non-type partially transformed female E, K (ECOSUR P2827). A Whole specimen, dorsal view B Anterior end, dorsal view C Posterior end, dorsal view D, E An- terior ends, dorsal view F Parapodium 1, anterior view G Parapodium 6, anterior view H Parapodium 10, frontal view I Parapodium 18, anterior view J Parapodium 31, anterior view K Parapodium 10, anterior view L Parapodium 24, anterior view M Parapodium 52, anterior view N Sesquigomph natatory chaetae, parapodium 31. Scale bars: 1 mm (A, D–E); 0.5 mm (B); 0.2 mm (C); 0.1 mm (F–M); 10 μm (N). Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 25 sal lamellae small, smallest in females, progressively smaller toward posterior region. Notopodial dorsal and ventral ligules subequal, subconical, notoacicular papillae con- spicuous; ventral ligules with lamellae without projections along ventral margins. Neu- roacicular ligules subconical, subequal to notopodial ventral ones; postchaetal lobes developing into flabellate lamellae, increasing in size posteriorly, decreasing in posteri- or-most chaetigers, some with slight projections in dorsal or ventral edge; neuropodial ventral ligules digitate, medially attached to neuroacicular ones. Ventral cirri subulate, subequal to neuroacicular ligules, with two basal lamellae of different sizes; dorsal cirri wider than ventral ones. Prenatatory region with noto- and neurochaeta as in atokes, homogomph falcigers not observed, blades missing in most chaetae, number progressively reduced. In nata- tory region, noto- and neurochaetae sesquigomph chaetae with finely serrated, paddle- like blades (Fig. 2N). In fully transformed males pygidium with anus surrounded by rosette of papillae (Fig. 2C); anal cirri sinuous, as long as last 5–6 chaetigers (Fig. 2C). Transformation in females discrete; small lamellae on base of dorsal cirri, neuroacicu- lar ligules and ventral cirri (Fig. 2K); size of lamellae increasing toward posterior end and becoming inconspicuous in far posterior chaetigers; other features as in atokous female. Variation. The results of the analysis of body variation and paragnath numbers are summarized in Tables 1 and 2. The effect of fixation techniques on the shape of specimens have been recently evaluated by Oliveira et al. (2010) for Laeonereis acuta (Treadwell, 1923). The authors concluded that techniques of fixation can influence the shape and body proportions, especially if specimens were not previously relaxed, leading to erroneous identifications. In the case of N. oligohalina, fixation with 96% ethanol clearly affected the anterior portion of specimens with strong contraction of first segments, hence the tentacular cirri can reach more posterior chaetigers but without modifying their lengths, reaching up to chaetiger 14 (Table 1). Differences are not significant for10 L , W3 and W10 (P =

>0.05), but significant for L3 and rTC (P = <0.001, P = <0.0001). Maximum rTC for formalin specimens was preferred for the identification key because the data were less variable (Table 1). Nevertheless, these differences would not cause misidentification, because parapodial topology is not affected appreciably. Also pigmentation is very use- ful for recognizing the species. In the maxillary ring, area I showed the least variation (Fig. 6J), and in oral ring areas V and VI rarely vary in one paragnath only (Fig. 6K), such that these areas can be regarded as the most stable ones. The fingerprint-like pattern starts in chaetigers 10–11 (Fig. 6L); it is size-independent, but in smaller specimens this pattern is faint; however, as shown below it is absent in the two other species. The divergence between parapodial rami reported by Rioja (1946) is evident in posterior chaetigers but only in some specimens, forming a furrow (Fig. 6O). Glandular masses appear more visible and also on neuropodial ligules (Fig. 6Q); perhaps these glands fade in specimens fixed with formalin, such as the type material. In mature specimens, the natatory region starts in chaetiger 15–17 in males, and 24–25 in females. 26 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015)

Table 1. Ranges, means and standard deviation of some body measures in three Nereis species (TL: total length, L3 and L10: length at chaetiger 3 and 10, respectively; W3 and W10: width at chaetiger 3 and 10, respectively; nC: number of chaetigers; rTC: reach of largest tentacular cirrus; SD: standard deviation).

Nereis oligohalina TL L3 L10 W3 W10 nC rTC Minimum 10.0 1.1 1.6 0.6 0.6 55.0 4 Formalin Maximum 40.0 3.0 6.3 2.2 1.9 80.0 7 specimens (n = 23) Mean 24.8 2.3 3.8 1.6 1.4 66.9 5.7 SD 9.3 0.5 1.3 0.5 0.4 7.8 0.9 Minimum 9.0 0.6 1.5 0.7 0.6 57.0 5 Ethanol Maximum 36.0 3.0 6.5 2.2 2.0 78.0 14 specimens (n = 19) Mean 20.4 1.5 3.5 1.4 1.3 67.0 9.3 SD 9.3 0.6 1.3 0.5 0.5 8.8 2.5

N. garwoodi TL L3 L10 W3 W10 nC rTC Minimum 9.0 0.7 2.4 1.0 0.8 46.0 6 Formalin Maximum 36.0 3.0 6.9 2.0 2.1 96.0 13 specimens (n=33) Mean 22.1 1.8 3.8 1.5 1.5 72.7 9.3 SD 9.4 0.5 1.0 0.3 0.3 11.8 1.5

N. confusa sp. n. TL L3 L10 W3 W10 nC rTC Minimum 13.0 1.4 2.6 1.0 0.9 66.0 4 Formalin Maximum 35.0 2.5 5.8 1.9 1.7 89 7 specimens (n=20) Mean 24.5 1.8 4.0 1.5 1.3 78.6 5.3 SD 6.6 0.4 0.9 0.3 0.3 6.0 0.9

Remarks. Nereis oligohalina (Rioja, 1946) is considered as a widespread species and even amphiamerican, but this stems from taxonomic confusion and the lack of type material. Designation of a neotype for N. oligohalina was considered necessary because there are no type specimens and being a problematic species, there must be an objective definition for it (ICZN 1999, Art. 75.1). Consequently, a neotype has been selected, described and illustrated (ICZN 1999, Art. 75.3.3); this neotype fits the original descrip- tion by Rioja (1946) (ICZN 1999 Art. 75.3.5). Because Rioja did not designate holo- type, his material became syntypes and the species had two type localities (ICZN 1999 Art. 73.2.3, 76.1): Estero de Larios, Tecolutla, and El Cocal, Estero Casitas, both in Veracruz, Mexico. Although topotypic specimens from Estero Casitas are available, they are in poor condition, and therefore better specimens collected from nearby Actopan River were preferred once they were shown to conform to the same species (ICZN 1999, Recomm. 75A). The proposed neotype was collected in a similar environment and on oysters, as the original specimens (ICZN 1999, Art. 75.3.6); but the neotype locality is modified accordingly (ICZN 1999, Art. 76.3). The neotype was deposited in ECOSUR (ICZN 1999, Art. 75.3.7), including ‘paraneotypes’ and part of the additional material. Neanthes oligohalina Rioja, 1946 was correctly transferred to Nereis by Hartman (1951) because there are notopodial homogomph falcigers in posterior chaetigers. Hart- Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 27 man (1951, 1954) suggested that N. oligohalina was ‘inseparable’ from N. pelagica occiden- talis Hartman, 1945, but not synonymized. Salazar-Vallejo (1989) noted that although Rioja clearly recognized the presence of notopodial homogomph falcigers, he maintained the species under Neanthes, even in later publications (Rioja 1947, 1960, 1962). The first synonymy involving these species was made by Pettibone (1956); she con- sidered N. p. occidentalis different from N. pelagica and raised it to species level as Nereis (Nereis) occidentalis Hartman. Also, she regarded Neanthes oligohalina as a junior syno- nym of N. occidentalis being regarded as a variety. The detailed description provided by Pettibone allowed us to recognize differences in comparison to the Laguna Madre, Texas specimens. She recognized slight but important differences among these variants, mainly in paragnath number in areas V and VI; adding the relative size of neuropodial ligules in middle and posterior chaetigers. In the same work, she determined that Nereis largoensis Treadwell, 1931 was a junior synonym of N. pelagica, and that other material identified asN. largoensis based upon material examined by Treadwell corresponds to N. occidentalis (Pettibone 1956). We follow, however, González-Escalante and Salazar- Vallejo (2003), who concluded that these three species are not synonyms. Nereis oligohalina differs from N. occidentalis in some diagnostic features. In N. oligohalina there are 8–15 paragnaths on area I and 1–2 on area V, whereas in N. oc- cidentalis there are 2–3 paragnaths on area I and no paragnaths on area V. Further, in N. oligohalina neuropodial ventral ligules are 2–3 times longer than neuroacicular ligules, but in N. occidentalis neuropodial ventral ligules from posterior chaetigers are shorter than neuroacicular ones. Regarding chaetae, and as an additional difference, in N. oligohalina the notopodial homogomph falciger has the distal tooth less developed than in N. occidentalis. Another synonymy was made by Day (1973), who regarded N. pelagica occidenta- lis Hartman and N. occidentalis fide McCloskey as junior synonyms of Nereis falsa de Quatrefages, 1865; however, in the list of synonyms of these two species he did not include Pettibone (1956), therefore N. oligohalina was not considered by him. Nereis falsa, has a rather complex or confusing delineation; according to Fauvel (1923) it differs from similar species by having different numbers of paragnaths in areas I and V, and different proportions in parapodial ligules. In fact, N. falsa is another species regarded as widely distributed and requires a critical revision and we could anticipate a restriction of its distribution to the Mediterranean region because its type locality is the Black Sea. On the other hand, N. oligohalina has been recorded along American Atlantic coasts from northeastern Brazil, chiefly in ecological (community assemblages on Spartina alternifloraand mangroves), or population studies (secondary production and population dynamics), as well as part of taxonomic or genetic studies (Amaral et al. 2012). A detailed record was made by Santos and Lana (2003); unfortunately, a com- mentary and one plate was based upon specimens collected in Todos Los Santos Bay, and indicated that their material agrees with the original description regarding neu- ropodial ventral ligules in posterior parapodia, and the feature was less developed in specimens from other localities but were regarded as the same because the prostomial 28 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015) pigmentation and paragnath number remained constant (Santos and Lana 2003). Lana et al. (2006) however, considered their previous record as a probable misidentification requiring a revision. Liñero-Arana and Díaz (2007) recorded N. oligohalina from Ven- ezuela in La Restinga Lagoon, Margarita Island, associated with Crassostrea rhizophorae (Guilding), and recognized that their specimens resembled Brazilian ones and differed from those described by Rioja, mainly in parapodial morphology. These two publica- tions pointed out the need for a revisionary work, recognizing N. oligohalina as valid species, but that their specimens were probably not the same as those described from Mexico because of parapodial features; at least the specimens from South America were regarded as a different species that should be clarified elsewhere. Records of N. oligohalina from the Eastern Tropical Pacific (Berkeley and Berkeley 1958, 1960) belong to a new species described below. Habitat. The species is associated with red mangrove Rhizophora mangle and with oysters. It has been found in Gulf of Mexico estuaries, including Tecolutla, Casitas- Nautla and Actopan (these estuaries have sand bars in their respective mouth rivers), and from coastal lagoons such as Mandinga and La Mancha. These systems have direct connection with the sea, some with seasonal closure of their mouths, with polyhaline to mesohaline waters (Lara-Domínguez et al. 2011). The neotype and associated specimens were found in Crassostrea virginica (Gme- lin) reef, in the Actopan river mouth. The specimens studied by Rioja (1946) from the Tecolutla estuary were found on mangrove roots covered by cirripedians, whereas specimens from the Casitas-Nautla estuary were collected between oysters (possibly C. virginica) and mytilids as Ischadium recurvum (Rafinesque) (reported as Mytilus recur- vatus (sic) by Rioja); as Rioja indicated, N. oligohalina specimens cohabit with an Alitta species in the Actopan river mouth. This species has been reported as N. occidentalis Hartman, together with Polydora websteri Hartman, as epifauna of C. virginica (Ruiz- Guerrero and López-Portillo Guzmán 2006), and on Rhizophora mangle roots (Ruiz and López-Portillo 2014), from La Mancha. Distribution. Restricted to the southwestern Gulf of Mexico.

Nereis garwoodi González-Escalante & Salazar-Vallejo, 2003 Figures 3, 4, 6B, E, M, I, R

Nereis garwoodi González-Escalante and Salazar-Vallejo 2003: 156–160, figs 1a–k, 2a–h.

Type material. Quintana Roo, Mexico. Lectotype ECOSUR 0065 and paralecto- types ECOSUR 0066 (7), Chetumal (18°29'38.88"N, 88°17'22.89"W), Chetumal Bay, 1 m depth, in calcareous sedimentary rocks, mixed bottom, September 24 1999, Coll. L.E. González-Escalante, S.I. Salazar-Vallejo. Additional material. Chetumal Bay, Quintana Roo, Mexico. ECOSUR P2829 (3), Alacranes (18°34'28.51"N, 88°14'24.21"W), May 1 1999, Coll. LEGE, SISV. Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 29

ECOSUR P2830 (14), Chetumal, May 3 1999, Coll. LEGE, SISV. ECOSUR P2831 (29), Chetumal, 21 May 1999, Coll. LEGE, SISV. ECOSUR P2832 (17), Alacranes (18°34'28.51"N, 88°14'24.21"W), May 21 1999, Coll. LEGE, SISV. ECOSUR P2833 (2), Luis Echeverría (18°39'04’’N 88°12'07’’W), May 21 1999, Coll. LEGE, SISV. ECOSUR P2834 (28), Chetumal (18°29'38.88"N, 88°17'22.89"W), June 30 1999, Coll. LEGE, SISV. ECOSUR P2835 (9), Chetumal (18°29'38.88"N, 88°17'22.89"W), August 27 1999, Coll. LEGE, SISV. Type locality. Chetumal Bay, Mexico, Caribbean Sea, on rocks in mixed bottoms, 1 m depth. Description. Lectotype complete (ECOSUR 0065), atokous female, damaged with incisions at level of chaetigers 1, 15, and 30. Body tapering, 36 mm long, 1.7 mm wide, 95 chaetigers. Body pale, pigmentation faint, brown rectangle present dorsally on middle of anterior chaetigers, striated, discoloring toward end of body, lateral pale lines in anterior chaetigers only, oocytes present. Prostomium with brown pigment along inner margins of palps, two lines extending from antennae toward anterior pair of eyes separated by a longitudinal pale area, and two oval lateral patches; peristomium slightly pigmented, pale lines present (Fig. 3A). Prostomium 1.5 times longer than wide; antenna cirriform, slightly passing palps; eyes subequal, black, in a rectangle (Fig. 3A). Peristomium twice longer than first chaetiger; tentacular cirri with short ceratophores, left cirri broken; dorsal longer than ventral ones, posterodorsal ones reaching chaetiger 12 (Fig. 3A). Pharynx everted, jaws pale brown with 13 rounded teeth, extending to base (Fig. 3I). Maxillary ring: I = 15 pyramids in diamond, II = 31–31 pyramids and cones in arc, III = 44 cones in an ellipse, IV = 35–35 pyramids in arc (Fig. 3A). Oral ring: V = 1 cone, VI = 4–4 pyramids in diamond, VII-VIII = 42 in two irregular rows, P-bars alternating with small pyramids in anterior-most row, pyramids and cones with similar size alternating in posterior-most row. Parapodial cirri pattern: Dorsal cirri longer than upper dorsal ligules throughout body; basally inserted on anterior region, displaced medially in midbody region, be- comes subdistal in posterior chaetigers. Ventral cirri longer than neuropodial ligules in a few anterior chaetigers, progressively reduced throughout body; basally inserted on anterior region, barely migrating ventrally throughout body. First two chaetigers uniramous, remaining ones biramous. In uniramous parapodia (Fig. 3D), dorsal cirri basal, slightly longer than notopodial ligules. Dorsal and neuro- podial ventral ligules subequal, digitate, three times longer than neuroacicular ligules; neuroacicular ligules subconical, postchaetal lobes rounded. Ventral cirri subequal to neuropodial ventral ligules; both dorsal and ventral cirri with similar length and width. In anterior parapodia (Fig. 3E), dorsal cirri medial, longer than notopodial ligules, extending beyond them. Notopodial dorsal ligules subconical, slightly longer than ventral ones; notopodial ventral ligules subconical, slightly longer than neuroacicular ligules, notoacicular papillae very conspicuous. Neuroacicular ligules subconical, sub- equal to ventral ones, postchaetal lobes rounded, slightly shorter than neuroacicular ligules; neuropodial ventral ligules digitate, basally attached to neuroacicular ligules. 30 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015)

Ventral cirri shorter than neuropodial ventral ligules; dorsal cirri twice wider than ventral ones. In middle parapodia (Fig. 3F), dorsal cirri medial, as long as notopodial ligules, extending beyond them. Notopodial dorsal and ventral ligules subequal, subconical, notoacicular papillae inconspicuous. Neuroacicular ligules subconical, slightly shorter than remaining ones, postchaetal lobes rounded, shorter than neuroacicular ligules; neuropodial ventral ligules digitate, basally attached to neuroacicular ones. Ventral cirri half as long as neuropodial ligule; both dorsal and ventral cirri with similar width. In posterior parapodia (Figs 3G, H), dorsal cirri medial, slightly longer than no- topodial ligule. Notopodial dorsal ligules become broad, longer than ventral ones; notopodial ventral ligules become large, 2–3 times longer than neuroacicular ligules, notoacicular papilla inconspicuous. Neuroacicular ligules subconical, half as long as neuropodial ventral ones, postchaetal lobes inconspicuous; neuropodial ventral ligules digitate, basally attached to neuroacicular ligules. Ventral cirri up to half as long as neuropodial ligule; dorsal and ventral cirri with similar width. Glandular masses con- spicuous on notopodial ligules. In anterior and midbody parapodia, notochaetae homogomph spinigers; neuro- chaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles. In posterior parapodia, notochaetae homogomph spinigers and falcigers; neurochaetae as in anterior parapo- dia. Number of chaetae decreasing toward posterior end. Notopodial homogomph spinigers pectinate, teeth decreasing in size distally (Fig. 3O). Notopodial homogomph falcigers with sigmoid blade, pectinate, distal tooth incurved, fused to blade (Fig. 3C). Neuropodial homogomph spinigers pectinate or basally serrate (Fig. 3J), heterogomph spinigers pectinate (Fig. 3P); both with teeth decreasing in size distally. Neuropodial heterogomph falcigers pectinate, distal tooth incurved, fused to blade, supra-acicular slightly broader than sub-acicular (Fig. 3L– N); supra-acicular falcigers narrow in midbody chaetigers, becoming broad posteriorly (Fig. 3L, M). Pygidium without modification; anal cirri cirriform, as long as last 4–5 segments (Fig. 3B). Epitokes. Paralectotype fully transformed male (ECOSUR 0066) complete, body tapering, 9 mm long, 1 mm wide, 46 chaetigers; paralectotype partially transformed fe- male (ECOSUR 0066) incomplete, body tapering, 16 mm long, 2 mm wide, 63 chaeti- gers; fully transformed female (ECOSUR P0066) complete, body tapering, 12 mm long, 1.6 mm wide, 63 chaetigers. All with body yellowish with brown pigmentation present dorsally on first quarter of body, discoloring toward midbody chaetigers (Fig. 4E, F). Prostomium and peristomium with pigmentation similar to atokes (Fig. 4A, B). Prostomium longer than wide; antennae cirriform, as long as palps; eyes black, subequal, in a rectangle, three times larger than antennal basal width (Fig. 4A, B). Peristomium twice length of first chaetiger; tentacular cirri with short ceratophores, dorsal tentacular cirri longer than ventral ones, posterodorsal reaching to chaetiger 9 in male, 13 in female (Fig. 4A, B). Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 31

Male with pharynx everted, jaws amber with 9 teeth. Maxillary ring: I = 6 pointed cones in triangle, II = 19–20 pointed cones in arc, III = 28 pointed cones in rectangle, IV = 22–19 pointed cones in arc. Oral ring: V = 1 pointed cone, VI = 4–3 pyramids in diamond, VII-VIII = 42 in two irregular rows, P-bars alternating with small pyramids in most-anterior row, pyramids alternating with cones with similar size in most-posterior. Male body divided into two regions (Fig. 4E); pre-natatory region includes chaeti- gers 1–16, natatory region from chaetiger 17 to end of body. Fully transformed female body with two regions; pre-natatory includes chaetigers 1–26, natatory region from chaetiger 27 to end of body (Fig. 4F). Partially transformed female divided in two inconspicuous regions, lamellae start in chaetiger 25. Parapodial cirri pattern: Anterior parapodia with dorsal cirri modified in chaetigers 1–7 in males, 1–5 in females; ventral cirri modified in chaetigers 1–5 in males, 1–4 in females. Dorsal cirri subequal to upper dorsal ligules in anterior chaetigers, become subequal throughout body; basally inserted in most-anterior region, displaced medially throughout body. Ventral cirri shorter than neuropodial ligules in unmodified chaeti- gers, subequal in modified region; basally inserted in anterior region, barely migrating ventrally throughout body. Chaetigers 1–2 with uniramous (Fig. 4G, L), modified dorsal cirri basal, subpy- riform in males, cattail-like in females, subequal to dorsal ligules. Dorsal and neuro- podial ventral ligules subequal, subconical, twice longer than wide in male, 1.5 times longer than wide in female. Neuropodial lobe subconical, shorter than dorsal ligules; postchaetal lobes rounded. Modified ventral cirri shorter than neuropodial ventral lig- ules; dorsal and ventral cirri subequal with similar width and length. Chaetigers 3–7 in males (Fig. 4H) and 3–5 in females (Fig. 4M) with slightly modified biramous parapodia. Modified dorsal cirri medial, cattail-like, slightly longer than notopodial ligules, extending beyond them; broader section as long as narrower one. Notopodial dorsal ligules subconical, slightly longer than notopodial ventral lig- ules; notopodial ventral ligules subconical, twice longer than neuroacicular ligules, notoacicular papillae conspicuous. Neuroacicular ligules subconical, postchaetal lobes rounded; neuropodial ventral ligules digitate, longer than neuroacicular ones. Modi- fied ventral cirri cattail-like, subulate in chaetigers 6–7 in males and 5 in females, shorter than neuropodial ventral ligules. Parapodial proportions as in atokous from chaetigers 8–16 in male and 6–26 in female (Fig. 4I, N). Remaining parapodia modified (Fig. 4J, K, O, P). Dorsal cirri medial, subulate, ventral margins sinuate in males only, subequal to notopodial ventral ligules; basal la- mellae large in males, small in females, increasing size toward posterior chaetigers and decreasing in most-posterior ones. Notopodial dorsal ligules subconical, longer than ventral ones in male, subequal in female; notopodial ventral ligules subconical, developing a large ventral lamella in males only, with a round projection. Neuroacicular ligules subconical, shorter than notopodial ventral ones; postchaetal lobes developing into flabellate lamellae with a round projection in dorsal edge in males, small lamellae in females, progressively in- 32 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015)

Figure 3. Nereis garwoodi. Lectotype female A–B, D–P (ECOSUR 0065); paralectotype C (ECOSUR 0066). A Anterior end, dorsal view B Posterior end, dorsal view C Notopodial homogomph falciger, para- podium 75 D Parapodium 2, anterior view E Parapodium 9, anterior view F Parapodium 28 G Parapodium 56, anterior view H Parapodium 80, anterior view I Left jaw, dorsal view J Supra-acicular homogomph spiniger, parapodium 80 K Sub-acicular heterogomph spiniger, from same L Supra-acicular heterogomph falciger, parapodium 56 M Supra-acicular heterogomph falciger, parapodium 80 N Sub-acicular hetero- gomph falciger, from same O Supra-acicular heterogomph spiniger, from same P Sub-acicular heterogomph spiniger, from same. Scale bars: 1 mm (A–B, I); 50 μm (C, J–N); 0.1 mm (D–H); 0.3 mm (O–P). Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 33

Figure 4. Nereis garwoodi. Paralectotype male A, C, E, G–K (ECOSUR 0066); paralectotype female B, D, F, L–P (ECOSUR 0066). A, B Anterior ends, dorsal view. C, D Posterior ends, dorsal view E, F Whole specimens, dorsal view G Parapodium 2, anterior view H Parapodium 6, anterior view I Para- podium 10, anterior view J Parapodium 18, anterior view K Parapodium 36, anterior view L Parapodium 2, anterior view M Parapodium 5, anterior view N Parapodium 10, anterior view O Parapodium 26, anterior view P Parapodium 36, anterior view. Scale bars: A–D = 0.5 mm; E–F = 1 mm; G–P = 0.1 mm. 34 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015) creasing in size and decreasing in far posterior segments; neuropodial ventral ligules digitate, basally attached to neuroacicular ones. Ventral cirri subulate, slightly longer than neuroacicular ligules, with two basal lamellae of different sizes; dorsal cirri wider than ventral ones. Prenatatory region with noto- and neurochaetae as in atokes, homogomph falcigers not observed. In natatory region, notochaetae and neurochaetae sesquigomph chaetae with finely serrated, paddle-like blades; atokous chaetae not completely replaced in specimens of either sex, homogomph falcigers observed in male (Fig. 4J, K, O, P). Pygidium modified, anus surrounded by rosette of papillae in male, unmodified in female (Fig. 4C, D); anal cirri cirriform, as long as last 5–6 segments (Fig. 4C, D). Variation. The results of the analysis of body variation and analysis of paragnath numbers are summarized in Tables 1 and 2. The width measurements reported here differ from the original description because chaetiger width without parapodia was used, instead of measuring them including parapodia. The arrangement and number of paragnaths have similar ranges as those reported for N. oligohalina (Fig. 6B, E; Table 2); however, area I has a larger range, and the arrangement is somewhat variable, often in a triangle (Fig. 6I). Also, paragnaths are more robust than in N. oligohalina and N. confusa sp. n. Regarding pigmentation, the striated rectangle seen in lectotype is more conspicu- ous in some specimens (Fig. 6M), which is also sometimes present in N. oligohalina, but the color is much more intense whereas the fingerprint-like pattern of the latter species was not observed. In mature specimens, the natatory region starts from chaeti- ger 17 only in males and 25–27 in females, which differs from the original description (22 in males and 21 in females). One specimen presented a duplicated ventral cirrus, but it was regarded as abnormal (Fig. 6R). Remarks. González-Escalante and Salazar-Vallejo (2003) indicated that they had six atokes and two epitokes as syntypes. Five atokous syntypes were expected to be sent to four foreign museums, but were never dispatched. Further, these syntypes were not formally deposited and labeled, and parts of the descriptions and illustrations are too imprecise to enable separation of N. garwoodi from N. oligohalina. In an attempt to redefine the species, a lectotype has been selected (ICZN 1999, Art. 74.1) to avoid future confusion; although the syntype series has better preserved specimens, the lectotype matches the original description and illustration, and was therefore preferred (ICZN 1999, Recomm. 74B). In order to ensure their validity, the term has been introduced in the material section and in the description (ICZN 1999, Art. 74.7.1, 74.7.3), and the lectotype has been described, illustrated and their data updated for its recognition (ICZN 1999, Art. 74.7.2, Recomm. 74C, 74E); the remaining syntypes are regarded as paralectotypes (ICZN 1999, Recomm. 74F). These specimens are deposited in ECOSUR. Nereis garwoodi is closely allied with N. oligohalina, but they differ in some features in both atokous and epitokous forms, and in their habitats. In atokes, N. garwoodi never shows the dark brown coloration nor the fingerprint-like pattern found in N. oligohalina. The ranges of paragnath numbers of both species overlap and therefore are Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 35

Table 2. Ranges, means and standard deviations (SD) in number of paragnaths in three Nereis species (r: right, l: left).

Pharynx areas N. oligohalina (n=27) I II-r II-l III IV-r IV-l V VI-r VI-l VII–VIII Minimum 8 25 25 32 22 22 1 4 3 40 Maximum 15 38 37 71 45 39 2 4 4 47 Mean 11.7 31.0 30.4 49.1 31.5 31.4 1.04 4.0 3.9 43.9 SD 2.3 2.9 3.1 9.2 4.9 3.6 0.2 0.00 0.2 2.0 N. garwoodi (n=29) I II-r II-l III IV-r IV-l V VI-r VI-l VII–VIII Minimum 4 19 20 28 22 19 1 4 3 42 Maximum 19 42 40 59 43 38 1 4 4 46 Mean 10.7 30.3 30.5 44.6 31.0 30.1 1.0 4.0 3.9 44.0 SD 3.4 5.3 5.7 6.4 5.3 4.5 0.00 0.00 0.3 1.3 N. confusa sp. n. (n=30) I II-r II-l III IV-r IV-l V VI-r VI-l VII–VIII Minimum 4 30 28 39 34 35 1 3 3 42 Maximum 11 36 42 62 58 54 1 5 6 45 Mean 7.3 32.3 34.2 51.2 45.6 45.1 1 3.6 4.0 42.9 SD 2.1 2.3 3.9 8.1 8.9 6.6 0 0.7 0.9 1.1 not useful to separate them, and the relative length of tentacular cirri would be useful if fixation method is the same (Table 1). In N. garwoodi, both dorsal and neuropodial ventral ligules are twice as long as neuroacicular ligules in uniramous chaetigers, whereas in N. oligohalina these are sube- qual and slightly longer, respectively. Also, in N. garwoodi the neuropodial postchaetal lobes are visible in the anterior and midbody only, whereas in N. oligohalina they are visible throughout body. Further, N. garwoodi has notopodial ventral ligules twice as long as neuropodial ventral ones in posterior chaetigers, whereas in N. oligohalina these ligules are sub- equal to each other; further, in N. oligohalina neuropodial ventral ligules are medi- ally attached in posterior chaetigers, whereas in N. garwoodi they are basally attached throughout body. Moreover, in N. garwoodi notopodial homogomph falcigers have more teeth and they are narrower than in N. oligohalina; also, in N. garwoodi the blades of supra-acicular heterogomph falcigers become broader and shorter in poste- rior chaetigers, but this modification is not present in N. oligohalina. In epitokes, N. garwoodi has modified, cattail-like dorsal cirri present in bira- mous chaetigers with the basal sections as long as distal ones, whereas in N. oligo- halina basal sections are longer. Also, in general N. garwoodi have better developed lamellae in natatory chaetigers than N. oligohalina, especially the basal lamellae of the dorsal cirri, the lamellae of both notopodial ventral and neuroacicular ligules. Moreover, epitokal transformation is more pronounced in females of N. garwoodi than in females of N. oligohalina. On the other hand, N. garwoodi is associated with calcareous rocks, while N. oligohalina is associated with reef-building bivalves and the mangrove R. mangle. 36 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015)

Habitat. Chetumal Bay is a semi-closed, dynamic system linked to the Caribbean Sea by several freshwater tributaries, having a salinity gradient ranging 7–18 practical salinity units (psu) (Carrillo et al. 2009). The species bores into calcareous sedimen- tary rocks, building mucous tubes, and has been regarded as a sedentary herbivore (González-Escalante and Salazar-Vallejo 2003); to obtain the specimens, rocks must be broken. Although the Bay has extensive zones of mangroves, N. garwoodi has never been found among them. Distribution. Apparently restricted to Chetumal Bay. González-Escalante and Salazar-Vallejo (2003) report a gradient of decreasing abundance from the southern to the northern regions of the bay, probably related to organic matter load.

Nereis confusa sp. n. http://zoobank.org/5048FF4A-0F6A-4B03-BCF5-352F41EDBC39 Figures 5, 6C, F–H, N, P

Nereis (Neanthes) oligohalina Berkeley and Berkeley 1958: 402 (non Rioja, 1946). Nereis oligohalina Berkeley and Berkeley 1960: 359 (non Rioja, 1946).

Type material. Gulf of California, Baja California Sur. Holotype ECOSUR 0174 and paratypes ECOSUR 0175 (5), Bahía de La Paz (24°08'38.68"N, 110°20'44.40"W), March 1 2004, 70 m from shore, on wrinkled pen shell Pinna rugosa, sponges, PVC tube, and filamentous green algae, Coll. M.A. Tovar-Hernández, P. Salazar-Silva. Additional material. Gulf of California, Baja California. ECOSUR P2836 (16), Bahía de Los Ángeles (28°58'6.72"N, 113°32'43.24"W), Gulf of California, March 17 1985, on Atrina maura, Coll. E. Aguirre, C. Garza. Gulf of California, Sinaloa. ECOSUR P2837 (10), Estero el Yugo (23°18'8.30"N, 106°29'0.53"W), Mazatlán, February 24 2004, 50 cm depth, fine sediment, on filamentous green algae on man- grove roots, Coll. S. Rendón-Rodríguez, Nuri M., M.A. Tovar-Hernández, P. Salazar- Silva. Baja California Sur. ECOSUR P2838 (32), same data as holotype. ECOSUR P2839 (2), Bahía La Paz (24°12'6.51"N, 110°17'59.26"W), Gulf of California, March 2 2004, 1 m depth, on basalt rocks, sponges and algae, Coll. M.A. Tovar-Hernández, P. Salazar-Silva. Etymology. The specific name (L.confusa : confused, perplexed, troubled) indi- cates an earlier problematic delineation of the species; it is a noun in apposition. Description. Holotype complete (ECOSUR 0174), atokous female. Body taper- ing, 34 mm long, 1.7 mm wide, 81 chaetigers, inmature. Body yellowish, reddish brown pigmentation present dorsally on first quarter of body as three spots pattern, two lateral ones, and the other less pigmented, middorsal, forming discontinuous transverse bands up to chaetiger 10, remaining segments pale; lateral pale lines in ante- rior chaetigers only (Fig. 5A). Prostomium with pigmentation brown on inner margins of palps and with two oval patches (Fig. 5A); peristomium dorsally pigmented, varie- gated (Fig. 5A), with very short pale lines on posterior margin. Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 37

Figure 5. Nereis confusa sp. n. Holotype A–P (ECOSUR 0174). A Anterior end, dorsal view B Posterior end, dorsal view C Notopodial homogomph falciger, parapodium 72 D Parapodium 2, anterior view E Parapodium 11, anterior view F Parapodium 28, anterior view G Parapodium 51, anterior view H Pa- rapodium 72, anterior view I Left jaw, dorsal view J Supra-acicular homogomph spiniger, parapodium 51 K Sub-acicular heterogomph spiniger from same L Sub-acicular heterogomph falciger from same M Supra-acicular heterogomph falciger, parapodium 72 N Supra-acicular heterogomph falciger from same O Supra-acicular homogomph spiniger from parapodium 51 P Sub-acicular heterogomph spiniger from same. Scale bars: 1 mm (A–B, I); 50 μm (C, J–N); 0.1 mm (D–H); 0.3 mm (O–P). 38 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015)

Prostomium longer than wide; antennae cirriform, extending beyond palps; eyes subequal, black, in a rectangle (Fig. 5A). Peristomium twice longer than first chaetiger; tentacular cirri with short ceratophores, dorsal cirri longer than ventral ones, postero- dorsal ones reaching to chaetiger 5 (Fig. 5A). Pharynx dissected, jaws with 8 teeth, restricted to anteromedial edge, light brown (Fig. 5I). Maxillary ring: I = 5 cones in rectangle, II = 30–30 cones in arc, III = 49 cones in an ellipse, IV = 41–36 cones and some merged, in sigmoidal. Oral ring: V = 1 cone, VI = 5–5 cones in round, VII-VIII: 42 in two irregular rows, P-bars and small cones alternating in most anterior row, pyramids and small cones in most-posterior row. Parapodial cirri pattern: Dorsal cirri longer than upper dorsal ligules throughout body; basally inserted on anterior region, displaced medially on midbody and poste- rior regions. Ventral cirri as long as neuropodial ligules in a few anterior chaetigers, progressively reduced throughout body; basally inserted in anterior region, migrating ventrally throughout body. First two chaetigers uniramous, remaining biramous. In uniramous parapodia (Fig. 5D), dorsal cirri basal, slightly longer than dorsal ligules. Dorsal ligules subconi- cal; neuroacicular ligules subconical, subequal to dorsal ligules; neuropodial ventral ligules digitate, shorter than neuroacicular ligules. Ventral cirri subequal than neuro- podial ligule; dorsal cirri slightly wider than ventral ones. In anterior parapodia (Fig. 5E), dorsal cirri medial, as long as notopodial dorsal ligules, extending beyond them. Notopodial dorsal ligules subconical; notopodial ven- tral ligules globose, subequal to dorsal ones, notoacicular papillae conspicuous. Neu- roacicular ligules subconical, postchaetal lobe rounded, subequal than neuroacicular ligules; neuropodial ventral ligules digitate, shorter than neuroacicular ones. Ventral cirri shorter than neuropodial ligule; dorsal cirri slightly wider than ventral ones. In midbody and posterior parapodia (Figs 5F–H), dorsal cirri medial, subequal to notopodial dorsal ligules. Notopodial dorsal and ventral ligules subequal, slightly enlarged in posterior parapodia, subconical, longer than wide, notoacicular papillae conspicuous in middle parapodia. Neuroacicular ligules subconical, postchaetal lobes rounded, shorter than notopodial ligules; neuropodial ventral ligules digitate, slightly shorter than neuroacicular ones. Ventral cirri shorter than neuropodial ventral ligules; dorsal cirri slightly wider than ventral ones (Fig. 5F–H). In anterior and midbody parapodia, notochaetae homogomph spinigers; neuro- chaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles. In posterior parapodia, notochaetae homogomph spinigers and falcigers; neurochaetae as in anterior parapodia. Notopodial homogomph spinigers pectinate, teeth decreasing in size distally. No- topodial homogomph falcigers pectinate, 9 teeth, distal tooth stout, incurved, fused to blade (Fig. 5C). Neuropodial homogomph spinigers basally serrate (Fig. 5J), het- erogomph spinigers pectinate or serrate (Fig. 5K); both with teeth decreasing in size distally. Neuropodial heterogomph falcigers pectinate, distal tooth incurved, fused to blade, very conspicuous (Fig. 5L–N); in posterior parapodia with short (Fig. 5M) or long (Fig. 5N) blades in both fascicles, missing in most chaetae. Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 39

Pygidium with broad margin, anus crenulated; anal cirri cirriform, as long as last 3–4 segments (Fig. 5B). Variation. The results of the analysis of body variation and paragnath numbers are summarized in Table 1 and 2. The arrangement and number of paragnaths is similar to that in N. oligohalina and N. garwoodi, but in N. confusa sp. n. paragnaths in areas III and IV are more numerous than in the other two species (Fig. 6C, F; Table 2), and cones have rounded tips; further, N. confusa sp. n. is the only species of the three with merged paragnaths. All specimens examined show the same dorsal spotted pigmentation, but in some, especially the largest specimen, the middorsal spot disappears and only two discon- tinuous lines are visible along chaetigers 1–10; fingerprint-like or striated patterns were not observed. In mature specimens, the transformation starts in chaetiger 18 in males, 22 in females as previously noticed by Berkeley and Berkeley (1958, 1960). Remarks. Nereis confusa sp. n. has been recorded as N. oligohalina; however, there are several differences between these two species. In N. confusa sp. n. the spotted pig- mentation pattern extends up to chaetigers 10–14, and the jaws have 8 teeth restricted to the distal part of its inner edge, whereas in N. oligohalina the pale areas are replaced by fingerprint-like patterns from chaetiger 11, and its jaws have 11 teeth along its inner edge. Further, in N. confusa sp. n., both notopodial ligules and neuroacicular ligules are subequal to, or slightly longer than, neuropodial ventral ligules in midbody and posterior parapodia, whereas in N. oligohalina, they are twice as long as the neuroac- icular ligules in midbody and posterior parapodia. On the other hand, N. confusa sp. n. has falcigers with broad blades, whereas in N. oligohalina they are narrower. Further- more, in N. confusa sp. n. distal tooth of notopodial homogomph falciger is short and well developed, whereas in N. oligohalina it is longer and weakly developed. The first records forN. confusa sp. n. (as N. oligohalina) from the Mexican Pacific were made by Berkeley and Berkeley; first, they reported Nereis (Neanthes) oligohalina males from Hipolito Bay (Berkeley and Berkeley 1958), and males and females from La Paz (Berkeley and Berkeley 1960). They argued that prostomium, anterior chaeti- gers and arrangement of paragnaths all matched Rioja’s descriptions. In their brief comments, they indicated the start of the modified region or first epitokous parapo- dium (17 in males, 22 in females), and a spotted pattern of pigmentation in males. These features match with N. confusa sp. n. rather than N. oligohalina. Rioja (1962) cited Berkeley and Berkeley (1958), and recorded N. confusa sp. n. (as N. oligohalina) from El Mogote, Ensenada de La Paz, Baja California Sur; he men- tioned a slight discrepancy in number of paragnaths in area I, and that paragnaths in the periphery of area III were larger than the rest of the group forming a borderline; also, he regarded the glandular parapodial masses as typical. Despite the fact that he did not provide more information, we regard his specimens as belonging to N. con- fusa sp. n. Other Mexican Pacific reports ofN. pelagica occidentalis by Bastida-Zavala (1993, 1995) from nearby localities might also be conspecific. Dean (2001) reported N. oligohalina from Pacific Costa Rican coasts, noticed the problems in the taxonomic history of the species, and regarded it as different from N. oc- 40 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015) cidentalis. According to his description Costa Rican specimens differ fromN. oligohalina in the number of paragnaths, mainly in areas I, III and IV. Also, in his specimens the long- est tentacular cirri reached chaetiger 3, and the notopodial dorsal and neuropodial ven- tral ligules were subequal to or shorter than notopodial ventral and neuroacicular ligules throughout body, whereas in N. oligohalina the longest tentacular cirri reaches chaetiger 7, and their ligules are larger in midbody and posterior chaetigers. Likewise, Costa Rican specimens resemble N. confusa sp. n. and probably belong to the same species and this might also include the record from Cocos Island (Dean et al. 2012). Nevertheless, these records cannot be assigned to N. oligohalina unequivocally until specimens are evaluated. Habitat. Holotype found on wrinkled penshell Pinna rugosa Sowerby, 1835, sponges and filamentous green algae; other specimens were found in sponges and green algae near the type locality. Specimens from Bahía de Los Ángeles were associated with another penshell, Atrina maura (Sowerby, 1835), and specimens from Estero El Yugo were found on filamentous green algae on mangrove roots. Bastida-Zavala (1995) found specimens on corals. Distribution. Gulf of California, Eastern Pacific coasts of Mexico. Probably ex- tends to Costa Rica, in shallow water. Reproduction patterns and dispersal in Nereis species studied. Some authors have emphasized the utility of reproductive patterns in taxonomy for species discrimination in closely related taxa (Smith 1958, Clark 1977). Also, strategies for larvae survival would be relevant, especially because they determine larval transport and its dispersal potential. An interesting strategy is when only males form epotikes and females remain atokous or are barely modified, as in Alitta virens (Sars, 1835) and Websterinereis glauca (Claparède, 1870). In A. virens, males form epitokes but females remain atokous or present very slight changes, also females spawn into or in the opening of their burrows (Bass and Brafield 1972). In W. glauca, female transformation is reduced, and females produce mucous tubes to deposit and incubate their eggs, while males can swarm (Pettibone 1971). Early studies considered that N. garwoodi presented a similar reproductive mode as A. virens or W. glauca, because the paralectotype female of N. garwoodi has a slight transformation; however, after further revision of additional material, a fully transformed female was found. We have no further details about its capacity for building mucous tubes or if females do not emerge to the water column or if gametes are retained in tubes. In the material available of N. oligohalina, no completely transformed females were encountered, but perhaps with further sampling efforts they may appear. An- other important consideration is the reduced size of males compared to females, even in the same sample; this has been noted for H. diversicolor (Bartels-Hardege and Zeeck 1990). In this case, N. oligohalina males can swarm while females remain inside cavi- ties, as in A. virens and W. glauca; the fact that there are many large females filled with oocytes points toward this direction. Reproduction modes in estuarine species play a crucial role in their dispersal be- cause the formation of planktonic larvae can determine their distribution range. Bilton et al. (2002) proposed two life-cycle models for estuarine species that have larvae: ex- port vs retention strategies. In the former, the adults respond to physical or biological Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 41

Figure 6. Variation of Nereis species studied. N. oligohalina A, D, J–L, O, Q from ECOSUR–OH– P0760. N. garwoodi B, E from paralectotype ECOSUR 0066; I, M, R from ECOSUR P2834. N. confusa sp. n. C, F–H, N, P from ECOSUR P2838. A–C Pharynges everted, dorsal view D–F Pharynges everted, ventral view G Pharynx everted, anterior view, showing merged paragnaths (arrows) H Close-up of merged paragnaths on area IV I–J Variations on area I K Variations on areas V and VI L Fingerprint- like pattern, dorsal view M–N Pigmentation patterns on anterior ends, dorsal view O Parapodial furrow in posterior end, lateral view (arrow indicates start) P Specimen on tube, dorsal view Q Parapodium from posterior chaetiger, anterior view R Ventral cirrus duplicated, chaetiger 5, anterior view. Scale bars: 0.5 mm (A–F, L); 1 mm (M–P); 0.1 mm (Q–R). 42 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015) factors by releasing gametes or larvae in the lower estuary; larvae are driven out from the estuary, mainly by tides, and later juveniles or adults return to the estuary. In the retention strategy, adults release their gametes or larvae in the upper estuarine areas, then they undergo early development in middle estuary; there, larvae have vertical mi- grations during circadian ebb-flood tidal regimes such that larvae are not exported but retained within the estuary (Bilton et al. 2002). Based on the above, we hypothesize that N. oligohalina has an export strategy; it could disperse thanks to surface currents running parallel to the continental margin, and this would explain its presence along Gulf of Mexico estuaries. Similarly, distribu- tion of N. confusa sp. n., with mainly marine habitats, could be due to current patterns along the Gulf of California. However, sometimes the distribution patterns cannot be explained by currents and tidal dynamics. For example, De Jesús-Flores et al. (2015) determined that Laeonereis nota (Treadwell, 1941), described for Galveston, Texas, is also present in Chetumal Bay; L. nota spawn into their burrows, limiting their dispersal by currents. The expla- nation for this discontinuous distribution lies in passive dispersal through migratory birds, because they use nereidids as food (De Jesús-Flores et al. 2015). Similarly N. garwoodi could have a classical retention strategy, but a wide, fragmented distribution caused by migratory birds. Further considerations. The present study demonstrates the need to encourage re- descriptions of closely related and widely distributed species and, should it be necessary, the establishment of new species if there are conspicuous morphological differences. Further, clarifying species delineation and distribution are urgent because they are es- sential for biogeography and phylogenetics. Reuscher and Shirley (2014) studied the dis- tribution patterns of polychaetes from the Gulf of Mexico; a recent species list was used (Fauchald and Solís-Weiss 2009) and current taxonomy verified in (Read and Fauchald 2015). They found that among all species recorded, 32% were cosmopolitan, 15% Pan-American and 9% Pan-Atlantic (N. oligohalina was regarded as Pan-American, which is incorrect as shown above). They concluded: “Most polychaete families are in need of global and regional revisions. Clear species boundaries have to be established by means of taxonomic research based on morphology and genetic analyses. Geographical ranges of spe- cies should be revised in order to eliminate false conclusions about distributions of species.”

Key to species of Nereis from the Grand Caribbean Region (Modified from González-Escalante and Salazar-Vallejo 2003)

1 Area V without paragnaths...... 2 – Area V with paragnaths...... 9 2(1) Parapodial ligules long, slender; ceratophores distinct...... N. caymanensis Fauchald, 1977 Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 43

– Parapodial ligules not enlarged; ceratophores indistinct...... 3 3(2) Notopodia with prechaetal lobes ...... N. goajirana Augener, 19331 – Notopodia without prechaetal lobes...... 4 4(3) Dorsal cirri shorter than notopodial ligules; notopodial homogomph falcigers with oval blades...... N. grayi Pettibone, 1956 – Dorsal cirri subequal or longer than notopodial ligules; notopodial homogo- mph falcigers with falcate blades...... 5 5(4) Paragnaths on area I absent; notopodial homogomph falcigers with bifid blade...... N. panamensis Fauchald, 1977 – Paragnaths on area I present, notopodial homogomph falcigers with entire blade...... 6 6(5) Area VII-VIII with few paragnaths, usually 5–7...... 7 – Area VII-VIII with numerous paragnaths, more than 40...... 8 7(6) Longest tentacular cirri reaching chaetiger 7; area VI usually with 9 parag- naths...... N. riisei (Grube & Ørsted in Grube, 1858)2 – Longest tentacular cirri reaching chaetiger 4; area VI usually with 3 parag- naths...... N. allenae Pettibone, 19563 8(6) Notopodial homogomph falciger with numerous teeth, distal tooth recurved, one quarter of blade embedded in shaft...... N. occidentalis Hartman, 1945 – Notopodial homogomph falciger with few teeth, without distal tooth, one- half of blade embedded in shaft...... N. pelagica Linnaeus, 17584 9(1) Notopodial homogomph falcigers with cutting edge smooth...... N. largoensis Treadwell, 19315 – Notopodial homogomph falcigers with cutting edge denticulate...... 10 10(9) Dorsal pigmentation as a striated pattern, usually along first 10 chaetigers; tentacular cirri reaching beyond chaetiger 7...... 11 – Dorsal pigmentation different; tentacular cirri reaching up to chaetiger .7.... 12 11(10) Area I with paragnaths in an oval; longest tentacular cirri reaching up to chaetiger 9...... N. rigida Grube & Ørsted in Grube, 18586 – Area I with paragnaths in a triangle; longest tentacular cirri reaching up to chaetiger 12...... N. garwoodi González-Escalante & Salazar-Vallejo, 2003 12(10) Dorsal pigmentation with fingerprint-like pattern from chaetigers 10–11; noto- podial ligules enlarged in posterior chaetigers...... N. oligohalina (Rioja, 1946) – Dorsal pigmentation with spotted pattern; notopodial ligules with similar proportions along body...... N. confusa sp. n.7

1 This species has not been found recently (Báez and Ardila 2003). 2 It has a doubtful amphiamerican status and there is no recent redescription. Here, descriptions by Renaud (1956) with specimens from Miami, and by Hoagland (1919) for N. glandulata from Porto Rico, were used. Hartman (1956) considered N. glandulata as a junior synonym of N. arroyensis Treadwell, 1901, but they differ in number of paragnaths on area VII-VIII. 44 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015)

3 De León-González et al. (1999) detailed the arrangement of paragnath number of N. allenae Pettibone, 1956. 4 After Chambers and Garwood (1992). 5 In the original description of N. largoensis as N. brevicirrata, Treadwell (1929) re- ported area V without paragnaths and homogomph falciger with lenticular blades. However, Hartman (1956) found one paragnath on area V and homogomph fal- ciger six times longer than wide. 6 Augener (1906) introduced the species name for the Caribbean fauna; however, Grube (1858) used a single specimen from Pacific Costa Rica (Puntarenas). Here the original description by Grube was used, and it was included as a contrast for N. confusa sp. n. 7 Distributed in the Mexican Pacific, often confused withN . oligohalina.

Additional comments. Several species recorded from the Grand Caribbean are questionable, such as N. falcaria (Willey, 1905), N. jacksoni Kinberg, 1866, N. victori- ana Augener, 1918, N. falsa de Quatrefages, 1865, N. callaona (Grube, 1857) and N. lamellosa Ehlers, 1868; consequently, they were not included in the key because their type localities are distant and different form the tropical American conditions. OnlyN . pelagica was included in order to contrast it with N. occidentalis, but its records from the Grand Caribbean Sea might belong to a different species.

Acknowledgements

We thank María A. Tovar-Hernández, Jesús. Á. de León-González (UANL), and Tulio F. Villalobos-Guerrero, for making available some of the material of N. oligohalina than made this study possible. We also thank Astrid E. Te-Gómez and Concepción Licona-Rosado help during fieldwork, and Luis F. Carrera-Parra by offering helpful suggestions about type-materials. This publication is part of the requirements for ob- taining a M.S. degree at El Colegio de la Frontera Sur, Unidad Chetumal (ECOSUR), and supplied by a fellowship from CONACyT.

References

Amaral ACZ, Nallin SAH, Steiner TM, Forroni T de O, Filho DG (2012) Catálogo das espé- cies de Annelida Polychaeta do Brasil. http://www.ib.unicamp.br/museu_zoologia/sites/ www.ib.unicamp.br.museu_zoologia/files/Cat%C3%A1logo_Polychaeta_Brasil_Amaral_ et_al_2013_1a.pdf [2013-03-06] Augener H (1906) Reports on the results of dredging, under the supervision of Alexander Agas- siz, in the Gulf of Mexico and the Caribbean Sea, and on the east coast of the United States, 1877 to 1880, by the U.S.S. Coast Survey Steamer “Blake,” Lieut. Commander C.D. Sigsbe, Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 45

U.S.N., and the commander J.R. Bartlett, U.S.N., commanding. XLII. Westindische Poly- chaeten. Bulletin of the Museum of Comparative Zoology 43(4): 91–196. Báez DP, Ardila NE (2003) Poliquetos (Annelida: Polychaeta) del Mar Caribe colombiano. Biota Colombiana 4: 89–109. Bakken T (2007) Revision of Pseudonereis (Polychaeta, Nereididae). Zoological Journal of the Linnean Society 150: 145–176. doi: 10.1111/j.1096-3642.2007.00289.x Bakken T, Glasby CJ, Wilson RS (2009) A review of paragnath morphology in Nereididae (Polychaeta). Zoosymposia 2: 305–316. Bakken T, Wilson RS (2005) Phylogeny of nereidids (Polychaeta, Nereididae) with paragnaths. Zoologica Scripta 34: 507–547. doi: 10.1111/j.1463-6409.2005.00200.x Bartels-Hardege H, Zeeck E (1990) Reproductive behavior of Nereis diversicolor (Annelida: Polychaeta). Marine Biology 106: 409–412. doi: 10.1007/BF01344320 Bass NR, Brafield AE (1972) The life-cycle of the polychaete Nereis virens. Journal of the Marine Biological Association of the United Kingdom 52: 701–726. doi: 10.1017/ S0025315400021664 Bastida-Zavala JR (1993) Taxonomía y Composición Biogeográfica de los Poliquetos (Annelida: Polychaeta) de la Bahía de la Paz, B.C.S., México. Revista de Investigación Científica UABCS 4: 11–39. Bastida-Zavala JR (1995) Poliquetos (Annelida: Polychaeta) del Arrecife Coralino de Cabo Pulmo-Los Frailes, B.C.S., México. Revista de Zoología 6: 9–29. Berkeley E, Berkeley C (1958) Some notes on a collection of polychaeta from the Northeast Pacific south of latitude 32° N. Canadian Journal of Zoology 36: 399–407. doi: 10.1139/ z58-034 Berkeley E, Berkeley C (1960) Notes on some Polychaeta from the West Coast of Mexico, Pan- ama, and California. Canadian Journal of Zoology 38: 357–362. doi: 10.1139/z60-040 Bilton DT, Paula J, Bishop JDD (2002) Dispersal, genetic differentiation and speciation in estuarine organisms. Estuarine, Coastal and Shelf Science 55: 937–952. doi: 10.1006/ ecss.2002.1037 Boilly-Marer Y (1972) Sur la signification et le déterminisme des caracteres sexuels somatiques chez les Nereidae (Annélides polychètes). Archives de Zoologie Expérimentale et Générale 113: 369–393. Carrillo L, Palacios-Hernández E, Ramírez AM, Morales-Vela JB (2009) Características hi- drometeorológicas y batimétricas. In: Espinosa-Ávalos J, Islebe GA, Hernández-Arana HA (Eds) El Sistema ecológico de la bahía de Chetumal/Corozal: costa occidental del Mar Caribe. ECOSUR, Chetumal, México, 12–20. Chambers S, Garwood P (1992) Polychaetes from Scottish Waters. A Guide to Identification, 3. Family Nereidae. National Museum, Edinburgh, 65 pp. Charrier H (1920) Recherches sur la “Nereis fucata” Sav. Actes de la Société Linnéenne de Bordeaux 72: 5–144. Clark RB (1977) Reproduction, speciation and polychaete taxonomy. In: Reish DJ, Fauchald K (Eds) Essays on the Polychaetous in Memory of Dr. Olga Hartman. Allan Hancock Foundation, University of South California, Los Angeles, 477–501. 46 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015)

Darbyshire T (2014) Intertidal and nearshore Nereididae (Annelida) of the Falkland Islands, southwestern Atlantic, including a new species of Gymnonereis. Zookeys 427: 75–108. doi: 10.3897/zookeys.427.7296 Day JH (1973) New Polychaeta from Beaufort, with a key to all species recorded from North Carolina. NOAA Technical Report NMFS CIRC-375: 1–153. doi: 10.5962/bhl.title.62852 De Jesús-Flores C, Salazar-González SA, Salazar-Vallejo SI (2015) Morphological distinction between estuarine nereidid polychaetes: Laeonereis culveri (Webster, 1879) and L. nota (Treadwell, 1941). Revista de Biología Tropical (sent for evaluation: Mar. 2015). De León-González JÁ, Solís-Weiss V, Ochoa-Rivera V (1999) Nereidids (Polychaeta) from the Caribbean Sea and adjacent coral islands of the southern Gulf of Mexico. Proceedings of the Biological Society of Washington 112: 667–681. Dean HK (2001) Some Nereididae (Annelida: Polychaeta) from the Pacific Coast of Costa Rica. Revista de Biología Tropical 40: 37–67. Dean HK, Sibaja-Cordero JA, Cortés J (2012) Polychaetes (Annelida: Polychaeta) of Cocos Is- land National Park, Pacific Costa Rica. Pacific Science 66: 347–386. doi: 10.2984/66.3.8 Evenhuis NL (2008) A compendium of zoological type nomenclature: A reference source. Bishop Museum Technical Report 41: 1–25. Fauchald K (1977) The polychaete worms: Definitions and keys to the orders, families and genera. Natural History Museum of Los Angeles County, Science Series 28: 1–190. Fauchald K, Solís-Weiss V (2009) Polychaeta (Annelida) of the Gulf of Mexico. In: Felder DL, Camp DK (Eds) Gulf of Mexico origin, waters and Biota: volume I, biodiversity. Texas A&M University Press, College Station, 751–788. Fauvel P (1923) Polychètes Errantes. Faune de France 5: 1–488. Glasby CJ (1999) The Namanereidinae (Polychaeta: Nereididae). 1. Taxonomy and Phylo­ geny. Records of the Australian Museum, Supplement 25: 1–129. doi: 10.3853/j.0812- 7387.25.1999.1354 Glasby CJ, Timm T, Muir AI (2009) Catalogue of non-marine Polychaeta (Annelida) of the World. Zootaxa 2070: 1–52. Glasby CJ, Wilson RS, Bakken T (2011) Redescription of the Indo-Pacific polychaeteNeanthes pachychaeta (Fauvel, 1918) n. comb. (Annelida, Phyllodocida, Nereididae) and its syno- nyms. Zoosystema 33(3): 361–375. doi: 10.5252/z2011n3a5 Glasby CJ, Wei N-WV, Gibb KS (2013) Cryptic species of Nereididae (Annelida: Polychaeta) on Australian coral reefs. Invertebrate Systematics 27: 245–264. doi: 10.1071/IS12031 González-Escalante LE, Salazar-Vallejo SI (2003) A new estuarine species, Nereis garwoodi (Poly- chaeta: Nereididae), from Bahía Chetumal, Mexican Caribbean coast. Revista de Biología Tropical 51: 155–164. Grube AE (1858) Annulata örstediana. Enumeratio Annulatorum, quae in itinere per Indiam occidentalem et Americam centralem annis 1845-1848 suscepto legit cl. A.S. Örsted, adjectis speciebus nonnullis a cl. H. Kröyero in itinere ad Americam meridionalem collectis, (Fortsæ). Videnskabelige Meddelelser fra Dansk Naturhistorisk Forening for 1857: 158–186. Hartman O (1948) The marine annelids erected by Kinberg with notes on some others types in the Swedish State Museum. Arkiv för Zoologi 42A(1): 1–156. Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 47

Hartman O (1951) The littoral marine annelids of the Gulf of Mexico. Publications of the Institute of Marine Science 2: 7–124. Hartman O (1954) Polychaetous annelids of the Gulf of Mexico. In: Galtsoff PS (Ed.) Gulf of Mexico. Its Origin, Waters, and Marine Life. Fishery Bulletin of the Fish and Wildlife Service. Volume 55, 413–417. Hartman O (1956) Polychaetous annelids erected by Treadwell, 1891 to 1948, together with a brief chronology. Bulletin of the American Museum of Natural History 109: 239–310. Herpin R (1925) Recherches biologiques sur la reproduction et le développement de quelques annélides polychètes. Bulletin de la Société des Sciences Naturelles de l’Ouest de la France, série 4ème, 5: 1–250. Hoagland RA (1919) Polychaetous annelids from Porto Rico, the Florida Keys, and Bermuda. Bulletin of the American Museum of Natural History 41: 571–591. ICZN (1999) International Code of Zoological Nomenclature, 4th ed. International Trust for Zoological Nomenclature (The Natural History Museum), London. Lana P da C, Santos CSG, Garraffoni ARS, Oliveira VM, Radashevsky V (2006) Checklist of Polychaete Species from Paraná State (Southern Brazil). Check List 2: 30–63. doi: 10.15560/2.3.30 Lara-Domínguez AL, Contreras-Espinosa F, Castañeda-López O, Barba-Macías E, Pérez- Hernández MA (2011) Lagunas costeras y estuarios. In: La biodiversidad en Veracruz: Estudio de caso. Comisión para el Conocimiento y Uso de la Biodiversidad (CONABIO), Gobierno del Estado de Veracruz, Universidad Veracruzana (UV), Instituto de Ecología A.C. (INECOL), 301–317. Linnaeus C (1758) Systema Naturae per Regna Tria Naturae, secundum Classes, Ordines, Genera, Species, cum Characteribus, Differentiis, Synonymis, Locis. 10th Ed., Laurentii Salvii, Holmie (Stockholm), 823 pp. Liñero-Arana I, Díaz-Díaz Ó (2007) Nuevas adiciones de Nereididae (Annelida: Polychaeta) para las costas de Venezuela. Boletín del Instituto Oceanográfico de Venezuela 46: 149–159. Oliveira VM, Santos CSG, Lana P da C, Camargo MG (2010) Morphological variations caused by fixation techniques may lead to taxonomic confusion inLaeonereis (Polychaeta: Nerei- didae). Zoologia (Curitiba) 27: 146–150. doi: 10.1590/S1984-46702010000100022 Pettibone MH (1956) Some polychaete worms of the families Hesionidae, Syllidae, and Nereidae from the east coast of North America, West Indies, and Gulf of Mexico. Journal of the Washing- ton Academy of Sciences 46: 281–294. Pettibone MH (1971) Revision of some species referred to Leptonereis, Nicon, and Laeonereis (Polychaeta: Nereididae). Smithsonian contribution to Zoology 104: 1–53. doi: 10.5479/ si.00810282.104 Read GB (1980) A new species of Nereis (Polychaeta: Nereididae) from Wellington, New Zealand, Rock Shores. Journal of the Royal Society of New Zealand 10: 185–193. doi: 10.1080/03036758.1980.10427174 Read GB (2007) Taxonomy of sympatric New Zealand species of Platynereis, with description of three new species additional to P. australis (Schmarda) (Annelida: Polychaeta: Nereididae). Zootaxa 1558: 1–28. 48 Víctor M. Conde-Vela & Sergio I. Salazar-Vallejo / ZooKeys 518: 15–49 (2015)

Read G, Fauchald K (Eds) (2015) World Polychaeta database. http://www.marinespecies.org/ polychaeta [2015-05-26] Renaud JC (1956) A report of some polychaetous annelids from the Miami-Bimini area. Amer- ican Museum Novitates 1812: 1–40. Reuscher MG, Shirley TC (2014) Diversity, distribution, and zoogeography of benthic poly- chaetes in the Gulf of Mexico. Marine Biodiversity, 1–14. doi: 10.1007/s12526-014-0222-7 Rioja E (1946) Estudios anelidológicos. 15. Nereidos de agua salobre de los esteros del litoral del Golfo de México. Anales del Instituto de Biología, Universidad Nacional Autónoma de México 17: 205–214. Rioja E (1947) Estudios anelidológicos. 19. Observaciones sobre algunos nereidos de las costas de México. Anales del Instituto de Biología, Universidad Nacional Autónoma de México 28: 247–266. Rioja E (1960) Estudios anelidológicos. 24. Adiciones a la fauna de anélidos poliquetos de las costas orientales de México. Anales del Instituto de Biología, Universidad Nacional Autónoma de México 31: 289–316. Rioja E (1962) Estudios anelidológicos. 26. Algunos anélidos poliquetos de las costas del Pacífico de México. Anales del Instituto de Biología, Universidad Nacional Autónoma de México 33: 131–229. Ruiz M, López-Portillo J (2014) Variación espacio-temporal de la comunidad de macroin- vertebrados epibiontes en las raíces del mangle rojo Rhizophora mangle (Rhizophoraceae) en la laguna costera de La Mancha, Veracruz, México. Revista de Biología Tropical 62: 1309–1330. doi: 10.15517/rbt.v62i4.12582 Ruiz-Guerrero M, López-Portillo Guzmán J (2006) Los invertebrados. In: Moreno-Casasola P (Ed.) Entornos Veracruzanos: La Costa de La Mancha. Instituto de Ecología, Xalapa, 341–361. Salazar-Vallejo SI (1989) Enrique Rioja y su contribución al estudio de los poliquetos (Anelida: Polychaeta) en México. Brenesia 30: 39–65. Salazar-Vallejo SI (2011) Revision of Piromis Kinberg, 1867 and Pycnoderma Grube, 1877 (Polychaeta: Flabelligeridae). Zootaxa 2819: 1–50. Santos CSG, Lana P da C (2003) Nereididae (Polychaeta) da costa Nordeste do Brasil, 3 Gêneros Ceratonereis e Nereis. Iheringia, Série Zoologia 93: 5–22. Santos CSG, Pleijel F, Lana P da C, Rouse GW (2005) Phylogenetic relationships within Nereididae (Annelida: Polychaeta). Invertebrate Systematics 19: 557–576. doi: 10.1071/ IS05001 Sendall K, Salazar-Vallejo SI (2013) Revision of Sternaspis Otto, 1821 (Polychaeta, Sternaspidae). Zookeys 286: 1–74. doi: 10.3897/zookeys.286.4438 Smith RI (1958) On reproductive pattern as a specific characteristic among nereid polychaetes. Systematic Zoology 7: 60–73. doi: 10.2307/2411793 Treadwell AL (1929) New species of polychaetous annelids in the collections of the American Museum of Natural History, from Porto Rico, Florida, Lower California, and British So- maliland. American Museum Novitates 392: 1–13. Villalobos-Guerrero TF, Carrera-Parra LF (2015) Redescription of Alitta succinea (Leuckart, 1847) and reinstatement of A. acutifolia (Ehlers, 1901) n. comb. based upon morphologi- Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi... 49

cal and molecular data (Polychaeta: Nereididae). Zootaxa 3919: 157–178. doi: 10.11646/ zootaxa.3919.1.7 Yáñez-Rivera B, Carrera-Parra LF (2012) Reestablishment of Notopygos megalops McIntosh, de- scription of N. caribea sp. n. from the Greater Caribbean and barcoding of “amphiamerican” Notopygos species (Annelida, Amphinomidae). Zookeys 223: 69–84. doi: 10.3897/zook- eys.223.3561

Capítulo 2

Nereididae (Polychaeta: Phyllodocida) from the Grand Caribbean and other related

regions

43

Nereididae (Polychaeta: Phyllodocida) from the Grand Caribbean and other related regions

VÍCTOR MANUEL CONDE-VELA1

1El Colegio de la Frontera Sur, CONACyT, Departamento de Sistemática y Ecología Acuática, Chetumal, México

Email: [email protected]

Abstract

The nereidid polychaetes from the Grand Caribbean region have been widely studied but their current taxonomy is far from being clarified. In order to improve the current taxonomy of nereidids from the Grand Caribbean region, a regional faunistic study was performed. The materials examined include mainly specimens collected during the University of Miami Deep Sea Expeditions along the Grand Caribbean, together with the study of types or non-type specimens deposited in ECOSUR, UANL, USNM and LACM-AHF, covering almost all regions from the Grand Caribbean. One hundred eight species of nereidids have been recorded from the Grand Caribbean region; 44 species were considered valid and 38 are questionable records; 7 species were reinstated but 4 of them informally, based upon original literature only. Thirteen Nereis species require further revisions to ensure their status as valid species, five species have special features that would imply transferring them to other genera, and 11 new species were described. Modifications in the current parapodial terminology, a classification for first two chaetigers and a standardized method to measure parapodial structures are also proposed.

Introduction

The Grand Caribbean region comprises the Caribbean Sea, the Gulf of Mexico, and the Bermuda, with Florida as the northern border and the northern coasts of Brazil as the southern one (Salazar- Vallejo 2000). Based upon a recent approach, the Grand Caribbean region is part of the Tropical Atlantic realm, which is comprised by 3 provinces and 16 ecoregions (Spalding et al. 2007).

44

Nowadays, some efforts have stressed the under estimation of species richness in all marine taxa in diverse areas of the Grand Caribbean. In addition, as part of a diminishing trend, regional forms have not been described as new species, or they were misidentified, due to previously prevailing perspective that of cosmopolitanism was very common among polychaetes and, more recently, because they could be regarded as exotic ones, but also the lack of taxonomic experts in all marine groups (Miloslavich et al. 2010). All these explain why Reuscher & Shirley (2014) have highlighted the problem in benthic polychaetes from Gulf of Mexico, and made a call to revisionary works for the region.

The family Nereididae de Blainville, 1818, is well represented in Tropical America, with 25 genera and 104 species (Salazar-Vallejo & Londoño-Mesa 2004; Salazar-Vallejo 1996; de León- González 2009). Members of this family are distinguished by having an eversible pharynx, often ornamented with paragnaths or papillae, and by well-developed parapodia. Currently, three subfamilies are recognized: Nereidinae de Blainville, 1818, Namanereidinae Hartman, 1959, and Gymnonereidinae Banse, 1977. Previously, the first two were regarded as monophyletic, and the third one paraphyletic (Glasby 1991), but recent approaches demonstrate that there are several conflicts in the current definitions of subfamilies and genera.

On the other hand, nereidids seem to be a well-known family due to the great amount of literature available; however, many taxonomic problems are still present. Some efforts have been performed trying to solve main problems, focused in a phylogenetic overview, as the ontogenetic origin of body structures and terminology for naming them, more accurate definitions for subfamilies and genera, and use of new tools for recognizing species as genetic and reproductive units (Sato & Nakashima 2003, Ackermann et al. 2005, Bakken & Wilson 2005, Santos et al. 2005, among many others).

In the Grand Caribbean, several names have been proposed for regional species; however, they have been regarded as synonyms of other species described from very different and distant regions such as the North Sea or the Mediterranean, Australia, or Indonesia. An special variant of these species is the called amphiamerican ones, but their existence has been regarded as questionable (Salazar-Vallejo 2000), and their rejection has been documented concluding that

45

Pacific species differ from Atlantic ones (Carrera-Parra & Salazar-Vallejo 2011, Yáñez-Rivera & Carrera-Parra 2012, Conde-Vela & Salazar-Vallejo 2015, Villalobos-Guerrero & Carrera-Parra 2015).

The main goals of this contribution are the following. First, to improve the current terminology, some emendations are proposed based upon definitions instead of common usage, and new terms are introduced to avoid future confusions, based upon in the patterns observed in different genera and subfamilies. Second, a method to perform parapodial measurements is proposed, trying to standardize the way they are measured since several works do not indicate it. Third, to perform an exhaustive revision of regional and introduced names, with the reinstatement of several regional species and the description of some new.

Material and methods

Type, topotype and non-type material was examined to perform the descriptions. Specimens are deposited in the following collections: Natural History Museum of Los Angeles County, Allan Hancock Foundation, Los Angeles, California, United States (LACM-AHF), National Museum of Natural History, Smithsonian Institution, Washington D.C., United States (USNM), the Polychaetological Collection of the Universidad Autónoma de Nuevo León (UANL), and in the Colección de Referencia, El Colegio de la Frontera Sur, Chetumal (ECOSUR). Some type specimens will be deposited in the Marine Invertebrates Museum, Rosenstiel School of Marine and Atmospheric Science, University of Miami, Florida, United States (UMML), and in USNM. Some species from other regions are included because 1) it has been regarded as a synonym of some species of the region, or 2) the name was based upon specimens from different regions but introduced into regional checklists. In both cases, the species were described with topotypic material or, if none available, some comments were added explaining the doubts about their synonymy or wide distribution. Body measurements include long, wide (without parapodia) and number of chaetigers. Occasionally methyl green staining was employed for improving the contrast, mainly in small specimens. Some parapodia along the body were removed and mounted in semi-permanent slides, to be examined under compound microscope. Whenever possible, one of the first two

46 chaetigers was illustrated and described. When not everted, pharynx was dissected to observe pharyngeal ornamentation, and in some cases, a jaw was mounted and photographed. Photographs were made using a digital camera provided by an adapter. Photographs were cleaned digitally and, in many cases, a set of photographs was combined to improve the composition of the images. Terminology used in this study and the procedures to estimate parapodial proportions are explained in morphology section below. If present, descriptions of pigmentation pattern were made because they are useful for recognition of species (Glasby et al. 2013, Conde-Vela & Salazar-Vallejo 2015). In the legends of figures, trying to make them as brief as possible, some indications were omitted such as whole specimens, anterior or posterior ends, jaws and pharynges, are all depicted in dorsal view, unless otherwise noted. Right parapodia from selected chaetigers were removed and are shown in anterior view; therefore in legends only appear the number of chaetiger. Further, ‘notopodial’ was abbreviated as ‘Not.’; and for neuropodia ‘supra-acicular’ and ‘sub-acicular’ were abbreviated as ‘Sup.’ and ‘Sub.’, respectively.

Morphology

Head, structures and origin. The head of nereidids is composed by two main structures, the prostomium and the peristomium, currently named achaetous ring. The prostomium bears the antennae, palps and eyes, and their relative lengths and shapes are diagnostic features. The antennae generally are two and are absent in Micronereis and Namanereis malaitae only (Glasby 1999; Paxton 1983); here the ratio between the length of antennae respect to the length of prostomium was used. The eyes generally are subequal in size, but sometimes one pair (anterior or posterior) is longer than the other one and in some cases, the eyes are absent. The two palps have little variation regarding palpophores, but palpostyles can be conical or spherical; the relative length of antennae in relation to palps was reported in this study, but since palpostyles can contract, it must be used cautiously. Respect to their origin, the prostomium and its structures arise from various cell lineages (Ackermann et al. 2005). Traditionally in nereidids and related genera, the term peristomium refers to the achaetous ring surrounding the mouth, although the fate of larval peristomium in adult nereidids is unclear yet. It

47 is a well-known fact that the cirri of the achaetous ring have distinct origins. The anterior-most cirri are supposedly peristomial-derived, i.e., with pre-segmental origin and formed first during ontogeny, and consequently called peristomial cirri; the posterior-most ones are derived from the first true segment, fusing with the peristomium during cephalization, and therefore called tentacular cirri (Fauchald 1977c; Wilson 1932). Wilson (1932) pointed out that, the posterior pair tentacular cirri became from lobes and not from parapodial cirri in Nereis pelagica ((Wilson 1932). On the contrary, Rouse & Pleijel (2001) and Santos et al. (2005) regarded all as segmentally-derived, using the name anterior cirri introduced in previous studies. Later, Ackermann et al. (2005) clearly demonstrated in larvae of Platynereis dumerilii that ‘peristomial’ cirri arise from a metameric cell lineage (1D blastomere), instead of a peristomial origins, and Winchell et al. (2010) found the same in , a species with direct development; both studies regarded as doubtful the presence of a remnant of larval peristomium in adult nereidids. After, Steinmetz et al. (2011) corroborated the metameric origin of first pair of tentacular cirri in P. dumerilii, and Kulakova et al. (2007) demonstrated that the same Hox gen (Hox1) is expressed in first three segments, including the tentacular cirri. Therefore, the evidence is conclusive in that all tentacular cirri in nereidids are segmentally- derived, therefore the term is adequate because it implies the metameric origin of cirri. On the other hand, the current terms achaetous or apodous ring is also used in cirratulids, capitellids, dorvilleids, lumbrinerids, among others families (e.g. Arias & Carrera-Parra 2014; Bakken et al. 2010; Haddad et al. 2013; Paxton & Åkesson 2007), without any resemblance to the structure found in nereidids. The term tentacular segment as defined by Mikkelsen & Virnstein (1982) seems to be more adequate for naming this ring, and it has been used previously in nereidids and other phyllodocideans, but in aphroditiforms there can be chaetae and even an acicular lobe (Fitzhugh & Wolf 1990; Imajima 2003; Pettibone 1971, 1976, 1989). To avoid more confusion about it, the term tentacular ring is used herein to ring-like, cephalized segments bearing tentacular cirri only, and its restricted use in nereidids is proposed. Furthermore, only two genus, Namanereis and Lycastonereis, have three pairs of tentacular cirri instead of four, so their origin has little controversy in the relationship within nereidids, and is relevant in homology issues at family and upper levels.

48

Pharynx patterns. Hylleberg et al. (1986) focused on the importance of the precise arrangement and number of paragnaths for separating species; they made schematic representations by using symbols for arrangements and rectangles representing pharyngeal areas. However, this scheme was not followed because it failed to provide information about the real position on pharynx surfaces. The relative position of paragnaths and papillae on pharyngeal rings is conservative among all genera (Bakken & Wilson 2005), but there are a few exceptions on the oral ring. In some cases, the areas V and VI are poorly defined as in Pseudonereis pseudonoodti (Fauchald, 1977) and Perinereis cariacoensis (Liñero-Arana, 1983). In other cases, there is a continuous band of paragnaths along the oral ring, covering areas V, VI and VII-VIII, sometimes forming a broad belt, as in Neanthes acuminata and related species. Further, areas VII-VIII have other variants. They can have a broad belt, not fused with areas V and VI as in N. pelagica; or there can be one, two or more well defined parallel series, often with conical or pointed-bars paragnaths alternating in position and forming a jagged line, as in some Nereis and Pseudonereis species. The rows in areas VII-VIII may vary in number but paragnath arrangement is generally transverse, linear, and to describe how many rows may cause confusion. For example, in some Pseudonereis species, pointed bars paragnaths or P-bars alternate with conical paragnaths along areas VII-VIII; there is no consensus about how many rows there are for similar patterns. For example, P. ferox and P. gallapagensis have a similar pattern along VII- VIII; some authors regarded them as two series (Bakken 2007; Villalobos-Guerrero & Tovar- Hernández 2013) instead of a single jagged row. However, two jagged rows are clearly discernible in P. variegata (Rozbaczylo & Bolados 1980), useful for separate it from P. gallapagensis and P. ferox. On the other hand, papillae and their arrangement in nereidids have not been revised. Bakken et al. (2009) made two important conclusions regarding papillae. First, papillae and paragnaths are homologous because papillae have a similar distribution on pharynx, being possible to describe their positions with the same system for paragnaths. The second conclusion, that papillae are plesiomorphic respect to paragnaths, is discussed below. Sometimes the pharyngeal arrangement has been regarded as irrelevant because it is difficult to observe with accuracy, or because the same pattern can be found in other species, as in Platynereis where often the presence of pectinate rows is confirmed after revision under

49 compound microscope. However, the pharynx features, including type of paragnaths or papillae must not be overlooked because in most cases they are generic features.

Paragnaths and papillae number, forms and origin. The diagnostic relevance of using the number of paragnaths is shown by some keys made almost exclusively with pharyngeal features (e.g. Hutchings et al. 1991). Paragnaths number has been analyzed with advanced statistical procedures (Bakken et al. 2009; Coutihno et al. 2015). Often paragnaths number in areas I, V, VI and VII-VIII are sufficient for separating species and species group within a genus, e.g. the two main groups in Nereis having or not paragnaths on area V (Conde-Vela & Salazar-Vallejo 2015) or number of bars on areas VI in Perinereis (de León- González & Goethel 2013). Also, shape of paragnaths on areas VI and recently on VII-VIII are important because they are diagnostic, as the occurrence of point, shield-shaped and smooth bars. Bakken et al. (2009) performed an important effort to standardize the terms used for different paragnaths. Improvements and restrictions to the proposal were later made by Glasby et al. (2011), Villalobos-Guerrero & Carrera-Parra (2015) and Conde-Vela & Salazar-Vallejo (2015).

The morphology of papillae deserves some comments. They are generally rounded, and have a variety of sizes. As occur in Nereidinae, a species can have more than one type of papillae. Paraleonnates species have filiform and triangular or nearly so papillae, and in Laeonereis they are verticillated and triangular. On the other hand, the second phylogenetic assumption by Bakken et al. (2009), papillae are plesiomorphic respect to paragnaths, seems to be a more parsimonious hypothesis. Other indications for this is that members of the families Hesionidae and Syllidae, closely related to nereidids, also have papillae on their pharynx, but paragnaths are unique to nereidids, and at least in hesionids the smooth pharynx is plesiomorphic (Aguado & San Martín 2009; Pleijel 1998; Ruta et al. 2007); however, the homology regarding these papillae is unknown and at least their relative position differs. In addition, it is not clear if fleshy, papillae- like structures present in Ceratonereis and in Namalycastis and Namanereis are also homologues to the papillae present in gymnonereidins. Partially sclerotized structures have been regarded as papillae; however, the contrary assumption, that they are paragnaths not completely sclerotized, would be valid as well. For example, the current diagnosis of Paraleonnates include the presence of paragnaths (P. bolus) or papillae

50 partially sclerotized (P. uschakovi) on maxillary ring (Hong et al. 2012); in fact, these ‘papillae’ rather resemble paragnaths without basal, dark scleroprotein instead of papillae becoming hard. Other cases of high resemblance between these structures are the rod-like paragnaths in Solomononereis or the verticillated papillae in Laeonereis. The appearance of jaws is well known in some species; for example, it develops at nectochaeta stage (four days or three chaetigers) in nereidins as Nereis virens (Bass & Brafield 1972), Perinereis rullieri (Prevedelli & Cassai 2001), Platynereis dumerilii (Fischer et al. 2010), but also in the gymnonereidin Laeonereis culveri (Klesch 1970; Mazurkiewicz 1975). However, notes about the development and appearing of paragnaths and papillae during ontogeny are scarce. Bass & Brafield (1972) noted that in N. virens (now Alitta virens) paragnaths do not originate at same time, the first appears at 13-chaetiger stage and, judging from their text-figure 4F, paragnaths appear on area I only, and the fully set of paragnaths is present by the 40- chaetigers stage. Conversely, verticillated papillae appear later because they were visible at 20- chaetiger stage in L. culveri (Mazurkiewicz 1975). It must be emphasized that detecting pharyngeal structures, especially the colorless papillae at very small stages is difficult, and even in juveniles and adults. Reish (1957) for Neanthes caudata found that paragnaths appears first on maxillary ring at 14 days and in oral ring after 21 days, and become hardened after 26 days. The late appearing of paragnaths on oral ring might be relevant in phylogenetic analysis, especially to those genera having paragnaths restricted to either maxillary or oral rings. Finally, other assumption is that a smooth pharynx surface is plesiomorphic relative to pharynx with papillae. However, this smoothness can be a secondary loss; e.g., Glasby (1999) detailed the presence of papillae in juveniles of N. amboinensis and N. hummelincki in an unusual arrangement on oral ring, but they are not present in adult specimens. The simplification of structures in this subfamily is discussed in other sections of the present contribution. All these evidences indicate that pharyngeal armature, notwithstanding the both morphological and positional similarity, has different origins and evolved at different times; therefore the hyphotesys stated by Bakken et al. (2009) is very possible but not the unique event during the evolution of pharynx. Further studies of ontogeny in nereidids would clarify these and other doubts about the origin and significance of pharyngeal armature.

51

Parapodia, terminology. There are several recent efforts towards a standardization of parapodial terms. The terminology used by Glasby (1999) for Namanereidinae, Bakken & Wilson (2005) for Nereidinae, or Hylleberg & Nateewathana (1988) for Gymnonereidinae are often followed. As they noted, nereidid parapodia are very variable and the homology among structures is only supposed in most cases, such that some processes receiving the same name might have a different origin. Another important limitation is the insufficient information in original descriptions, requiring re-assessment of material. The topology of parapodial features is frequently described as size relationships involving lengths and sometimes width of cirri, ligules and lobes. However, the means to measure the beginning and ending points of lobes or ligules, especially being asymmetric structures, are not detailed in most works, and this is very important for indicating relative proportions. For these reasons, and trying to offer a simple, useful means to describe more appropriately the parapodial features, some general commentaries are added. Further, some glossaries were consulted trying to clarify some words commonly used (Fauchald 1977c; Glasby et al. 2000; Mikkelsen & Virnstein 1982). Naturally, the use of one or other term does not (or must not) affect the descriptions if every structure is well identified, and herein a short discussion is presented about which terms are the most appropriate to use, based upon definitions and not by their practical use. Because the first two chaetigers in nereidids are modified and give rise to the tentacular ring, no longer visible in adult specimens, the chaetiger counts start at 1 instead of 3, as the traditional way found in most literature. The first two chaetigers (chaetigers 3 and 4 in an ontogenetic view) have important information for taxonomical and phylogenetic issues, therefore they were classified, illustrated and included in the descriptions. As Santos et al. (2005) highlighted, there are problems about parapodial terminology for phylogenetic studies, and they recommended to describe parapodial structures instead to use terms involving different topologies. The Figure 1 shows the different varieties of first two chaetigers found in the genera studied, and because of their variation, different terms should be used for each pattern. Therefore, the terms uniramous was avoided, and a classification is proposed based upon the presence of main parapodial structures, including aciculae, as parapodial types A to E (Figs 1A–E, respectively). Type A parapodia are found in Stenoninereis and Micronereis; they have a neuropodial lobe and a ventral cirrus only, and the notopodia is represented with any other visible structures. Type B

52 parapodia are found in gymnonereidins as Ceratocephale and Gymnonereis; they include dorsal cirrus, neuropodial lobe and ligule and two ventral cirri; the variant is Gymnonereis having two ‘dorsal cirrus’, but here the second structure was regarded as a projection of cirrophore as Santos et al. (2005) stated in their Figs 3D–E instead of ‘accessory dorsal cirrus’, as it was referred to in text. Type C parapodia are present in most nereidins and in gymnonereidins; they have a dorsal cirrus, a notopodial ligule, neuropodial lobe and ligule, and a ventral cirrus. Type D parapodia are only present in Ceratonereis; these parapodia are very similar to type C ones but it has notoacicula crossing the dorsal ligule. Type E parapodia are exclusive to namanereidins; they have a dorsal cirrus, a neuropodial lobe and a ventral cirrus, and it is the only one present in all chaetigers. Type E parapodia are simplified because of the reduction of notopodia during ontogeny, and the current position of dorsal cirri on the neuropodial lobes in adults is the result of this. Feuerborn (1931) indicated this for Lycastis ranauaensis, and Glasby (1999) considered that parapodia of namanereidins are derived structures. Recently, Ezhova (2011) found that in 6-stage juveniles of Namanereis littoralis, two minor lobes with one acicula each, i.e. biramous parapodia, and therefore only ventral lobe develops and forming sub-biramous parapodia; another plus to this view is the presence of notochaetae in some namanereidins.

Parapodia, lobes and ligules. There are doubts about when the terms lobes or ligules must be used or if one have preference to the other. As defined, parapodial lobe is a major, conical process, with a rounded base or nearby; while ligule is a compressed process with an oval base (Glasby et al. 2000; Mikkelsen & Virnstein 1982). If Fauchald’s definitions (1977c) for ligules and lobes were used, most notopodial ligules should not be called as such because they are not digitiform but subconical; also, they can reach a large size, even larger than the whole parapodium as in Pseudonereis, Alitta and some Perinereis, and then fitting within the lobes definition instead. Furthermore, if one structure is conical, subconical or if it is compressed, is very subjective, additionally the same structure can modify its shape along the body and be conical in one region and digitiform in another. In the nereidids the separation between notopodial and neuropodial rami often is not so conspicuous and these difficult the selection of the term. As defined by Santos et al. (2005), true parapodial lobes carry aciculae, being the two most-medial structures in the parapodium, and divides it into notopodial and neuropodial rami, the remaining structures must be referred as

53 ligules. The main lobes often have processes derived from themselves, also referred as lobes, i.e. prechaetal or postchaetal lobes, depending of their relative position with the chaetae. The part of the lobe that carries the acicula must be referred as ‘acicular lobe’, as defined by Mikkelsen & Virnstein (1982). In this manner, a notopodial lobe is composed by a notoacicular lobe, and a pre- or postchaetal lobe can be present, whereas a neuropodial lobe is composed by a neuroacicular lobe and, in most cases, a postchaetal lobe. An additional structure found in the main lobes are the parapodial accessory cirri in Micronereis species, these are discussed in remarks of that genus. Generally, notoacicular lobes are not wholly crossed by the notoacicula (except in Micronereis), and chaetae arise from notacicular lobes bases. Notoacicular lobes can have prechaetal or postchaetal lobe. Sometimes, notopodial papillae are very prominent and can be confused with a prechaetal lobe, but they can be distinguished because they over acicular tips, which is not the case for prechaetal lobes. Notopodial papillae, if present, are more conspicuous on anterior and middle parapodia, but generally inconspicuous or absent in posterior segments. Neuropodia are more complex. Neuroacicular lobes are wholly crossed by neuroacicula, and can be distally subdivided into superior and inferior lobes (Fig. 1G). In some cases, these lobes can be hypertrophied and therefore very evident, as in Paraleonnates (Fig. 1G), but generally they are inconspicuous and, if present, restricted to the tip of neuroacicular lobe; superior and inferior lobes can be symmetrical or asymmetrical, giving a bilobed aspect to neuroacicular ones. On the other hand, sometimes the schematization of neuropodial postchaetal lobe can lead to confusion. As Hylleberg et al. (1986) drawn postchaetal lobes, they can be regarded as independent structures from neuroacicular lobes. However, postchaetal lobes are part of the main lobe (Figs1F, G), and in the material examined here, it never has been found as an independent structure. Generally it appears as an overlapped line near the tip or beyond neuroacicular lobes, but it is more conspicuous when seen from above, giving a bilobed appearance (Figs 1H–K). In most cases, chaetal bundles or a notch allow us to recognize this structure, and it is absent if chaetae originate along neuroacicular lobes, as in Namanereidinae, or when there is no clear notch. Bakken & Wilson (2005) considered that neuropodial prechaetal lobe as currently defined is absent among Nereidinae, and this is true because it cannot be found in the material examined nor in other subfamily studied here.

54

Parapodial morphology varies among genera mainly in the absence or presence of ligules and pre- or postchaetal lobes and their relative development, also the major differences are found among subfamilies: Gymnonereidinae generally has more complex parapodia than Nereidinae, and the latter more than Namanereidinae. In non-dissected parapodia, ligules and lobes resemble oblique cones with oval bases, or fingers if their sides are nearly parallel, with both rounded tips but the former being more acute. However, once mounted, the parapodia are compressed and the tridimensional aspect is lost, and this is the outline usually described. Of course, describing the profile of the lobes and ligules is easier for a tridimensional structure, and the resultant description will be more useful than only say ‘X structure is subconical/digitiform’, when most have this shape, but in several variants. The way we mount the dissected parapodium would affect the original shape, especially if it is very compressed or if their structures overlap, but not the length ratios because all structures are equally affected. Furthermore, in many cases it is necessary to describe the tridimensional aspect; for example, a structure can vary their thickness together with its length along the body as occur with the notopodial ligules in genera Alitta and Neanthes, or the dorsal cirri in Namalycastis, which is commonly regarded as foliose or lamellar.

Parapodial structures and measurements. As noted above, the procedure to indicate parapodial ratios is not clarified in most publications. The procedure herein used is depicted in Figure 1L, including the abbreviations used; this consists of placing three points to draw a triangle from the projected lobes tips to the center of their bases. Point A is the dorsal-most point, point B the ventral-most one, and point C the distal-most one (Fig. 1L). Points A and B are placed at base of structure, generally where there is an incision or intersection, and once placed they form the segment AB, regarded as the base (Ba). The midpoint between A and B (mAB) must be obtained, then C is defined as the farthest point from mAB (Fig. 1L). Once these points are set, some other segments are formed. Segment BC corresponds to the ventral extension (VeEx), and segment AC is the dorsal extension (DoEx). The partial contour of the structures was not used because it can increase if structure is rounded or has a convex side, but these would not increase their relative extension, as is the case for most notopodial ligules (Fig. 1L).

55

For parapodial cirri, three additional points and two distances are needed. Points a and b are placed where the dorsal or ventral cirrus are inserted into their respective ligule or lobe, and the segment ab forms the base of dorsal (bDc), or ventral (bVc) cirri. Point c is the farthest point from mab (Fig. 1L). As the parapodial cirri frequently are curved in mounted parapodia, the resultant curve between C and mab, called length of dorsal or ventral cirrus (LDc or LVc), must cross internally the cirrus in all its length (Fig. 1L). An additional segment results between C and mab, and it is called tip extension (tEx), useful for describing the reach of the dorsal cirrus against notopodial ligule (Fig. 1M). Formally, an arc can be formed taking LDc as the radio and mab as center; if the arc intersects with tEx, then dorsal cirri do not extend beyond notopodial ligules, on the contrary, dorsal cirri extend beyond notopodial ligules (Fig. 1M). The exact position was not determined, because describing the general pattern of modifications along body was regarded as more relevant, and this information is provided as the ‘parapodial cirri pattern’ paragraphs in descriptions. To indicate the insertion of dorsal cirri, a perpendicular line is drawn between mab and DoEx (Fig. 1N). Then, DoEx is divided into three segments: two for medial and one for subdistal insertion, thus where the perpendicular line falls will indicate the type of insertion (Fig. 1N). The point mab lies on the same point as, or are very close to A in a basal insertion, and to C for a distal one. These distances are used for calculating ratios and several can be obtained easily or in an intuitive way, but many others are more troublesome. Hence, numerical descriptions are not promoted, but ratios help to improve descriptions and making them more precise. In fact, if a numerical ratio can be replaced by a sentence, it was preferred, e.g. ‘twice longer’ and ‘half as long as’, and if the difference would be just a few decimals, ‘slightly’ or ‘subequal’ were used. Moreover, the configurations herein shown were useful for species described; however, there are more configurations in other species and genera, and this proposal must be taken as an example, not as rule. The estimation of ratios for lobes and ligules is performed by using the adjacent sections, concurrent in a common point. Thus, the ratio between notopodial ligule and notoacicular lobe is estimated with VeEx of the former and DoEx of the latter, and the ratio between notoacicular and neuroacicular lobes with VeEx and DoEx, respectively. Sometimes the structures do not share any adjacent point, as when neuropodial ligule is medially inserted to neuroacicular lobe, or the

56 structures are not adjacent to each other, as when notopodial and neuropodial ligules are compared. For the first case, VeEx of neuropodial ligule and VeEx of neuroacicular lobe were used, and VeEx of notopodial ligule and DoEx of neuropodial one for the second case. To estimate ratios of dorsal and ventral cirri, the sections LDc and DoEx of notopodial ligules or LVc and VeEx of neuropodial ligules were used. Although all measurements described above may be included in the descriptions, this is not encouraged because some ratios are more informative than others, and they can vary among genera. With this approach, more discriminant ratios were the length of dorsal and ventral cirri, given by LDc or LVc and DoEx or VeEx of notopodial or neuropodial ligules, respectively; and the DoEx of a ligule or lobe and the VeEx of the upper one, e.g. DoEx of notoacicular lobe and VeEx of notopodial ligule. Often in first two chaetigers, the notopodial and neuropodial ligules were directly compared because the neuroacicular lobe is reduced; in that case VeEx of notopodial ligule and DoEx of neuropodial one were used. Furthermore, in some cases the Ba of a structure was compared in the same structure along body or against the Ba of other one.

Chaetae. The nereidids have several types of chaetae, compound or simple ones, often being an important feature to specific and generic levels (Bakken & Wilson 2005). The chambered chaetae are a feature shared with hesionids and chrysopetalids (Pleijel & Gustavsson 2010). The compound or jointed chaetae are composed by two structures, the blade and the shaft. The shaft can be homogomph, heterogomph or sesquigomph, an here the proposal by Glasby (1999) to determines the type of articulation was used. The blades generally are dentate; two types of teeth are recognized based upon in their position on the blade. The teeth of the inned edge of blades can be present in all margin or be restricted to a portion of the margin, this feature was reported in the descriptions. The number of teeth was not reported since in most cases that number is related with the length of the blade, and therefore variable along body and even in the same fascicle. On the other hand, the distal teeth are stouter than the found in the inner edge, in many cases they develope a tendon attached to the inner edge, or have an accessory distal mucron. The reach of the tendon and the position and size of the mucron are useful as additional features to recognize species.

57

Pygidium. The pygidium often is disregarded as diagnostic; however, their shape and the length of the anal cirri could be useful. Glasby (1999) recognized two types of pygidium in namanereidins: multi-incised (here called crenulated), or tripartite (having three lobes). These shapes were also found in other subfamilies (see below). In epitokes, the pygidium develops a rosette of papillae, but can take other shapes as in some Platynereis species (Read 2007).

Systematic account

Family Nereididae de Blainville, 1818

Subfamily Nereidinae de Blainville, 1818

Alitta Kinberg, 1865

Type species. Nereis virens Sars, 1835, by monotypy.

Diagnosis (after Villalobos-Guerrero & Carrera-Parra 2015). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary ring with conical paragnaths in areas I-IV. Oral ring with conical, p-bars or pyramidal paragnath in areas VI–VIII, area V with or without paragnaths. First two chaetigers type C (¿?) or D. Notopodial ligules elongate, pennant-like toward posterior chaetigers. Notopodial prechaetal lobes present at least in anterior and middle chaetigers; neuropodial postchaetal lobe present throughout body. Notochaetae homogomph spinigers; neurochaetae homogomph spinigers and heterogomph falcigers in supra- acicular fascicles, heterogomph falcigers and spinigers in sub-acicular fascicles.

Remarks. One species, Alitta succinea (Leuckart, 1847), has been widely reported in several parts of the world, and recorded from the Grand Caribbean region as Nereis (Neanthes) succinea or Neanthes succinea. However, some regional names must be reinstated after the redescription of A. succinea by Villalobos-Guerrero & Carrera-Parra (2015).

58

The diagnosis they provided include notoaciculae present or absent in first two chaetigers, without further discussion about, since A. succinea and A. acutifolia (Ehlers, 1901) have notoacicular in first two chaetigers (type D); here the presence of type C chaetigers is regarded as doubtful (¿?).

Questionable records

Alitta succinea (Leuckart, 1847)

Nereis succinea Leuckart 1847: 154–156, Pl. 2, Figs 9, 11. Alitta succinea Bakken & Wilson 2005: 516–517 (partim). Villalobos-Guerrero & Carrera-Parra 2015: 161–167, Figs 1A–C, 2A–K.

Type locality. Helgoland, Germany.

Remarks. This species was recently redescribed by Villalobos-Guerrero & Carrera-Parra (2015) based on type and topotypic material. A. succinea has a cosmopolitan status, and commonly reported as invasive species; however these records must be re-assessed and several species synonymized with A. succinea must be reinstated.

Ceratonereis Kinberg, 1865

Type species. Ceratonereis mirabilis Kinberg, 1865, by subsequent designation (Hartman 1948).

Diagnosis (modified from Bakken & Wilson 2005). Prostomium with anterior margin deeply incised. Eyes present. Four pairs of tentacular cirri, cirrostyles long and slender, very conspicuous. Maxillary ring with conical paragnaths in areas II-IV, area I with or without paragnaths. Oral ring with cushion-shaped papillae on areas VI, V and VII-VIII bare. First two chaetigers type D. Notopodial ligules reduced or absent in posterior chaetigers. Postchaetal lobe present throughout body Notochaetae sesquigomph spinigers and falcigers; neurochaetae

59 sesquigomph spinigers and heterogomph falcigers in supra-acicular fascicles, sesquigomph or heterogomph falcigers and heterogomph spinigers in sub-acicular fascicles.

Remarks. Perkins (1980) observed “soft, cushion-shaped lobes” on Area VI in C. mirabilis and others species, also he noted these structures only appear in species with a cleft prostomium, and recognized these as relevant features for the erection of a new genus or subgenus for species without them. Hartmann-Schröder (1985), attending to Perkins’s observations, proposed three subgenera: Ceratonereis for species with anterior margin of prostomium incised, Composetia for species with anterior margin entire and Simplisetia for species also with ankylosed falcigers in neuropodia. Khlebovich (1996) regarded Simplisetia and Composetia as distinct genera to Ceratonereis. Also, she diagnosed Composetia and Simplisetia without structures on area VI, but the subgenus Ceratonereis as ‘usually with a soft cushion or papillae on area VI’. All species recorded here have papillae on area VI as in the type species; therefore, the oral ring was not reported as bare but having papillae on this area. The genus Unanereis Day, 1962 have close phylogenetic relationships with Ceratonereis, and regarded as distinct genus in a recent revision (Bakken & Wilson 2005). Its sole species U. macgregori also have cushion-shape papillae on Area VI, but the feature were disregarded as valuable for phylogeny. The species can be grouped of two main features, the presence of unidentate or bidentate falcigers and the presence or absence of notopodial ligules on posterior chaetigers, being four main groups. A distinctive feature of Ceratonereis species are the well-developed ceratophores of dorsal cirri and the sensible diminution of size and often migration of both notopodial and neuropodial ligules. Here is considered that ceratophore carry notopodial ligule and not the contrary, and the ratios were obtained with cirrostyle and notopodial ligule lengths because in anterior chaetigers is difficult to discern when ends the notopodial ligule and when begins the ceratophore. Generally, notopodial ligules have a glandular mass very visible and both diminish their sizes proportionally, not found in ceratophores. (e.g. Fig. 4J–N).

60

Ceratonereis gracilis (Webster, 1884) reinst. Figure 2A–Q

Nereis gracilis Webster 1884: 313–314, Pl. 9, Figs 29–35 (non N. gracilis Hansen, 1882). Ceratonereis gracilis Hartman 1938b: 13, 14. Ceratonereis mirabilis Perkins 1980: 4–1, Figs ?? (partim, non Kinberg, 1865) Ceratonereis excisa Salazar-Vallejo & Jiménez-Cueto 1997: 363, Figs 1,15,16,37,38 (partim, non Grube & Müller in Grube, 1873)

Type material. Northwestern Atlantic Ocean, Bermuda. Syntypes of Nereis gracilis USNM 4787 (2), Bermuda, 1876, Coll. G.B. Goode.

Additional material. Northwestern Atlantic Ocean. Bahamas. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7106, St. 1431 (21°41'N 73°51'W), Hogsty Reef, Bahamas, 27 m depth, 22 July 1971. Caribbean Sea, Mexico. ECOSUR P1118 (15), Buenavista, Quintana Roo. ECOSUR P1148 (1), Punta Allen, Quintana Roo. ECOSUR P1151 (7), Xahuayxol, Quintana Roo. ECOSUR P1152 (6), R/V Edwin Link, Banco Chinchorro, Quintana Roo, 20 August 1990. ECOSUR P1153 (5), R/V Edwin Link, Punta Herradura, Quintana Roo, 28 September 1996. ECOSUR P1154 (4), R/V Edwin Link, Banco Chinchorro, Quintana Roo (EPITOKES). ECOSUR P1156 (2), Chichorro Bank, Quintana Roo. ECOSUR P1160 (1), R/V Edwin Link, Banco Chinchorro, Quintana Roo. ECOSUR P1204 (2), Sabancuy, Campeche, 28 August 1984, Coll. E. Escobar. ECOSUR P0000 (1), Cayo Norte, Banco Chinchorro, Quintana Roo, May 1998. Jamaica. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7006, St. 1195 (17°24'N 76°02'W), SE Jamaica, 14 m depth, 3 July 1970. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7006, St. 1217 (17°33'N 77°13'W), S Jamaica, 27 m depth, 6 July 1970. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7006, Sta. 1230 (17°52'N 77°58'W), SW Jamaica, 31 m depth, 7 July 1970. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7006, Sta. 1233 (17°60'N 78°05'W), SW Jamaica, 21 m depth, 7 July 1970. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7006, St. 1258 (17°03'N 78°10'W), Pedro Bank, 15 m depth, 15 July 1970. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7006, St. 1259 (17°04'N 77°55'W), S Jamaica, 11 m depth, 15 July 1970. ESCOSUR P0000 (3), R/V Pillsbury, Cruise 7006, St. 1250 (17°20'N 78°48'W), SW Jamaica, 24 m depth, 14 July 1970. ECOSUR

61

P0000 (1), R/V Pillsbury, Cruise 7006, St.1196 (17°28'N 75°57'W), SE Jamaica, 26 m depth, 3 July 1970. Antigua and Barbuda. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 966 (17°08'N 62°03'W), NW Antigua, 18 m depth, 20 July 1969. ECOSUR P0000 (7), R/V Pillsbury, Cruise 6802, St. 975 (17°29'N 61°55'W), SW Barbuda, Lesser Antilles, 29 m depth, 21 July 1969. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 967 (17°16'N 62°02'W), NW Antigua, 22 m depth, 20 July 1969. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6806, St. 968 (17°18'N 61°52'W), N Antigua, Lesser Antilles, 18 m depth, 20 July 1969. Saint. Barthélemy. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 981 (18°01'N 62°55'W), Saint-Barthélemy Channel, Lesser Antilles, 22 m depth, 22 July 1969. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6806, St. 979 (17°51'N 62°39'W), ESE Saint-Barthélemy, Lesser Antilles, 37 m depth, 22 July 1969. Anguilla. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 983 (18°20'N 62°38'W), S Anguilla Valley, Lesser Antilles, 49 m depth, 22 July 1969. St. Kitts and Nevis. ECOSUR P0000 (2) R/V Pillsbury, Cruise 6802, St. 955 (17°04'N 62°34'W), S Nevis, Lesser Antilles, 11 m depth, 19 July 1969. ECOSUR P0000 (8), R/V Pillsbury, Cruise 6907, St. 961 (17°27'N 62°51'W), NW St. Kitts, Lesser Antilles, 11 m depth, 19 July 1969. Guadeloupe. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 939 (16°22'N 61°09'W), NW La Désirade, Lesser Antilles, 95 m depth, 16 July 1969. Caribbean Netherlands. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6907, St. 965 (17°26'N 63°23'W), Saba Bank, 24 m depth, 20 July 1969.

Type locality. Bermuda.

Description. Syntypes include two anterior ends, almost dried. One with deep cuts along some chaetigers caused by previous dissections (Fig. 2B), 55 mm long, 1.2 mm wide, 30 chaetigers; another syntype 76 mm long, 1.4 mm wide, 33 chaetigers. Non-type specimen from Punta Allen (P1148) used for description, incomplete, 21 mm long, 1.5 mm wide, 44 chaetigers. Body slightly reddish, without other pigmentation. (Fig. 2C). Prostomium wider than long, anterior margin deeply incised; antennae subulate, twice longer than prostomium, slightly passing palps; eyes rounded, subequal, in trapezoidal arrangement (Fig. 2C). Tentacular ring as long as first chaetiger; four pairs of tentacular cirri, posterodorsal ones incomplete, longest one reaches chaetiger 13 (Fig. 2C).

62

Pharynx dissected; jaws light brown, 6 teeth, restricted to distal half (Fig. 2Q). Maxillary ring: I=0, II= 9–10 cones in round, III= 5 cones in round, IV=8–9 cones in round. Oral ring: V and VII-VIII bare, VI= 1 cushion-shaped papillae. Parapodial cirri pattern: Dorsal cirri longer than notopodial ligules throughout body, passing greatly other parapodial structures throughout body, basally inserted in anterior-most chaetigers; notopodial ligules progressively smaller toward posterior end, without evident migrating toward basal joint of dorsal cirri. Ventral cirri shorter than neuroacicular lobes in anterior and middle body, becoming subequal or slightly longer than neuroacicular ones in posterior body, basally inserted throughout body. First two chaetigers type D, remaining biramous. In first two chaetigers (Figs 2D, G), dorsal cirri basal, cirrostyles 1.5 times longer than notopodial ligules. Notopodial ligules digitiform, blunt tip, twice longer than neuroacicular lobes. Neuroacicular lobes digitiform, twice longer than postchaetal ones; postchaetal lobes rounded; neuropodial ligules digitiform, blunt tip, 1.2 times longer than neuroacicular lobes. Ventral cirri basal, 0.8 length of neuropodial ligules. In anterior chaetigers (Figs 2E, H, I), dorsal cirri basal, cirrostyles 4-5 times longer than notopodial ligules. Notopodial ligules digitiform, acute tip, as long as notoacicular lobes; notoacicular lobes digitiform to subconical, as long as neuroacicular lobes. Neuroacicular lobes subconical, twice longer than postchaetal lobes; postchaetal lobes rounded, lamellar; neuropodial ligules subconical, subequal to neuroacicular lobes. Ventral cirri basal, subequal to neuropodial ligules. In middle chaetigers (Fig. 2J), dorsal cirri basal, cirrostyles several times longer than notopodial ligules. Notopodial ligules subconical, reduced; notoacicular lobes subconical, 3 times longer than notopodial ligules. Neuroacicular lobes subconical, wide base, 1.5 times longer and twice wider than notopodial ones, twice longer than postchaetal ones; postchaetal lobes rounded, lamellar; neuroacicular lobes digitiform, acute tip, reduced, medially inserted to neuroacicular lobes, 1/3 as long as neuroacicular lobes. Ventral cirri basal, 0.6 times length of neuroacicular lobes. In posterior chaetigers (Fig. 2K), dorsal cirri basal, cirrostyles several times longer than notopodial ligules. Notopodial ligules subconical, very reduced; notoacicular lobes subconical, 5 times longer than notopodial ligules. Neuroacicular lobes subconical, slightly shorter than notoacicular ones; postchaetal lobes reduced; neuropodial ligules subconical, acute tip, reduced,

63 medially inserted to neuroacicular lobes, 1/3 as long as neuroacicular lobes. Ventral cirri basal, subequal to neuroacicular lobes. Notochaetae sesquigomph spinigers and falcigers. Neurochaetae sesquigomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular fascicles. Notopodial sesquigomph spinigers pectinate, teeth decreasing in size toward distal end; sesquigomph falcigers shallowly bidentate (Fig. 2N), pectinate, fine teeth, increasing their length toward distal end, blades of falcigers with different length, even in the same fascicle. Neuropodial sesquigomph spinigers as notopodial ones; heterogomph spinigers pectinate, teeth decreasing in size toward posterior end. Heterogomph falcigers in anterior-most chaetigers (Figs L, M) distally unidentate, pectinate, fine teeth, blades longer and slender than in posterior-most chaetigers, teeth shorter than in posterior-most ones; blades becoming bidentate, shorter and stouter toward posterior chaetigers (Figs 2O,P). Both supra- and sub-acicular falcigers with distal teeth extending slightly beyond tips; sub-acicular falcigers with tip stouter than supra-acicular ones, distal tip less developed, rounded (Fig. 2P). Pygidium unknown.

Remarks. Syntypes are in poor conditions, most dorsal cirri are missing and the remaining ones are incomplete. Perkins (1980) revised the type materials of C. mirabilis Kinberg and concluded that C. gracilis is a junior synonym, arguing minor differences between them. Although both species appear to be inseparable, the enormous distance between both type localities (9°S, coasts of Brazil for C. mirabilis) make us hesitate they are conspecific. The available specimens and descriptions do not allow a complete separation; however, C. gracilis is considered as a valid species, redescribed and reinstated. Some differences can be noticed with available data. With regard to the cirrostyle/notoacicular lobe ratio, in C. gracilis, this ratio is lower than in C. mirabilis in all body. Also, in C. gracilis, notopodial sesquigomph falcigers have blades with different length throughout body, even in the same fascicle; whereas in C. mirabilis they are of the same length in the same fascicle, if the blade have different sizes along the body need to be assessed. There are no posteriorly complete specimens in material examined, but illustrations and description performed by Webster (1884) provide some diagnostic features useful for their

64 recognition. Posterior parapodia drawn by Webster (1884, Pl. 9, fig. 32) indicate that ventral cirri become slightly longer than neuroacicular lobes, as indicated in parapodial cirri pattern section. This condition is also present in C. mirabilis and in other species of the genus (Perkins 1980, Hartmann-Schröder 1985).

Distribution. Bahamas, Mexico, Jamaica, Antigua and Barbuda, Saint Barthélemy, Anguilla, St. Kitts and Nevis, Guadeloupe, Caribbean Netherlands.

Ceratonereis longicirrata Perkins, 1980 Figure 3A–O

Ceratonereis longicirrata Perkins 1980: 26–32, Figs 11, 12. Salazar-Vallejo & Jiménez-Cueto 1997: 363, Figs 4, 19, 40, 41.

Type material. Northwestern Atlantic Ocean, United States. Holotype USNM 58740, Hutchingson Island, St. Lucie County, Florida (27°19.1’N 80°13.2’W), 8.2 m depth. Paratypes USNM 58741 (5), data as in the holotype.

Additional material. Caribbean Sea, Mexico. ECOSUR P1111 (3), Laguna Nichupté, Quintana Roo, 5 July1988, Coll. S. Jiménez-Cueto. ECOSUR P1149 (2), Laguna Nichupté, Quintana Roo, 30 October 1987, Coll. M.S. Jiménez-Cueto. ECOSUR P1155 (1), R/V Edwin Link, Mahahual, Quintana Roo, 24 August 1990. ECOSUR P1159 (2), Mahahual, Quintana Roo, 1 October 1996, Coll. S.I. Salazar-Vallejo, L.F. Carrera-Parra. ECOSUR P1161 (2), R/V Edwin Link, Punta Changuay, Quintana Roo, 26 August 1990. ECOSUR P1162 (2), Laguna Nichupté, Quintana Roo, 30 October 1987, Coll. M.S Jiménez-Cueto. ECOSUR P1163 (3), Laguna Nichupté, Quintana Roo, 30 October 1987, Coll. M.S. Jiménez-Cueto. ECOSUR P1164 (1) Laguna Nichupté, Quintana Roo, 19 April 1988, Coll. M.S. Jiménez-Cueto. ECOSUR P1165 (2), Laguna Nichupté, Quintana Roo, 30 October 1987, Coll. M.S. Jiménez-Cueto. ECOSUR P1205 (1), Laguna Nichupté, Quintana Roo, 28 October 1987, Coll. M.S. Jimémez-Cueto. ECOSUR P1207 (4), Playa Aventuras, Quintana Roo, 18 May 1986, Coll. S.I. Salazar-Vallejo. ECOSUR P1217 (17), Punta Nizuc, Quintana Roo, 1 September 1997, Coll. L.F. Carrera-Parra, S.I. Salazar-

65

Vallejo, M.A. Ruiz-Zárate. ECOSUR P0000 (2), Puente Las Coloradas, Yucatán (21°35'36.82"N 88° 3'21.75"W), on dark greenish sponges, in mangrove roots, 31 May 2005, Col. S.I. Salazar- Vallejo, L.F. Carrera-Parra. ECOSUR P0000 (1), Puente Las Coloradas, Yucatán (21°35'36.82"N 88° 3'21.75"W), on algae in piles, 15 October, 2009, Coll. S.I. Salazar-Vallejo, L.F. Carrera- Parra. Jamaica. ECOSUR P0000 (2), R/V Pillsbury, Cruise 7006, St. 1252 (17°09'N 78°57'W), W Pedro Bank, Jamaica, 26 m depth, 14 July 1970. Dominican Republic. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7001, Sta. 1156 (19°20'N 69°10'W), N Cabo Samaná, 27 m depth, 16 January 1970.

Type locality. Hutchinson Island, Florida (27°19.1’N, 80°13.2’W).

Description. Holotype complete, in two portions, 20 mm long, 0.7 mm wide, 59 chaetigers; body whitish, without pigmentation (Fig. 3C). Non-type material, the most incomplete with both tentacular and parapodial dorsal cirri missing. Specimen from Nichupté (ECOSUR P1164) used for description in good conditions, incomplete posteriorly, tapered. Body pale, 14 mm long, 0.9 mm wide, 54 chaetigers, pigmentation absent. Other specimen from Las Coloradas (ECOSUR P0000) used to illustrate the pharynx. Prostomium as long as wide, anterior margin deeply incised; antennae subulate, as long as prostomium, slightly passing palps; eyes rounded, subequal, in trapezoidal arrangement (Figs 3A, D). Tentacular ring as long as first chaetiger; four pairs of tentacular cirri, cirrophores conspicuous, longest one reaches chaetiger 15 (Figs 3A, D). Pharynx everted; jaws amber, 8 teeth, restricted to first two-thirds of cutting edge (Fig. 3E). Maxillary ring: I= 0, II= 8–9 cones in round, III= 9 cones in round, IV= 14–18 cones in ellipse. Oral ring: V and VII-VIII bare, VI= 1 cushion-shaped papilla (Figs 3F, G). Parapodial cirri pattern: Dorsal cirri longer than notopodial ligules throughout body, passing greatly all parapodial structures throughout body, basally inserted in anterior-most chaetigers; notopodial ligules progressively reduced in size greatly toward posterior end, with evident migration toward basal joint of dorsal cirri. Ventral cirri shorter than neuroacicular lobes throughout body, basally inserted throughout body. First two chaetigers type D, remaining biramous. In first two chaetigers (Fig. 3K), dorsal cirri basal, cirrostyles 1.5 times longer than dorsal ligules. Dorsal ligules linear, blunt tip, twice longer

66 than neuroacicular lobes. Neuroacicular lobes subconical; postchaetal lobes reduced, papillae- like; neuropodial ligules digitiform, twice longer than neuroacicular lobes. Ventral cirri basal, 4/5 as long as neuropodial ligules. In anterior chaetigers (Figs 3L-O), dorsal cirri basal, cirrostyles 3.5 to several times longer than notopodial ligules and all parapodial structures toward middle chaetigers. Notopodial ligules linear, blunt tip, slightly longer than notoacicular lobes in chaetigers 3-10, becoming shorter than notopodial ones in chaetiger 17, very reduced and migrating to basal joint of dorsal cirri by chaetiger 19, absent from chaetiger 20; notoacicular lobes digitiform, becoming subconical toward middle chaetigers, slightly longer or subequal toward middle chaetigers. Neuroacicular lobes subconical, blunt tip; postchaetal lobes reduced, papillae-like; neuropodial ligules digitiform, displacing medially and becoming shorter to neuroacicular lobes toward middle chaetigers. Ventral cirri basal, shorter than neuroacicular lobes. Notochaetae sesquigomph spinigers and falcigers. Neurochaetae sesquigomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular fascicles. Notopodial sesquigomph spinigers pectinate, fine teeth, decreasing their length toward distal end; sesquigomph falcigers distally bidentate, pectinate, coarse teeth, increasing their length toward distal end, distal teeth barely extended beyond distal tooth (Fig. 3H). Neuropodial sesquigomph and heterogomph spinigers with blades as in notopodial ones. Neuropodial heterogomph falcigers pectinate, distal teeth barely extended beyond distal tooth, distally unidentate along first 17-20 chaetigers (Fig. 3I), bidentate in following ones (Fig. 3J); sub-acicular heterogomph falcigers with blades longer than supra-acicular ones in anterior chaetigers, becoming shorter toward posterior ones. Pygidium tripartite; anal cirri cirriform, as long as last 10 segments (Fig. 3B).

Variation. In some specimens, notopodial ligules disappear from chaetiger 18. Unidentate falcigers are replaced by bidentate ones from chaetigers 17–20. Other features vary little.

Remarks. The disappearing of the notopodial ligules and the appearing of the notopodial sesquigomph falcigers are common attributes to distinghish species in Ceratonereis. The original description is complete and attributes therein described match well with the type material

67 examined. However, some features were not illustrated or described, as the first two chatetigers, the pharynx including jaws, and posteror end. In order to improve the original description about these issues, and to show the actual state of type material, a brief description including figures was made. The species is easily recognizable by the combination of the disappearing of notopodial ligule toward posterior chaetigers and the shape of blades of notopodial sesquigomph falcigers and the very enlarged dorsal cirri in all chaetigers. Salazar-Vallejo & Jiménez-Cueto (1997) inverted the appearing of unidentate and bidentate falcigers in their description.

Ceratonereis setosa sp. n. Figure 4A–O

Ceratonereis sp. Perkins 1980: 32–34, Figs 13a–f. Ceratonereis singularis Salazar-Vallejo & Jiménez-Cueto 1997: 365 (non Treadwell, 1929). Ceratonereis excisa Salazar-Vallejo & Jiménez-Cueto 1997: 363, Figs 1, 15, 16, 37, 38 (partim, non Grube & Müller in Grube, 1873).

Type material. Caribbean Sea, Mexico. Holotype ECOSUR 0000, Isla Contoy, Quintana Roo, Mexico, 1 March 2001, Coll. Paratypes ECOSUR 0000 (1), Cabo Catoche, 2.5 m depth, 10 May 2009. ECOSUR 0000 (1), R/V Edwin Link, St. 2786 (18°20.73'N 87°46.51'W), Punta Gavilán, Quintana Roo, 261 m depth, 25 August 1990, Coll. E. Escobar, L. Soto. ECOSUR 0000 (3), R/V Edwin Link, St. 2786 (18°20.73'N 87°46.51'W), Punta Gavilán, Quintana Roo, 25 August 1990, Coll. E. Escobar, L. Soto. ECOSUR 0000 (1), Punta Allen, Quintana Roo (19°47'58''N 87° 28'29''W), 2 June 1986, Coll. E. Donath. ECOSUR 0000 (1), Ría Lagartos, Yucatán (21°34'5.15''N 88°13'58''W), 18 February 1999, Coll. S.I. Salazar-Vallejo, J.R. Bastida-Zavala.

Additional material. Caribbean Sea, Mexico. ECOSUR P1150 (4), Laguna Yalahau, 8 September 1993, Coll. S.I. Salazar-Vallejo. ECOSUR P1157 (1), Punta Gorda, 27 August 1984, Coll. E. Escobar. ECOSUR P1158 (1), Laguna Nichupté, Quintana Roo (21°06'11.6"N 86°47'21.1"W), 3 February 1988, in seagrasses, Coll. M.S. Jiménez-Cueto. ECOSUR P1167 (2), Laguna Nichupté, Quintana Roo (21°06'11.6"N 86°47'21.1"W), 30 October 1987, Coll. M.S.

68

Jiménez-Cueto. ECOSUR P1197 (2), Laguna Nichupté, Quintana Roo (21°06'11.6"N 86°47'21.1"W), 5 July 1988, Coll. M.S. Jiménez-Cueto. ECOSUR P1200 (1), R/V Edwin Link, St. 2774 (18°45.63'N 87°15.84'W), Cayo Norte, Banco Chinchorro, Quintana Roo, 20 August 1990, Coll. E. Escobar, L. Soto. ECOSUR P1208 (2), Laguna de Términos, Campeche, 27 July 1984, Coll. E. Escobar. ECOSUR P1220 (1), Buenavista, Quintana Roo (18°30'42''N 87°45'30"W), 27 September 1996, Coll. S.I. Salazar-Vallejo, L.F. Carrera-Parra. ECOSUR P0000 (1), Isla Contoy, Quintana Roo, on seashell, 26 February 2008. ECOSUR P0000 (1), Puerto Viejo, Isla Contoy, on mangrove roots, 24 February 2008. ECOSUR P0000 (2), Isla Contoy, Quintana Roo, Mexico, 1 March 2001, Coll. Jamaica. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7006, Sta. 1258 (17°03'N 78°10'W), Pedro Banks, Jamaica, 15 m depth, 15 July 1970. Nicaragua. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7101, St. 1330 (11°51'N 83°27'W), SE Nicaragua, 24 m depth, 28 January 1971. Venezuela. USNM 47771 (1), Cumaná Key, Mochima, 12 m depth, 21 October 1971, Coll. Edwards. USNM 49709 (1), Las Gabarras Beach, Mochima Bay, 8 m depth, 11 May 1971, Coll. R. Edwards. British Virgin Islands. LACM-AHF 7370 (2) White Bay, Guana Island (18°28’32”N, 64°34’39”W), on artificial reef matrix structures (ARMs), 1.5 m depth, 9 July 2000, Coll. T. Zimmerman, G. Hendler, R. Ware, K. Fitzhugh.

Other material examined. Tropical Eastern Pacific, Mexico. Non-type specimens of Ceratonereis singularis Treadwell, 1929. LACM-AHF 7408 (2), R/V Velero III, Sta. 1093-40 (25°49’25”N 111°18’35”W), Puerto Escondido, Baja California Sur, 15–27 m depth, 10 February 1940, Coll. Allan Hancock Foundation (AHF). LACM-AHF 7410 (2), R/V Velero III, Sta. 1903- 40 (25°49’25”N 111°18’35”W), Puerto Escondido, Baja California Sur, 15-27 m depth, 10 February 1940, Coll. AHF. LACM-AHF 7411 (5), R/V Velero III, Sta. 662-37 (25°31’35”N 111°01’45”W), off San Marcial Reef, Agua Verde Bay, Baja California Sur, 15 m depth, 11 March 1937, Coll. AHF. LACM-AHF 7412 (1), R/V Velero III, Sta. 683-37 (26°53’50”N 111°52’25”W), off Bahía Concepción, Baja California, 22 m depth, 15 March 1937, Coll. AHF. LACM-AHF 7413 (1), R/V Velero IV, Sta. 2024-51(27°48’33”N 114°42’30”W to 27°49’00”N 114°42’09”W), Scammon's Lagoon, Baja California Sur, 13-16 m depth, 18 Apil 1951, Coll. AHF. LACM-AHF 7415 (1), R/V Velero III, Sta. 277-34 (21°51’35”N, 105°54’30”W), Isabel Island, Nayarit, 18-46 m depth, 5 March 1934, Coll. AHF. LACM-AHF 7416 (3), R/V Velero IV, Sta. 1101-40 (25°31’00”N, 111°01’45”W), Agua Verde Bay, Baja California Sur, 18 m

69 depth, 12 February 1940, Coll. AHF. LACM-AHF 7417 (2), R/V Velero III, Sta. 1103-40 (25°31’05”N, 111°02’30”W), Agua Verde Bay, Baja California Sur, shore, 12 February 1940, Coll. AHF.

Etymology. The specific name stems from the Latin adjective setosus, a, um, meaning bristly, referring to the high amount of neurochaetae, particularly spinigers.

Description. Holotype complete, atokous female, 240 mm long, 2.6 mm wide, 130 chaetigers. Body pale, tapered, yellowish (Fig. A). Paratype with brown pigmentation; prostomium with spots in the incision of anterior margin, two spots over the eyes, and a cordiform spot between posterior pair of eyes; tentacular ring and anterior segments with two transverse narrow bands, giving a striate appearance, pigmentation progressively reduced toward posterior end (Fig. 4C). Prostomium wider than long, anterior margin deeply incised; antennae subulate, as long as prostomium, extending slight beyond palps; eyes rounded, anterior pair sometimes reniform, subequal, in trapezoidal arrangement (Figs 4A, C). Tentacular ring twice longer than first chaetiger; four pairs of tentacular cirri, one posterodorsal incomplete, longest one reach chaetiger 16 (Figs 4A, C). Pharynx everted (Fig. A); jaws brown with 8 teeth extended along cutting edge (Fig. 4I). Maxillary ring: I=0, II= 8–7 cones in arc, III= 8 cones in round, IV= 9–10 cones in round. Oral ring: V and VII-VIII bare, VI= 1 cushion-shaped papillae. Parapodial cirri pattern: Dorsal cirri longer than notopodial ligules throughout body, passing greatly all parapodial structures throughout body, medially inserted in anterior-most chaetigers; notopodial ligules progressively reduced toward posterior end, migrating toward basal joint of dorsal cirri. Ventral cirri shorter than neuroacicular lobes in anterior and middle body, becoming longer than neuroacicular ones in posterior body, basally inserted throughout body. First two chaetigers type D, remaining ones biramous. In first two chaetigers (Fig. 4K), dorsal cirri basal, 1.5 times longer than dorsal ligules. Dorsal ligules digitiform, twice longer than neuroacicular lobes. Neuroacicular lobes digitiform, twice longer than postchaetal ones; postchaetal lobes rounded; neuropodial ligules digitiform, 1.2 times longer than neuroacicular lobes. Ventral cirri basal, 4/5 as long as neuropodial ligules.

70

In anterior chaetigers (Fig. 4L), dorsal cirri basal, 3.5 times longer than notopodial ligules. Notopodial ligules subconical, as long as notoacicular lobes; notoacicular lobes subconical, 1.5 times longer than neuroacicular ones. Neuroacicular lobes digitiform, twice longer than postchaetal ones; postchaetal lobes rounded, lamellar; neuropodial ligules subconical, blunt tip, as long as postchaetal lobes. Ventral cirri basal, 4/5 as long as neuropodial ligules. In middle chaetigers (Fig. 4M), dorsal cirri several times longer than notopodial ligules and all parapodial structures. Notopodial ligules subconical, subequal to notoacicular lobes; notoacicular lobes subconical, twice longer than neuroacicular ones. Neuroacicular lobes rounded, subequal to postchaetal ones, 1.5 times longer than neuroacicular ligules; postchaetal lobes rounded, lamellar; neuropodial ligules digitiform. Ventral cirri basal, 1.3 times longer than neuropodial ligules. In posterior chaetigers (Figs 4N, O, P), dorsal cirri several times longer than notopodial ligule and all parapodial structures. Notopodial ligules subconical, becoming linear and very reduced, absent in posterior-most chaetigers (holotype only); notoacicular lobes subconical, very reduced, slightly longer than notopodial ligules. Neuroacicular lobes subconical; postchaetal lobes rounded, lamellar, as long as neuroacicular ones; neuroacicular ligules digitiform, very reduced, medially inserted to neuroacicular lobes. Ventral cirri basal, 1.2 times longer than neuroacicular lobes. Notochaetae sesquigomph spinigers and falcigers. Neurochaetae sesquigomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular fascicles. Notopodial sesquigomph spinigers serrate, teeth decreasing in size toward distal end (Fig. 4H); sesquigomph falcigers unidentate, serrate, teeth increasing in size toward distal end, distal teeth extending beyond distal tooth (Fig. 4D). Neuropodial sesquigomph spinigers as notopodial ones; heterogomph spinigers pectinate, coarse teeth, basal-most ones shorter, increasing rapidly in size, decreasing in size toward distal end, with convex appearance (Fig. 4G). Neuropodial heterogomph falcigers pectinate, distally unidentate throughout body, distal teeth extending beyond distal tooth (Figs 4E, F, I); blades longer and slender in anterior chaetigers (Fig. 4I); supra-acicular falcigers slightly longer and stouter than sub-acicular ones (Figs E, F). Pygidium swollen, funnel-shaped; anal cirri missing (Fig. 4B).

71

Variation. Pigmentation was lost in several specimens, and the striate pattern is shared with other species. The holotype is the largest specimen, being several times larger than remaining material. The absence of notopodial ligules in posterior-most chaetigers in the holotype is attributed to the latter, because all smaller specimens have notopodial ligules in posterior-most segments. Most specimens lost posterior ends, and some of them also lost tentacular or parapodial cirri, because they are dehiscent. Neuropodial chaetae and especially sub-acicular falcigers decreasing greatly their length toward posterior end, but this is common to all species of Ceratonereis studied.

Remarks. Ceratonereis setosa sp. n. resembles C. singularis as noted by Perkins (1980) who preferred not erected a new species due to no more material was available for his study. Specimens studied by him, and deposited in USNM, were examined and they demonstrate to be C. setosa sp .n. Additonal material of C. singularis from Western Mexico (AHF) were examined to corroborate the differences between these species. In C. setosa sp. n., notopodial ligules and neuroacicular lobes are subequal in anterior chaetigers but notopodial ligules are progressively smaller toward posterior chaetigers, becoming papillae- like in posterior-most chaetigers and even disappearing in large specimens; whereas in C. singularis both are subequal throughout body, becoming slightly shorter in posterior-most chaetigers only. On the other hand, in C. singularis the cirrostyle of dorsal cirri are few times longer than notoacicular lobes and the ratio is similar in anterior and posterior chaetigers, while in C. setosa sp. n. the cirrostyles are several times longer than notoacicular lobes and this ratio increases markedly toward posterior chaetigers. In addition, in C. singularis ventral cirri are shorter than neuroacicular lobe throughout body, while in C. setosa sp. n. they are longer than neuroacicular lobes in posterior-most chaetigers. In addition, the blades of notopodial sesquigomph falcigers in C. setosa sp. n are subequal in the same fascicle, whereas in C. singularis, falcigers have different lengths. Normally in the species of Ceratonereis described herein, the neuropodial falcigers from anterior chaetigers have thin, long blades, but decreasing in size and becoming thicker toward posterior segments. This is the case of C. setosa sp. n.; however, in C. singularis, thin, long-bladed falcigers were observed in posterior chaetigers.

72

Finally, the pygidium in C. setosa sp. n. is funnel-shaped and distally swollen, whereas in C. singularis the pygidium is minute, without any projection. Probably most records of C. singularis and some of C. excisa from Grand Caribbean region belong to this new species.

Questionable records

Ceratonereis excisa (Grube & Müller in Grube, 1873)

Nereis excisa Grube 1873: 72–73.

Type locality. Santa Catarina Island (Desterro), Brazil.

Remarks. The authority was modified following the argumentation by Conde-Vela (2013) for Lycastis abiuma and Lycastis littoralis; however names were inverted as noted by Grube in several parts of their work. Perkins (1980) redescribed this species with type material, and he and Hartmann-Schröder (1985) concur that it is restricted to Brazilian coasts. Material identified as C. excisa by Salazar-Vallejo & Jiménez-Cueto (1997) belong to C. gracilis or C. setosa sp. n., and probably other records from Caribbean Sea belong as well.

Ceratonereis mirabilis Kinberg, 1865

Ceratonereis mirabilis Kinberg 1866: 170. Hartman 1948: 71 (partim). Perkins 1980: 4–11, Figs 1–4 (partim).

Type locality. Brazil (09°S), 33 m depth.

Remarks. Hartman (1948:71) revised type material of both C. mirabilis and C. tentaculata Kinberg, 1865, the latter from Honolulu, Hawaii and synonymized them; however, Perkins (1980) reinstated C. tentaculata. Bakken & Wilson (2005) wrongly attributed Galapagos Islands as the type locality of C. mirabilis.

73

The synonymy with Nereis gracilis after Perkins (1980) extended the distribution to Bermuda. However, C. gracilis is herein regarded as a distinct species, while C. mirabilis would be restricted to Southwestern Atlantic coasts. Indeed, both species are very difficult to separate with available data. Further commentaries in remarks section of C. gracilis above.

Ceratonereis singularis Treadwell, 1929

Ceratonereis singularis Treadwell 1929: 1–3, Figs 1–8.

Type locality. San José Island, Gulf of California.

Remarks. Perkins (1980) redescribed this species with type material from the Gulf of California and additional specimens from other localities from Eastern Pacific coasts, but he also included material from the Atlantic, such that he regarded it as amphiamerican. However, his Atlantic specimens rather resemble C. setosa sp. n. as well as the material identified as C. singularis by Salazar-Vallejo & Jiménez-Cueto (1997). Other comments in remarks section of C. setosa sp. n. above.

Composetia Hartmann-Schröder, 1985

Type species. Nereis costae Grube, 1840, by subsequent designation (HartmannSchröder, 1985).

Diagnosis (after Bakken & Wilson 2005). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary ring with conical paragnaths; in areas II-IV, area I with or without paragnaths. Oral ring bare. First two chaetigers type C. Notopodial ligules similar in size throughout body. Notopodial prechaetal lobes present, reduced or absent posteriorly. Neuropodia postchaetal lobe present throughout body. Notochaetae homogomph spinigers; neurochaetae homogomph or sesquigomph spinigers and sesquigomph or heterogomph falcigers in supra-acicular fascicles, heterogomph or homogomph spinigers and heterogomph or sesquigomph falcigers in sub-acicular fascicles.

74

Remarks. Bakken & Wilson (2005) found that this genus is poorly supported in their phylogenetic analysis. Supposedly, the distinctive chaetae for the genus are notopodial homogomph falcigers (Hartmann-Schröder 1985); however, they are not present in the genus (Bakken & Wilson 2005), species described here only have notopodial homogomph spinigers. However, the absence of notoaciculae in chaetigers 1 and 2 and papillae on area VI, prostomium with entire margin, and the distinct parapodial morphology clearly separate Composetia from Ceratonereis.

Composetia versipedata (Ehlers, 1887) Figure 5A–O

Nereis (Ceratonereis) versipedata Ehlers 1887: 116–117, Pl. 36, Figs 5–10. Ceratonereis versipedata Monro 1933a:45–46; 1933b: 256.

Type locality. W Dry Tortugas (24°43'N 83°25'W), Florida Keys, 68 m depth.

Material examined. Northwestern Atlantic Ocean, Bahamas. ECOSUR P0000 (2), R/V Gerda, Cruise 6804, St. 989 (24°07'N 80°11'W), N Cay Sal Bank, Bahamas, 104 m, 6 March 1968. Gulf of Mexico, United States. Topotypes ECOSUR P0000 (2), R/V Gerda, Cruise 6518, St. 570 (24°23'N 82°54'W), S Dry Tortugas, Florida, 101 m, 12 April 1965. Caribbean Sea, Mexico. ECOSUR P0000 (2), R/V Edwin Link, St. EL-2777 (18°26.02'N 87°18.82'W), S Banco Chinchorro, Quintana Roo, 66 m depth, 21 August 1990. Jamaica. ECOSUR P0000 (2), R/V Pillsbury, Cruise 7006, St. 1195 (17°24'N 76°02'W), SE Jamaica, 14 m depth, 3 July 1970. ECOSUR P0000 (2), R/V Pillsbury, Cruise 7006, St. 1252 (17°09'N 78°57'W), W Pedro Bank, Jamaica, 26 m depth, 14 July 1970. Nicaragua. ECOSUR P0000 (2), R/V Pillsbury, Cruise 7101, St. 1339 (13°04'N 82°24'W), E Nicaragua, 33 m depth, 29 January 1971. Puerto Rico. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 962 (17°29'N 65°54'W), SE Puerto Rico, 23 m depth, 19 July 1969. Antigua and Barbuda. ECOSUR P0000 (19), R/V Pillsbury, Cruise 6802, St. 975 (17°29'N 61°55'W), SW Barbuda, Lesser Antilles, 29 m depth, 21 July 1969. ECOSUR P0000 (6), R/V Pillsbury, Cruise 6806, St. 968 (17°18'N 61°52'W), N Antigua, Lesser Antilles, 18 m

75 depth, 20 July 1969. Saint Kitts and Nevis. ECOSUR P0000 (1) R/V Pillsbury, Cruise 6907, St. 961 (17°27'N 62°51'W), NW St. Christopher, Lesser Antilles, 11 m depth, 19 July 1969. Saint Vincent and the Grenadines. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6907, St. 875 (13°10'N 61°05'W), SE St. Vincent, Lesser Antilles, 134 m depth, 6 July 1969. Grenade. ECOSUR P0000 (3), R/V Pillsbury, Cruise 6907, St. 857 (12°23'N 61°22'W), SE Carriacou, Lesser Antilles, 178 m depth, 3 July 1969. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 853 (11°55'N 61°43'W), S Grenada, Lesser Antilles, 17 m depth, 3 July 1969.

Description. Material from Dry Tortugas incomplete, in poor condition, paragnaths faded; remaining material from other localities in variable condition. All without pigmentation, some with body reddish, with body translucent toward posterior end (Fig. 5B). Specimen from Bahamas complete, fairly preserved, middle body damaged, 46 mm long, 2.4 mm wide, 62 chaetigers, tapered posteriorly. In midbody and posterior chaetigers, glandular masses very conspicuous in ligules and lobes, and in dorsal side of each segment (Fig. 5B). Prostomium as long as wide, anterior margin entire; antennae cirriform, half as long as prostomium, not passing the palps; eyes rounded, subequal, in trapezoidal arrangement (Fig. 5A). Tentacular ring 1.6 times longer than first chaetiger; four pairs of tentacular cirri, cirrophores conspicuous, longest one reaches chaetiger 3 (Fig.5 A). Pharynx everted; jaws brown with 7 blunt teeth, restricted to distal half (Fig. 5E). Maxillary ring: I= 0, II= 8–8 cones in two rows, III= 4 cones, IV= 16–16 cones in round. Oral ring bare (Figs 5C, D). Parapodial cirri pattern: Dorsal cirri longer than notopodial ligules in most-anterior segments, becoming slightly longer toward posterior end, passing notopodial ligules throughout body, basally inserted in anterior segments, displaced medially throughout body. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 5F), dorsal cirri basal, 1.4 longer than notopodial ligules. Notopodial ligules subconical, blunt tip, subequal to neuropodial postchaetal lobes. Neuroacicular lobes subconical, blunt tip, half as long as postchaetal ones; postchaetal lobes digitiform, longer than wide; neuropodial ligules digitiform, 1.5 times wider than postchaetal lobes. Ventral cirri basal, as long as neuropodial ligules, half as long as dorsal cirris.

76

In anterior chaetigers (Fig. 5G), dorsal cirri medial, as long as notopodial ligules. Notopodial ligules subconical, 1.5 times longer and 2.5 wider than notoacicular lobes; notoacicular lobes rounded, twice longer than prechaetal ones; prechaetal lobes digitiform. Neuroacicular lobes subconical, blunt tip, as wide as notoacicular ones; postchaetal lobes subconical, 1.2 times longer than neuroacicular ones; neuropodial ligules digitiform, subequal to neuroacicular lobes. Ventral cirri basal, 0.7 times as long as neuropodial ligules, 0.6 times as long as dorsal cirri. In middle and posterior chaetigers (Figs 5H, I), dorsal cirri medial, as long as notopodial ligules, extending twice times beyond tip of the same. Notopodial ligules subconical, 1.0–1.3 times longer and twice wider than notoacicular lobes; notoacicular lobes subconical, 2.5 times longer than prechaetal lobes; prechaetal lobes rounded. Neuroacicular lobes rounded to subconical with blunt tip, 1.5–2 times wider than notoacicular ones; postchaetal lobes subconical, half as long as neuroacicular ones; neuropodial ligules subconical, as long as neuroacicular lobes. Ventral cirri basal, half as long as neuropodial ligules, 0.4 times as long as dorsal cirri. In posterior-most chaetigers (Fig. 5J), dorsal cirri medial, as long as notopodial ligules. Notopodial ligules subconical, 1.5 longer and 2.5 times wider than notoacicular lobes; notoacicular lobes subconical, 1.7 times longer than neuroacicular lobes; prechaetal lobes papillae-like. Neuroacicular lobes subconical, blunt tip, as wide as notoacicular ones; postchaetal lobes subconical, half as long as neuroacicular ones; neuropodial ligules subconical, as long as neuroacicular lobes. Ventral cirri basal, as long as neuropodial ligule, 0.5 times length of dorsal cirri. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers and heterogomph falcigers in both supra- and sub-acicular fascicles. Notopodial homogomph spinigers serrate, teeth decreasing in size toward distal end. Neuropodial homogomph spinigers as notopodial ones (Fig. 5O). Heterogomph falcigers pectinate, fine teeth, supra-acicular (Figs 5K, M) stouter than sub-acicular ones (Figs 5L, N); blades with falcate tips, one-sixth of the margin edentulate (Figs 5K, L), becoming shorter toward posterior chaetigers, with blades weakly falcate, half of the margin edentulate (Figs 5M, N). Pygidium crenulated; anal cirri cirriform, as long as last six segments (Fig. 5B).

Variation. Material examined show slight divergences from the above description. The species is easily recognized by the distinctive pattern of glandular masses on dorsal side of chaetigers and

77 parapodia and translucent body toward posterior end. The arrangement of paragnaths and their shape vary little, only in specimens from Grenade (ECOSUR P0000) they are conical with pointed tips. Notopodial prechaetal lobes are slightly longer than represented in Figure G in some specimens, especially after first two chaetigers, diminishing their length rapidly.

Remarks. Composetia versipedata have several differences that separate it from the other species Northwestern Atlantic species, C. irritabilis. In C. versipedata, paragnaths on area III are arranged in a round, while in C. irritabilis they are arranged in a large rectangle, almost reaching areas III. Hartman (1938a) argued that in the type of C. versipedata paragnaths of area III are arranged as in C. irritabilis, but it requires confirmation. In addition, in C. versipedata, the dorsal cirri are longer than notopodial ligules throughout body, while in C. irritabilis they are shorter throughout body. Furthermore, jaws of C. versipedata have sharp teeth, while in C. irritabilis they are blunt. Hartman (1938a) considered this species as synonym of C. irritabilis after the revision of type material. However, she changed her mind and in another publication (Hartman 1945) she did not include C. versipedata as synonym and even highlighted differences between them. Also, Hartman in the same work and later Fauchald (1977a) argued that the report by Monro (1933a) from Panama could not belong to C. versipedata, probably being a undescribed species. Monro stated in the text that their specimens have two notopodial processes only in anterior parapodia, although he drawn a prechaetal (Monro 1933a: text fig. 19); here his report is considered as valid because other described features match with actual description.

Distribution. Bahamas, Caribbean Sea.

Composetia brasiliensis (McIntosh, 1885)

Nereis (Ceratonereis) brasiliensis McIntosh 1885: 230–231, Pl. 36, Figs 1–3, Pl. 17a, Figs 3,4. Ceratonereis brasiliensis Salazar-Vallejo & Jiménez-Cueto 1997: 363. Composetia brasiliensis Hartmann-Schröder 1985: 49. Ceratonereis hircinicola Santos & Lana 2003: 8–9, Figs 13–18 (non Eisig, 1870).

78

Type locality. Barra Grande (9°9’S 34°53’W), Brazil, at 59 m depth.

Remarks. Hartmann-Schröder (1985) placed this to Composetia because the entire margin of prostomium present in the species. The specimen illustrated by Santos & Lana (2003) as Ceratonereis hircinicola from Paraíba is identical to that described and illustrated by McIntosh (1885). Also, material reported from off Salvador, Brazil by Rullier & Amoureux (1979) might belongs as well. Unfortunately material for a redescription was not available; Salazar-Vallejo & Jiménez-Cueto (1997) revised type material and included some features, they considered unnecessary a redescription, however it is required to avoid future confusions. This species has not been recorded from other locality.

Questionable records

Composetia irritabilis (Webster, 1879) Figure 6A–N Nereis irritabilis Webster 1879: 231–234, Pl. 5, Figs 56–64, Pl. 6, Figs 65–69. Ceratonereis irritabilis Hartman 1938b: 13–15. Composetia irritabilis Bakken & Wilson 2005: 521–522.

Type locality. Virginia, Northwestern Atlantic Ocean.

Type material. Northwestern Atlantic Ocean, United States. Syntypes of Nereis irritabilis USNM 531 (2), 532 (3), 533 (8) and 534 (4), Virginia. (Data in original publication: Northampton Co., Va, 21 July for atoke females and 1-12 August for swarming epitokes, 1874 to 1876).

Description. Syntypes consist in epitokes and atokes specimens, all in good conditions. Atokes (USNM 531) complete, in fragments, partially dehydrated, jaws used for illustration. One specimen (USNM 533) male complete, 54 mm long, 1.7 wide, 146 chaetigers. Female specimens (USNM 532) two incomplete, complete used for description. Body pale, pigmentation in palpophores only, on lateral edges, glands green in parapodia along body (Figs 6A, D–F).

79

Prostomium as long as wide, anterior margin entire; antennae cirriform, 1/3 or 1/4 as long as the prostomium, not passing the palps; eyes black, rounded, subequal (Figs 6A, D). Tentacular ring 1.5 to 2 times longer than first chaetiger; four pairs of tentacular cirri, longest one reach chaetigers 11 in female and 15 in male (Figs 6A, D). Pharynx everted in male and atoke; jaws of atoke dark brown, 10-12 teeth with incisions very shallow (Fig. C). Maxillary ring: I= 0, II= 8-9 in round; III= 34 in a large rectangle; IV= 10-11 in round. Oral ring bare (Figs 6A, B, D). Parapodial cirri pattern. Dorsal cirri shorter than notopodial ligule throughout body, basally inserted in most-anterior chaetigers, displaced medially throughout body. Ventral cirri basal and shorter than neuropodial ligule throughout body. In male epitokes, anterior chaetigers with dorsal cirri modified in chaetigers 1–7 in males, ventral cirri modified in chaetigers 1–5 in males; females epitokes without modified parapodial cirri throughout body. Body of male and females divided into three regions. Pre-natatory region includes chaetigers 1– 27 in male and 1–31 in females, natatory region includes chaetigers 28–128 in male and 32–107 in female, post-natatory region from chaetiger 128 to end of body in males and from chaetiger 108 to end of body in females. First two chaetigers type C, remaining biramous. In first two chaetigers (Figs 6G, K), dorsal cirri basal, 3/4 as long as notopodial ligules. Notopodial ligules subconical, subequal to neuropodial ones. Neuroacicular lobes subconical, blunt tip, half as long as notopodial ligules; postchaetal lobes digitiform, as long as neuroacicular ones; neuropodial ligules subconical. Ventral cirri basal, half as long as neuropodial ligules. In chaetigers 2–7 in males (Fig. 6H), dorsal cirri basal, vertically positioned, increasing their length posteriorly. Notopodial ligules subconical, dorsally enlarged, as long as notoacicular lobes; notoacicular lobes subconical, blunt tip, twice longer than neuroacicular ones; prechaetal lobes papillae-like. Neuroacicular lobes subconical, blunt tip, distally bilobated, superior lobes digitiform and twice longer than globose inferior ones; postchaetal lobes digitiform, as long as neuroacicular ones; neuropodial ligules subconical, blunt tip, as long as neuroacicular lobes. Ventral cirri 3/4 the length of neuropodial ligules. Chaetigers 8–27 in males and 1–31 in females (Fig. 6L), dorsal cirri medial, cirriform, 3/4 as long as notopodial ligule. Notopodial ligules subconical; remaining structures as in chaetigers 2–7 in males.

80

In chaetigers from natatory region modified (Figs 6J, M), dorsal cirri medial, half as long as notopodial ligules; basal lamellae rounded, as long as dorsal cirri in males, smaller in females, progressively shorter toward posterior chaetigers. Notopodial ligules and notoacicular lobes subequal, subconical, blunt tip; both with with surrounding lamellae, more developed in females. Neuroacicular lobes difitiform, with surrounding flabellate lamellae, more developed in males; neuropodial ligules digitiform. Ventral cirri half as long as neuroacicular lobe; male with basal lamellae, progressively shorter toward posterior chaetigers. In chaetigers from post-natatory region (Figs 6I, N), dorsal cirri medial, half as long as notopodial ligules. Notopodial ligules subconical, as long as notoacicular lobes; notoacicular lobes subconical, twice longer than neuroacicular lobes, prechaetal lobes inconspicuous. Neuroacicular lobes subconical, blunt tip; postchaetal lobes digitiform, as long as neuroacicular ones; neuropodial ligules subconical, 1.5 times longer than neuroacicular lobes. Ventral cirri basal, half as long as neuropodial ligules. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers and sesquigomph falcigers in both supra- and sub-acicular fascicles. Anterior chaetigers with homogomph spinigers only. Sesquigomph falcigers appearing in middle chaetigers and disappearing in most-posterior ones. Pygidium with anal papillae in males (Fig. 6E), slightly crenulated in females (Fig. 6F); anal cirri cirriform, as long as three segments (Figs 6E, F).

Remarks. Description provided here and that performed by Bakken & Wilson (2005) are in agreement. Webster (1879) included further data about the collecting of specimens and living observations. The dorsal cirri in males are unusually bare instead of crenulated as in epitokal transformation of most genera. The sucession of chaetae is alo unusual among nereidids. For further comments about synonymy with C. versipedata see remarks section of this species above. Reports of this species from Grand Caribbean region (Dean 2012) not included descriptions or illustrations to ensure their presence in the region, requiring reassessment.

Distribution. Virginia, United States.

81

Composetia hircinicola (Eisig, 1870)

Nereis hircinicola Eisig 1870: 103–104, Pl. 11, Figs 3–4.

Type locality. Portopí, Palma de Mallorca, Balearic Islands, Mediterranean Sea.

Remarks. The species was confused with the regional species C. brasiliensis, see remarks section of this species above.

Species requiring further revision

Composetia articulata (Ehlers, 1887)

Nereis articulata Ehlers 1887: 114–116, Pl. 36, Figs 1–4. Composetia articulata Hartmann-Schröder 1985: 49.

Type locality. Off Sand Key, Florida, at 220 m depth.

Remarks. Hartman (1938a) noted the poorly condition of the holotype (MCZ ANNa-746 in online database), and without pharynx, it is difficult to assess. Augener (1906), based upon an incomplete specimen from St. Vincent, tried to improve the original description, but their belonging to this species cannot be ensured with the provided data. The description by Ehlers is sufficient to regard this species as distinct, but it requires the revision of type material and additional specimens to ensure its validity.

Hediste Malmgren, 1867

Type species. Nereis diversicolor O.F. Müller, 1776, by original designation (Malmgren 1867).

82

Diagnosis (after Bakken & Wilson 2005). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary ring with conical paragnaths in areas II-IV, area I with or without paragnaths. Oral ring with conical paragnaths in areas VI-VIII, area V bare. First two chaetigers type C. Notopodial ligules similar in size throughout body. Notopodial prechaetal lobes present, reduced or absent posteriorly. Neuropodial postchaetal lobe present or absent. Notochaetae homogomph spinigers; neurochaetae homogomph spinigers, heterogomph and ankylosed falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular fascicles.

Remarks. The species present to this genus are very close morphologically, but they have different reproductive patterns that allow recognize species (Sato & Nakashima 2003). Some studies suggest that the species H. diversicolor has cryptic populations worldwide distributed, and even some mechanisms for their introduction have been proposed (Cossu et al. 2012; Einfeldt et al. 2014; Virgilio et al. 2009).

Questionable records

Hediste diversicolor (O.F. Müller, 1776)

Nereis diversicolor O.F. Müller 1776: 217.

Type locality. Southern Norway.

Remarks. The name was introduced for the Grand Caribbean in some reports (Dean 2012). The report by Hoagland (1919) was based upon paragnaths only, and the description by Treadwell (1939) was based on McIntosh (1910), and referring it to the report by Hoagland (1919). This species has a confused, problematic taxonomic history; problems must be solved to clarify the taxonomic status of synonymized species and the possible proposal of new species for the Grand Caribbean and others regions.

83

Leonnates Kinberg, 1865

Type species. Leonnates indicus Kinberg 1865, by monotypy.

Diagnosis (modified from Qiu & Qian 2000). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary ring with conical paragnaths, rarely papillae, in areas II-IV, area I with or without paragnaths or papillae. Oral ring with papillae, rarely conical paragnaths in areas VI–VIII, area V with or without paragnaths or papillae. First two chaetigers type C. Notopodial ligules similar in size throughout body. Prechaetal notopodial lobes generally present; neuropodial postchaetal lobe present throughout body. Notochaetae homogomph spinigers and rarely homogomph falcigers; neurochaetae homogomph spinigers, sesquigomph and heterogomph falcigers.

Remarks. The genus has been recently revised by Qiu & Qian (2000), most of the species are distributed in the Pacific Ocean. The authors redefined the genus based on the revision of the type species, L. indicus. However, they regarded as homogomph to neuropodial falcigers with a large tooth, instead of sesquigomph ones. Based upon the SEM illustrations, these falcigers match with the definition of sesquigomph chaetae adopted here, and therefore this feature was included in the diagnosis. Furthermore, the accurate composition of neuropodial supra- and sub-acicular fascicles was not provided by the authors, and therefore excluded in the diagnosis. On the other hand, Qiu & Quian (2000) included some species referred to Paraleonnates, but Paraleonnates guadalupensis Amoureux, 1985 from Caribbean Sea was excluded. P. guadalupensis was synonymized with Leonnates by Hong et al. (2012).

Leonnates guadalupensis (Amoureux, 1985)

Paraleonnates guadalupensis Amoureux 1985: 96–98, Figs 2A–C. Leonnates guadalupensis Hong et al. 2012: 60.

Type locality. Lagoon of Manche-a-Eau, Guadeloupe, Lesser Antilles.

84

Remarks. Hong et al. (2012) stated that Paraleonnates Amoureux, 1985 is a junior homonym to Paraleonnates Khlebovich & Wu, 1962, and transferred this species to Leonnates with some reserves. N. goajirana resembles to this species in having notopodial prechaetal or postchaetal lobes and the shape of blades of neuropodial falcigers, however they differ in length of tentacular cirri and paragnaths number. They might be congeneric; unfortunately, material for a redescription was not available. Amoureux (1985) proposed this genus due to the presence of paragnaths on oral ring, which are supossedly absent in Leonnates species; however, Qiu & Qian (2000) emended the diagnosis after L. simplex has paragnaths on areas VI. Also, Amoureux noted small papillae strongly colored by Bengal rose on areas VII-VIII, and the neuropodial falcigers above mentioned resemble those found in L. indicus and L. niestraszi (Qiu & Qian 2000). The parapodial features match with Leonnates, but this species requires revision for clarifying its generic affinity.

Micronereis Claparède, 1863

Type species. Micronereis variegata Claparède, 1863, by monotypy.

Diagnosis (after Paxton 1983). Prostomium with anterior margin entire. Antennae absent. Eyes present. Four pairs of tentacular cirri. Maxillary and oral rings with paragnaths, not arranged into discrete areas, often tiny or non-visible. First two chaetigers type A. Notopodial and neuropodial ligules absent throughout body. Notoacicular and neuroacicular lobes clearly divergent, often with ciliate surfaces. Parapodial accessory cirri present or absent. Notochaetae homogomph spinigers; neurochaetae homogomph spinigers and falcigers (?). Epitokes dimorphic, with simple capillary chaetae and sesquigomph spinigers absent or present.

Remarks. The diagnosis was formed from the original revision of the genus by Paxton (1983). Paxton (1983) noted that digitiform structures on notoacicular and neuroacicular lobes, found in males of some species, are a sexual dimorphic feature. In addition, she discussed that these structures cannot be regarded as ligules, retaining the term accessory parapodial cirri used by other authors, and this point of view is accepted in this study. Furthermore, Paxton (1983)

85 highlighted the absence of fascicles in both parapodial rami, arranged in 'fan-shaped bundles', the rare presence of sesquigomph spinigers and the simple chaetae in some epitokes. The accessory parapodial cirri, sesquigomph spinigers and capillary chaetae were included in the diagnosis. Presence of falcigers was regarded as a doubtful because their presence in the genus need to be reassessed, further comments in remarks of M. piccola below. The phylogenetic position of Micronereis is, as other genera, regarded as doubtful. Santos et al. (2005) weakly corroborated the monophyly of subfamily Notophycinae Knox & Cameron, 1970, originally erected for the genus Micronereis and related ones (including the family Notophycidae). The genus is represented with one species in the Grand Caribbean, but the presence of more species is suspected.

Micronereis piccola Paxton, 1983 Figure 7A–I

Micronereis piccola Paxton 1983: 7–8, Figs 3–10.

Type locality. Cherokee Sound, Abaco, Bahamas (26°N 77°W).

Type material Northwestern Atlantic Ocean, Bahamas. Holotype USNM 54514, Cherokee Sound, Abaco Island, 18 March 1972, Coll. A. Schoener. Paratypes USNM 60505 (1) and USNM 60506 (1), Cherokee Sound, Abaco Island, 10 April 1972, Coll. A. Schoener. Paratypes USNM 60507 (1) and USNM 60508 (1), Cherokee Sound, Abaco Island, 12 March 1972, Coll. A. Schoener.

Additional material. Caribbean Sea, British Virgin Islands. LACM-AHF 7380 (1 female), White Bay, Guana Island (18°28’21”N, 62°34’26”W), 0.5 m depth, on filamentous green algae ver sand, 16 July 2001, Coll. K. Fitzhugh. LACM-AHF 7386 (1 female), White Bay, Guana Island (18°28’26.97”N 64°34’35.04”W), 1 m depth, in imxed algal turf, 6 July 2000, Coll. K. Fitzhugh. LACM-AHF 7387 (1 female), Long Point, Guana Island (18°29.153’N 64°34.971’W), 1-2 m depth, from larga barnacles over grown with algae, tunicates, sponges, 19 July 2000, Coll. T. Zimmerman, T. Haney, J. Martin. LACM-AHF 7389 (1 female), Grand Ghut, Guana Island

86

(18°28’47”N, 64°33’47”W), on alga from hard substrate, 18 m depth, 6 July 2000, Coll. G. Hendler.

Description. Holotype male complete, 1.8 mm long, 0.5 mm wide, 16 chaetigers (Fig. 7A). Non- type females with oocytes, two per chaetiger (Fig. 7C); female (LACM-AHF 7387) complete, 2 mm long, 0.7 mm wide, 18 chaetigers, dissected for parapodia. Body pale, without pigmentation (Fig. 7A). Prostomium ovate, anterior margin entire; antennae absent; eyes rounded, anterior pair larger than posterior one, in trapezoidal arrangement (Fig. 7B). Tentacular ring surrounding prostomium; four pairs of tentacular cirri, longest reach chaetiger 4 (Fig. 7A, B). Pharynx observed by transparence. Female jaws subtriangular, four teeth restricted to distal end (Fig. 7D); male jaws large, sinuose, without dentate base (Fig. 7E). Parapodial ciri pattern: dorsal and ventral cirri shorter than and subdistally inserted to notoacicular lobes in anterior chaetigers, becoming subequal in posterior ones. First two chaetigers type A, remaining biramous. In first two chaetigers (Fig. 7G), notopodial structures absent. Neuroacicular lobes globose, 3 to 4 times longer than ventral cirri. Ventral cirri medially inserted to neuroacicular lobes. In remaining chaetigers (Figs 7H, I), dorsal cirri subdistal, subconical, blunt tip, half as long as notoacicular lobes in anterior chaetigers, subequal posteriorly. Notoacicular and neuroacicular lobes digitiform, distal end swollen, becoming wider toward posterior chaetigers. Ventral cirri subequal to dorsal ones throughout body. Chaetae homogomph spinigers in notoacicular and neuroacicular lobes throughout body (Fig. 7F).

Remarks. This description mostly matches the one made by Paxton, but some features deserve comments. The distal end of jaws was detailed, and differs little as figured by Paxton (1983). In M. piccola stylet-like jaws are common in males of the genus but male jaws from other species have a wider, dentate base (Paxton 1980), which is not present in M. piccola. The ‘falcigers’ illustrated by the same are spinigers as well. Some blades of spinigers are damaged in their medial or distal portions (Paxton 1983, Fig. 10), and when blades are observed at 40x magnification, they seem to have a rounded, spoon-like tip. However, at higher

87 magnification, blades are actually damaged and the falciger illustrated have a bifid tip; true falcigers were not found. Other 'falciger' was illustrated (Paxton 1980, Fig. 32) with a spoon-like tip; however, it seems to be an artifact caused by the recurvation of the tapered tip, sometimes a consequence of being mounted in a temporary slide. There are other Micronereis species described as having falcigers, M. bansei and M. halei, but because of the above reasons, they must be reassessed to discard possible artifacts.

Distribution. Bahamas, British Virgin Islands.

Neanthes Kinberg, 1865

Type species. Neanthes vaali Kinberg, 1865, by subsequent designation (Augener, 1922).

Diagnosis (after Glasby et al. 2011). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary ring with conical or smooth-bar paragnaths; present or absent in areas I-IV. Oral ring with conical paragnaths, sometimes arising from plate-like basements in areas V–VIII, sometimes missing. First two chaetigers type C. Notopodial ligules with different development types, present throughout body. Prechaetal notopodial lobe present or absent; neuropodial postchaetal lobe present throughout body. Notochaetae homogomph spinigers; neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph falcigers and spinigers in sub-acicular fascicles. Ankylosed falcigers in both neuropodial fascicles present or absent.

Remarks. The genus, as occurs in other genera, has merged several body patterns, and as result, it has high levels of homoplasy in phylogenetic analyses (Bakken & Wilson 2005, Santos et al. 2005). Species can have or not notopodial prechaetal lobes, or the notopodial ligules can be clearly widened of foliose toward posterior chaetigers, and even ankylosed, simple chaetae can be present. Other analyses are required for restriction of Neanthes and the possible restablishement or even the proposition of new genera (Glasby et al. 2011).

88

Neanthes galetae Fauchald, 1977 Figure 8A–K

Neanthes galetae Fauchald 1977: 26–27, Figs 6a–c.

Type locality. Galeta Island Reef, Panama, Caribbean Sea.

Type material. Caribbean Sea, Panama. Paratypes LACM-AHF 1132 (3), Isla Galeta, Galeta Reef, Colón (9°24'18"N 79°51'48.5"W), 23 October 1970, Coll. STRI.

Description. Paratypes in good conditions; one complete, 14 mm long, 0.8 mm wide, 57 chaetigers. Glands dark blue arranged in a thin transverse band per segment throughout body (probably after a previous treatment with methy-green under acidic conditions). Glands present on distal portion of prostomium, on palpophores and tentacular ring (Fig.8B); a line crossing the dorsum on each chaetiger, discoloring toward posterior end, and additional spots on the base of each parapodia and in ligules and lobes (Figs 8A–C). Prostomium longer than wide; antennae cirriform, half as long as prostomium; eyes rounded, subequal, in rectangular arrangement (Fig. 8B). Tentacular ring 1.5 times longer than first chaetiger; four pairs of tentacular cirri, longest reach chaetiger 2. Pharynx dissected. Maxillary ring: I=1 big cone, II=10-14 in curve, III= 8 cones and four rows of merged cones, IV=10-9 in a curve. Oral ring: V=0, VI=10-12 in round, VII-VIII= 9 in a row, small cones alternating with big ones. Parapodial cirri pattern: dorsal cirri shorter than notopodia ligules in anterior and middle chaetigers, becoming as long as in posterior chaetigers, basally inserted in anterior-most chaetigers, displacing medially and passing the notopodial ligules toward posterior chaetigers. Ventral cirri shorter and basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 8H), dorsal cirri basal, as long as notopodial ligules. Notopodial ligules digitiform, as long as neuropodial ligules. Neuroacicular lobes subconical, blunt tip; postchaetal lobes digitiform, twice longer than neuroacicular ones; neuropodial ligules digitiform, twice longer than neuroacicular lobes.

89

In anterior chaetigers (Fig. 8I), dorsal cirri medial, 3/4 as long as notopodial ligules. Notopodial ligules digitiform to rounded, slightly shorter than notoacicular lobes; notoacicular lobes rounded, twice wider than notopodial ligules and neuroacicular lobes. Neuroacicular ligules subconical to rounded, half as long as postchaetal lobes; postchaetal lobes digitiform; neuropodial ligules digitiform, as long as neuroacicular lobes. Ventral cirri basal, 3/4 as long as neuropodial ligules. In middle and posterior chaetigers (Figs 8J, K), dorsal cirri medial, 3/4 as long as notopodial ligules, becoming as long as those present in posterior chaetigers, distal end becoming narrow toward posterior chaetigers. Notopodial ligules subconical, blunt tip, becoming wide at the base and narrow at the tip in posterior chaetigers, subequal to notoacicular lobes; notoacicular lobes subconical, blunt tip, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip; postchaetal lobe digitiform, subequal to neuroacicular ones, inconspicuous in posterior chaetigers; neuropodial ligules digitiform, as long as neuroacicular lobes. Ventral cirri basal, 3/4 as long as neuropodial ligules. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles. Notopodial homogomph spinigers pectinate, fine teeth, decreasing in size toward distal end (Fig. 8F). Neuropodial homogomph spinigers as notopodial ones (Fig. 8G); heterogomph spinigers pectinate, coarse teeth, decreasing in size toward distal end. Heterogomph falcigers pectinate, fine and long teeth, distal teeth stout, with a fine tendon fused to margin; supra-acicular falcigers with 1/3 of inner margin edentate, sub-acicular ones with 1/5 edentate (Figs 8D, E). Pigydium tripartite; anal cirri subulate, as long as last three chaetigers (Fig. 8C).

Remarks. This description matches the one made by Fauchald (1977a); however, the description is accompanied with few figures and some features were not detailed. The paratype has a very conspicuous glandular pattern, the glands are dark green and probably they are stained with green-methyl, but if it was performed previously is unkown. N. galetae is easily separated from other Grand Caribbean species by having a few paragnaths on areas VII-VIII and by the parapodial topology.

90

Distribution. Curently restricted to type locality, Galeta Reef, Caribbean Panama.

Neanthes micromma (Harper, 1979) Figure 9A–K

Nereis (Neanthes) micromma Harper 1979: 92–95, Figs 1a–c, 2–6, 7a–e, 8a–c, 9 a–c, 10–11. Taylor 1984: 31-17, Figs 31-14a–h.

Type locality. Brazos-Colorado River Delta, Freeport, Texas, Gulf of Mexico (28°44.5’N 95°13.2’W), 21 m.

Type material. Gulf of Mexico, United States. Holotype of Nereis (Neanthes) micromma USNM 55575, Freeport, Texas, 2 December 1977, Coll. D.E. Harper. Paratypes USNM 55576 (15), Freeport, Texas, 2 December 1977, Coll. D.E. Harper.

Description. Type specimens in good condition. Holotype complete, 83 mm long, 1.1 mm wide, 256 chaetigers. Body long, tapered, pale, pigmentation absent. Paratypes numbered, paratype #6 dissected and used for description; specimen atokous female, incomplete posteriorly, 43 mm long, 1 mm wide, 124 chaetigers, body pale, pigmentation absent. Prostomium anterior margin entire; antennae subulate, 1/3 as long as prostomium, not passing the palps; palps divergent; eyes black, small, in rectangular arrangement (Fig. 9A). Tentacular ring 1.5 times longer than first chaetiger; four pairs of tentacular cirri, largest reach chaetiger 3 (Fig. 9A). Pharynx partially everted, paragnaths almost faded, very difficult to observe and count. Parapodial cirri pattern: Dorsal cirri shorter than notopodial ligules throughout body, basally inserted in anterior-most chaetigers, displaced medially in medial chaetigers, subdistally in posterior-most ones. Ventral cirri shorter than neuropodial ligules throughout body, basally inserted in anterior-most region, becoming distal toward posterior chaetigers. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 9H), dorsal cirri basal, 3/4 as long as dorsal ligules. Dorsal ligules subconical, longer than wide, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, as long as postchaetal ones;

91 postchaetal lobes digitiform; neuropodial ligules subconical, longer than wide, longer than all parapodial structures. Ventral cirri basal, 3/4 as long as neuropodial ligules. In anterior chaetigers (Fig. 9I), dorsal cirri medial, half as long as notopodial ligules. Notopodial ligules subconical, longer than wide, half as long as notoacicular lobes; notoacicular lobes subconical, longer than wide, twice longer than neuroacicular ones. Neuroacicular lobes subconical, blunt tip; postchaetal lobes reduced, button-shaped; neuropodial ligules subconical, longer than wide. Ventral cirri basal, half as long as neuropodial ligules. In middle chaetigers (Fig. 9J), dorsal cirri subdistal, half as long as notopodial ligules. Notopodial ligules subconical, wider than long, half as long as notoacicular lobes; notoacicular lobes subconical, longer than wide, 1.3 times longer than neuroacicular lobes. Neuroacicular lobes subconical, as long as neuropodial ligules; postchaetal lobes disappearing since chaetiger 26; neuropodial ligules subconical. Ventral cirri basal, 3/4 as long as neuropodial ligules. In posterior chaetigers (Fig. 9K), dorsal cirri subdistal, half as long as notopodial ligules. Notopodial ligules pennant-like, foliose, wider than long, as long as notoacicular lobes; notoacicular lobes subconical, longer than wide, as long as neuroacicular lobes. Neuroacicular lobes subconical, as long as neuropodial ligules; postchaetal lobes inconspicuous; neuropodial ligules subconical. Ventral cirri basal, 3/4 as long as neuropodial ligules. Notochaetae homogomph spinigers. Neurochaetae homogomph and heterogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph falcigers and spiniger in sub- acicular ones. Homogomph spinigers pectinate, coarse teeth, decreasing in size toward distal end (Fig. 9D). Neuropodial homogomph and heterogomph spinigers with dentition as in notopodial ones; blades of spinigers decreasing their lengths toward lower-most position in fascicles (Figs C, E). Heterogomph falcigers serrate to pectinate; blades of falcigers from anterior chaetigers with coarse teeth and blunt tips (Fig. 9F), becoming with fine teeth and acute tips toward posterior chaetigers (Fig. 9G). Pygidium crenulated; anal cirri short, half as long as pygidium width (Fig. 9B).

Variation. Chaetal blades have different lengths along body and in the same parapodium, as noted by Harper. Notopodial ligules become wider in posterior-most chaetigers in complete holotype (256 chaetigers) (Fig. 9B), than posterior parapodium showed (chaetiger 120) (Fig. 9K).

92

Remarks. The species has been characterized by Harper (1979); Taylor (1984) noted that N. micromma was often confused with Nicon, Ceratocephale, Ceratonereis and Nereis. Here, some features are included. The species have been reported from Pacific coasts, by Hernández- Alcántara & Solís-Weiss (1991) and Dean (2001); the latter included a detailed description that does not match with N. micromma, being an undescribed species.

Distribution. Gulf of Mexico.

Neanthes 1 sp. n. Figure 10A–K

Neanthes acuminata Liñero-Arana & Díaz-Díaz 2007: 158-159, Figs 4A–F (partim, non Ehlers, 1868)

Type material. Caribbean Sea, Mexico. Holotype ECOSUR 0000, and paratype ECOSUR 0000 (1), Isla Contoy, Quintana Roo, 22 February 2008, on mangrove roots, Coll. S.I. Salazar- Vallejo, L.F. Carrera-Parra. Jamaica. Paratype UMML P0000 (1), R/V Pillsbury, Cruise 7006, St. 1214 (17°36'N 77°03'W), S Jamaica, 24 m depth, 05 July 1970.

Additional material examined. Northwestern Atlantic Ocean, Bahamas. Neanthes cf. 1 sp. n. ECOSUR P0000 (1), R/V Gerda, Cruise 6429, St. 354 (25°41'N 79°31'W), S Bimini, 817 m depth, 24 August 1964.

Type locality. Isla Contoy, Quintana Roo, Mexico.

Description. Holotype incomplete posteriorly, 11 mm long, 1.2 mm wide, 34 chaetigers. Body pale, without pigmentation, tapered. Paratype (ECOSUR P0000) with everted pharynx used for paragnaths number. Prostomium as long as wide, anterior margin entire; antennae subulate, 1/3 as long as prostomium; eyes oval, anterior pair larger than posterior one, in trapezoidal arrangement (Fig.

93

10A). Tentacular ring 1.5 times longer than first chaetiger; four pairs of tentacular cirri, longest one reaches chaetiger 5 (Fig. 10A). Pharynx partially everted in holotype, fully everted in paratype; jaws amber, 10 teeth, eroded. Maxillary ring: I= 9 cones in round, II=18-17 cones in arc, III= 30 in rectangle, IV= 27-32 cones in arc. Oral ring: V–VII= a broad, continuous band of cones, ventral side twice broader than dorsal one, with variable arrangement and paragnaths number; areas VI generally with a paragnath larger than surrounding ones (Fig. 10B). Parapodial cirri pattern: Dorsal cirri slightly shorter than notopodial ligules throughout body, basally inserted in anterior-most chaetigers, displaced medially throughout body, passing the ligules throughout body. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body, progressively distal throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 10H), dorsal cirri basal, slightly shorter than dorsal ligules. Dorsal ligules subconical, blunt tip, longer than wide, as long as postchaetal lobes. Neuroacicular ligules digitiform, half as long as postchaetal lobes; postchaetal lobes digitiform; neuropodial ligules digitiform, slightly shorter than postchaetal lobes. Ventral cirri basal, half as long as neuropodial ligules. In anterior chaetigers (Fig. 10I), dorsal cirri 3/4 as long as notopodial ligules. Notopodial ligules subconical, as long as wide, twice longer than prechaetal lobes; notoacicular lobes subconical, 3/4 as long as prechaetal lobes; prechaetal lobes subconical. Neuroacicular lobes digitiform, 3/4 as long as postchaetal lobes; postchaetal lobes digitiform; neuropodial ligules subconical, as long as neuroacicular lobes. Ventral cirri basal, half as long as neuropodial ligules. In middle chaetigers (Fig. 10J), dorsal cirri half as long as notopodial ligules. Notopodial ligules subconical, as long as wide, 2.5 times longer than prechaetal lobes; notoacicular lobes subconical, twice longer than prechaetal lobes; prechaetal lobes digitiform. Neuroacicular lobes digitiform, half as long as postchaetal lobes; postchaetal lobes subconical, half as long as notoacicular lobes; neuropodial ligules subconical, as long as neuroacicular lobes. Ventral cirri basal, half as long as neuropodial ligules. In posterior chaetigers (Fig. 10K), dorsal cirri 3/4 as long as notopodial ligules. Notopodial ligules subconical, longer than wide, 3.5 times longer than prechaetal lobes; notoacicular lobes subconical, twice longer than prechaetal ones; prechaetal lobes digitiform. Neuroacicular lobes digitiform; postchaetal lobes digitiform, 1.2 times longer than neuroacicular ones; neuropodial

94 ligules digitiform, as long as postchaetal lobes. Ventral cirri basal, half as long as neuropodial ligules. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers and sesquigomph falcigers in both supra- and sub-acicular fascicles. Notopodial homogomph spinigers serrate, teeth decreasing in size toward distal end (Fig. 10G). Neuropodial homogomph spinigers as notopodial ones. Neuropodial sesquigomph falcigers pectinate, inner margin completely dentate, distal tooth stout, with a fine tendon fused to margin; blades of falcigers slightly decreasing toward lower-most position of neuropodial fascicles (Figs 10D-F). Praratype with pygidium crenulated; anal cirri missing (Fig. 10C).

Variation. The few specimens vary little. The notopodial ligules in complete paratype tend to be shorter than dorsal cirri toward posterior end.

Remarks. This species resembles N. arenaceodentata, but they differ in some features. In N. 1 sp. n., dorsal cirri are subequal or slightly longer than notopodial ligules, while in N. arenaceodentata they are shorter than notopodial ligules throughout body. Further, neuropodial falcigers of N. sp. 1 have blades almost half-shorter than in N. arenaceodentata; and in N. sp. 1, falcigers have a distal tooth recurved and fused to blade, not observed in N. arenaceodentata. The description by Liñero-Arana & Díaz-Díaz (2007) allows identify their specimens as belonging to this species. The description of N. acuminata provided by Arteaga-Flórez & Londoño-Mesa (2015) from Isla Providencia, Colombia, match with the current description in paragnaths number and chaetal features, but parapodial ones require further revision to ensure their belonging to N. 1 sp. n. The specimen from Bahamas, collected at 810 m depth, is an anterior portion of 18 chaetigers, in bad conditions, with the everted pharynx. It resembles to N. sp.1 in the belt of paragnaths on areas VII-VIII, but it is a distinct species because they differ in the follow features. In specimens from Bahamas, area I has paragnaths arranged in a triangle, while in N. sp. 1 they are in a round. Moreover, in specimen from Bahamas, the notopodial prechaetal lobes are as long as notoacicular ones in anterior chaetigers, whereas in N. sp. 1 they are shorter. However, the poor conditions of the only specimen is inadequate for describe it as a new species.

95

Distribution. Mexican Caribbean, Jamaica.

Neanthes arenaceodentata (Moore, 1903) Figure 11A–L

Nereis arenaceodentata Moore 1903: 720–723, Pl. 15, Figs 1–10. Nereis (Neanthes) arenaceodonta (sic) Pettibone 1963: 162–165, Figs 44i, 45e (partim) Nereis (Neanthes) acuminata Gardiner 1975: 149, Figs 15e, f (non Ehlers, 1868) Neanthes caudata Renaud 1956: 16–17, Fig. 11 (non delle Chiaje, 1828) Neanthes acuminata Salazar-Vallejo & Jiménez-Cueto 1997: 365, Figs 6, 22, 23, 45 (non Ehlers, 1868)

Type locality. Woods Hole, Massachusetts, United States.

Material examined. Northwestern Atlantic Ocean, United States. Topotypes USNM 28093 (13), Woods Hole, Massachusetts, 7 August 1954, Coll. M. Pettibone. USNM 28100 (35), Nonamesset Island, Lackey’s Bay, Massachusetts, 22 August 1950, Coll. M. Pettibone. Caribbean Sea, Mexico. ECOSUR P1216 (2), Laguna Nichupté, Quintana Roo, 5 July 1988, M.S. Jiménez-Cueto. ECOSUR P1215 (1), Laguna Nichupté, Quintana Roo, 27 October 1987, Coll. M.S. Jiménez-Cueto.

Description. Topotypes complete, all with similar length, in good conditions. One specimen selected for description. Specimen complete (Fig. 11C), 13 mm long, 0.8 mm wide, 52 chaetigers. Body pale, pigmentation consist in greenish spots in prostomium and tentacular ring; on tentacular ring and anterior chaetigers (Fig. 11A), spots form a transverse band in each chaetiger, spots more concentrated on lateral margins of chaetiger, becoming scarce toward middle chaetigers. Ventrally, pigmentation consists in several spots at base of parapodia and on middle of each chaetiger (Fig. 11D), darker in midbody, discoloring toward anterior and posterior ends. Prostomium longer than wide, anterior margin entire; antennae cirriform, short, 1/4 as long as prostomium; eyes black, anterior pair ovate, posterior one rounded, anterior pair slightly longer

96 than posterior one, both pairs slightly overlapped (Fig. 11A). Tentacular ring 1.5 times longer than first chaetiger; four pairs of tentacular cirri, longest one reaches chaetiger 3. Pharynx dissected. Maxillary ring: I= 23 cones in a round, dispersed, II= 32–34 cones in arc, III= 68 cones in rectangle, IV=38–39 cones in arc. Oral ring: V–VII= a broad, continuous band of cones, ventral side twice broader than dorsal one, with variable arrangement and paragnaths number; areas VI generally with a paragnath larger than surrounding ones (Fig. 11E). Parapodial cirri pattern: Dorsal cirri shorter than notopodial ligules throughout body, basally inserted in anterior-most chaetigers, displaced medially throughout body, not passing the notopodial ligules throughout body. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 11I), dorsal cirri basal, 3/4 as long as dorsal ligules. Notopodial ligules subconical, blunt tip, subequal to neuropodial ones. Neuroacicular lobes rounded; postchaetal lobes digitiform, 3 times longer than neuroacicular ones, 1.2 times longer than neuropodial ligules; neuropodial ligules digitiform. Ventral cirri basal, 3/4 as long as neuropodial ligules, as long as dorsal cirri. In anterior and middle chaetigers (Fig. 11J, K), dorsal cirri medial, 1/3 as long as notopodial ligules. Notopodial ligules subconical, 3 times broader than notoacicular lobes; notoacicular lobes subconical, subequal to notopodial ligules, slightly longer than prechaetal lobes, twice longer than neuroacicular lobes; prechaetal lobes digitiform. Neuroacicular lobes rounded to digitiform; postchaetal lobes digitiform, twice longer than neuropodial ones; neuropodial ligules subconical, blunt tip, as long as postchaetal lobes (slightly overlapped in Fig. 11K).Ventral cirri basal, half as long as neuropodial ligules (slightly overlapped in Fig. 11K), as long as dorsal cirri. In posterior chaetigers (Fig. 11L), dorsal cirri medial, half as long as notopodial ligules. Notopodial ligules subconical, blunt tip, as long as notoacicular ones; notoacicular lobes subconical, blunt tip; postchaetal lobes reduced, rounded. Neuroacicular lobes digitiform; postchaetal lobes digitiform, twice longer than neuroacicular ones, subequal to neuropodial ligules; neuropodial ligules digitiform. Ventral cirri basal, half as long as neuropodial ligules, half as long as dorsal cirri. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers and sesquigomph falcigers in both supra- and sub-acicular fascicles.

97

Notopodial homogomph spinigers pectinate, coarse teeth, decreasing in size toward distal end (Fig. 11H). Neuropodial homogomph spinigers as notopodial ones; neuropodial sesquigomph falcigers pectinate, coarse teeth, with hooked distal tooth (Figs 11F, G); supra-acicular falcigers slightly longer than sub-acicular ones. Pygidium slightly crenulated; anal cirri cirriform, as long as last 7 chaetigers (Fig. 11B).

Variation. Specimens’ measurements and parapodial proportions almost invariant. Further paragnaths counts not performed due to restrictions to dissect collection specimens, otherwise there are difficulties to observe and count them. At least two patterns of paragnaths were found (Fig. 11E), one with more paragnaths than the other one.

Remarks. This species is regarded as a junior synonym of N. acuminata, which has been regarded as a species complex; however, herein it is regarded as a distinct, valid species. Some morphological differences can be noticed by comparison with the original description of N. acuminata. In N. arenaceodentata, the dorsal cirri are much shorter than notopodial ligules throughout body and not pass the ligules, whereas in N. acuminata dorsal cirri are as long as, or slightly longer than, notopodial ligules throughout body and extending beyond them. Furthermore, in N. arenaceodentata the notopodial structures have similar length throughout body, while in N. acuminata notopodial ligules are longer than remaining notopodial processes throughout body. The specimens from Mexico are indistinguishable from topotypic material, but it is not surprising since others species from temperate latitudes are also present in the Gulf of Mexico and Florida.

Questionable generic affinity

Neanthes egregicirrata (Treadwell, 1924)

Leptonereis egregicirrata Treadwell 1924: 13–14, Fig. 24 (partim) Nereis egregicirrata Treadwell 1939: 233, Fig. 56. Pettibone 1956: 284–287, Figs 4a, 5a–d. Neanthes egregicirrata de León-González et al. 1999: 673–670, Figs 1a–h, 2a–f.

98

Type locality. English Harbor, Antigua.

Material examined. Caribbean Sea, Antigua and Barbuda. Lectotype of Leptonereis egregicirrata USNM 71733, English Harbor, Antigua Island, July 1918, Coll. C.C. Nutting. Paratypes USNM 20324 (4), data as in lectotype.

Remarks. Treadwell placed this species in Leptonereis because he did not note the paragnaths at first, but saw them later (Treadwell 1939), placing it in Nereis. Pettibone (1956) redescribed and revised the type material for this species, recognizing a second species, Nereis allenae. de León- González et al. (1999) redescribed newly the species with type and additional materials from Caribbean Sea, and they confirmed the absence of notopodial falcigers in atokous specimen from Little Cayman, transferring it to Neanthes. The modified chaetiger 6 cannot be regarded as merely epitokous modification since atokes also have the same feature. Moreover, epitokes have modified ligules and lobes in pre-natatory region, instead of retaining the atokous shapes and proportions. At least the chaetigers 6 are unique and does not match with the current definition of Neanthes, even if in other features resembles to that genus. The species and their afiinities will be assessed elsewhere, with the revision of additional material.

Distribution. Caribbean Sea.

Questionable records

Neanthes acuminata (Ehlers, 1868)

Nereis acuminata Ehlers 1868: 552-554, Pl. 22, Figs 23–28.

Type locality. Naples, Mediterranean Sea.

Remarks. This species has a cosmopolitan status. The main reason to this attribution is the morphological similarity assumed among specimens from distant localities, especially since the

99 continuous belt of paragnaths on oral ring have been regarded as a specific feature (e.g. Hoagland 1919). Also, this and other several species have been regarded as junior synonyms of N. caudata delle Chiaje, 1828. However, there are some differences in the parapodial morphology. Recent genetic studies indicated that there are differences among geographically isolated populations at a chromosome and gene levels, even they found some divergences in morphology, reproductive isolation and behavior (Pesch et al. 1988; Reish et al. 2014; Weinberg et al. 1990). Therefore, a detailed revision of these species is required to clarify the synonymies.

Neanthes caudata (delle Chiaje, 1822)

Spio caudatus delle Chiaje 1822: T. 2, Pl. 28, Figs 10, 15.

Type locality. Naples, Mediterranean Sea.

Remarks. This species has a confused, problematic taxonomic history; these problems must be solved to clarify the taxonomic status of synonymized species. The record from the Grand Caribbean region is by Renaud (1956) from Virginia Key, Florida, is not surprising since several species have been regarded as synonymies of N. caudata. The author included sufficient attributes to refer the record to N. arenaceodentata as shown above. de León González (2009) noted that N. caudata cannot be differentiated morphologically from N. acuminata; however, they have genetic differences (Weinberg et al. 1990).

Neanthes irrorata (Malmgren, 1867)

Praxithea irrorata Malmgren 1867: 51, Pl. 4, Figs 24, 24A–D.

Type locality. Bohüslan (Dyngö and Koster Islands), Skagerrak coast of Sweden.

Remarks. The name was introduced in the Caribbean region by Suárez & Fraga (1978); the authors argued that probably the species was present in Cuban waters. However, they did not provide any explanation for this neither the specific document they based to conclude this.

100

Probably some authors were influenced by previous publication like those made by Fauvel, Hartman and Rullier as the authors cited in the synonymy list.

Nereis Linnaeus, 1758

Type species. Nereis pelagica Linnaeus, 1758, by subsequent designation (Hartman 1948).

Diagnosis (modified from Bakken & Wilson 2005). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary ring with conical and sometimes pyramidal or smooth-bar paragnaths in area IV, areas I-III sometimes without paragnaths. Oral ring with conical, pyramidal or p-bars paragnaths, sometimes arising from plake-like basements in areas V–VIII, sometimes missing. First two chaetigers type C. Notopodial ligules with different development types, present throughout body. Prechaetal notopodial lobe present or absent; neuropodial postchaetal lobe present at least on anterior and middle chaetigers. Notochaetae homogomph spinigers and falcigers; neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph falcigers and spinigers in sub-acicular fascicles.

Remarks. For the Grand Caribbean region, most species are in taxonomic confusion, mainly because regional names have been regarded as junior synonyms of older, mainly Scandinavian or Mediterranean species, even with revision of type material. For this effort, revision of types or topotypes for all species was not possible, but each name contains comments for future revisions. The section have some subdivisions: the first contains species generally well known or with distinctive features that allow us to recognize them, species recently redescribed or whose material was available for this work; a subdivision was required to include species synonymyzed with N. riisei. The second contains species with distinctive features but with generic affinity questionable; third section has species requiring a detailed revision; and the last one contains species with doubtful records.

101

Note about nereidids described by Grube in Annulata örstediana. Grube (1858) proposed 11 new Nereis species and placed them into three subgenera: Nereis s.s. Örsted, including N. vallata, Gr. Kr., N. marginata Gr. Örsd, N. krebsii Gr. Örsd, N. rigida Gr. Örsd, N. riisei Gr. Örsd., N. puncturata Gr. Örsd and N. debilis Gr. Örsd; Nereilepas de Blainville, including N. variegata Gr. Kr. and N. callaona Gr. Kr.; and Heteronereis Örsted, including N. caudipunctata Gr. Örsd and N. pannosa Gr. Kr. Also, the latter scheme was followed in the index of all species (Grube 1859: 119). Therefore, these species were placed originally into Nereis, contrary to opinion of Salazar- Vallejo & Eibye-Jacobsen (2012); the confusion seems to have originated by following Hartman’s catalogue of the polychaetes of the world (Hartman 1959b).

Nereis alacranensis Ramírez-Hernández, Hernández-Alcántara & Solís-Weiss, 2015

Nereis alacranensis Ramírez-Hernández et al. 2015: 157–163, Figs 1A–G, 2A–I, 3A, B.

Type locality. Arrecife Alacranes, Yucatán, México, Gulf of Mexico.

Remarks. The presence of a ventral, hard plate in the tentacular ring is a remarkable feature for this species. One parapodium from chaetiger 1 was illustrated (Ramírez-Hernández et al. 2015, Fig. 1D) with two aciculae, probably a mutation. The species belong to Nereis species without paragnaths in areas I and V, few paragnaths in one row in areas VII-VIII and short-bladed falcigers; N. alacranensis resembles to N. grayi Pettibone, but differences between these and other species were discussed by the cited authors.

Nereis arroyensis Treadwell, 1901 Figure 12A–K

Nereis arroyensis Treadwell 1901: 193–194, Figs 30–31; 1939: 223. Hartman 1956: 280.

Type locality. Arroyo, Puerto Rico.

102

Type material. Caribbean Sea, Puerto Rico. Syntypes USNM 16005 (2), San Juan, Puerto Rico, 567 m depth, 13 January 1899, Coll. R/V Fish Hawk.

Description. Syntypes in good condition, one small, incomplete. Largest complete, 43 mm long, 2 mm wide, 70 chaetigers. Body pale, without pigmentation, translucent toward posterior end (Fig. 12B). Prostomium as long as wide, anterior margin entire; antennae broad, one broken, remaining subconical, blunt tip, not passing the palps; eyes slightly reddish, almost faded, in rectangular arrangement (Fig. 12A). Tentacular ring 1.5 times longer than first chaetiger; four pairs of tentacular cirri, posterodorsal tentacular cirri missing in larger specimen, in smaller syntype reach chaetiger 4 (Fig. 12A). Pharynx everted, corrugated, partially dehydrated. Jaws with inner margin dentate, more than 8 teeth (not extracted). Maxillary ring: I= 3 cones in a vertical row, II= 18-17 cones in arc, III= 42 in large rectangle, IV=32-38 cones in arc. Oral ring: V= 0, VI= 6-7 in a row, VII-VIII= 98 small cones in two irregular bands, anterior one shorter than posterior one (Figs 12C–E). Parapodial cirri pattern: Dorsal cirri slightly shorter than, or as long as, notopodial ligules in anterior-most segments throughout body, basally inserted in anterior-most segments, displaced medially throughout body. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 12H), dorsal cirri basal, slightly shorter than notopodial ligules. Notopodial ligules subconical, blunt tip, as long as neuropodial ones, three times longer than neuroacicular lobes. Neuroacicular lobes rounded; postchaetal lobes rounded, half as long as neuroacicular ones; neuropodial ligules subconical, blunt tip (tip broken in Fig. 12H). Ventral cirri basal, 3/4 as long as neuropodial ligules. In anterior chaetigers (Fig. 12I), dorsal cirri medial, slightly shorter than notopodial ligules. Notopodial ligules subconical, blunt tip, twice longer than notoacicular lobes; notoacicular lobes subconical, blunt tip three times longer than neuroacicular ones. Neuroacicular lobes subconical to digitiform, twice longer than postchaetal ones; postchaetal lobes rounded; neuropodial ligules digitiform, as long as neuroacicular lobes (overlapping to neuroacicular lobe in Fig. 12I). Ventral cirri basal, as long as neuropodial ligules.

103

In middle and posterior chaetigers (Figs 12J, K), dorsal cirri medial, as long as notopodial ligules. Notopodial ligules digitiform to subconical, being more acute toward posterior chaetigers, as long as notoacicular ones; notoacicular lobes digitiform to subconical, being more acute toward posterior chaetigers, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, slightly longer than postchaetal ones; postchaetal lobes rounded; neuropodial ligules digitiform, twice longer than neuroacicular lobes. Ventral cirri basal, half as long as neuropodial ligules. Notochaetae homogomph spinigers and falcigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular ones. Notopodial homogomph falcigers slightly falcate, six coarse teeth, one third of inner margin edentulate, distal tooth stout (Fig. 12G). Neuropodial sub-acicular falcigers pectinate, teeth increasing in size toward distal end, long distal tooth with almost inconspicuous recurved end (Fig. 12F). Pygidium swollen; anal cirri cirriform, long, as long as last seven segments (Fig. 12B).

Remarks. Two publications with different date of publication were found, meaning that there is more than one printing of the report. In order to avoid future confusions, a brief note about publication years was made. The earlier version (not found online, but in photocopies) was published in 1901, declaring it was extracted from U.S. Fish Commission Bulletin for 1900, being 1901 the date established by most authors. However, there is an online version (link on references) of the document stating that the first and second part of the activities for 1900 and the report for 1901 of the Bulletin of the United States Fish Commission were published in 1902. The species is easily distinguishable from other Nereis species from Grand Caribean in having a very distinctive paragnaths arrangement on areas VII-VIII, and by the shape of notopodial homogomph falcigers.

Distribution. Puerto Rico.

104

Nereis garwoodi González-Escalante & Salazar-Vallejo, 2003

Nereis garwoodi González-Escalante & Salazar-Vallejo 2003: 156–160, Figs 1a–k, 2a–h. Conde- Vela & Salazar-Vallejo 2015: 28–36, Figs 3, 4, 6B, E, M, I, R.

Type locality. Chetumal Bay, Mexico (18°29’40’’N 88°17’23’’W), 1 m depth.

Remarks. Recently redescribed by Conde-Vela & Salazar-Vallejo (2015). The species is closely related to N. oligohalina but it can be recognized by parapodial and cheatal features, and by different pigmentation.

Nereis grayi Pettibone, 1956

Nereis grayi Pettibone 1956: 282–284, Figs 3a–g.

Type locality. Hadley Harbor, Uncatena Island, Massachusetts.

Remarks. The report by Taylor (1984) from the Gulf of Mexico is doubtful because his material show some differences in parapodial features. However, Neanthes arenaceodentata from Massachusetts is confirmed from the Mexican Caribbean coasts, therefore the presence of N. grayi for this region cannot be discarded completely until revision of types and additional materials. On the other hand, Ramírez-Hernández et al. (2015) detailed diagnostic features from literature that allows recognizing this species from N. alacranensis, and some other related ones.

Nereis lanai Santos, 2007

Nereis lanai Santos 2007: 370–371, Figs 1a–l. Nereis falcaria Santos & Lana 2003: 13–14, Figs 41–46 (non Willey, 1905)

Type locality. Rocas Atoll, Rio Grande do Norte, Brazil.

105

Remarks. This species resembles N. panamensis by their notopodial homogomph falcigers with strongly bifid blade. However, they can be easily recognized by parapodial and chaetal features. For further comments, see remarks section of N. falsa below.

Nereis largoensis Treadwell, 1931

Nereis brevicirrata Treadwell 1929: 3–5, Figs 9–14. Nereis largoensis Treadwell 1931: 3 (replacement name) Nereis pelagica largoensis Hartman 1956: 280. Nereis gracilicirrata Warren 1942: 39–40.

Type locality. Key Largo, Florida Keys.

Remarks. Treadwell (1931) proposed the replacement name N. largoensis, because he had described other species with the same name for Southern Brazilian coasts. The replacement name might be reversed since N. brevicirrata from Brazil rather belongs in Perinereis (Hartman 1956); however, because N. brevicirrata was referred as a junior synonym it is permanently invalid (ICZN 1999, Art. 57.2 ) and does not meet the conditions for reestablishment (ICZN 1999, Arts. 57.2.1–3), then the name largoensis is retained. Hartman (1956) considered this species just a variant in Nereis pelagica Linnaeus, 1758, although she recognized differences in notoacicular lobes and paragnaths count; here it is regarded as a distinct species. The species has very distinctive features such as the slight change in lobes and ligules along body, dorsal cirri increasing their length toward posterior end, and the biconvex (lenticulated) blade of notopodial homogomph falcigers. Warren (1942) used the name gracilicirrata for a material collected in Grand Isle, Louisiana; however, the description is very general and it can just ensure the presence of his specimens to the family. Hartman (1954) concluded that N. gracilicirrata is a synonym of N. largoensis, and the synonymy is retained until a further revision of this species.

106

Nereis occidentalis Hartman, 1945

Nereis pelagica occidentalis Hartman 1945: 20, Pl. 4, Figs 1–6.

Type locality. Beaufort, North Carolina.

Remarks. N. oligohalina (Rioja, 1946) was considered as a junior synonym of N. occidentalis, thus several reports of this species could be attributed to the former, currently considered a Pan- American species. Conde-Vela & Salazar-Vallejo (2015) recently redescribed N. oligohalina and separated it from N. occidentalis, and their presence in the Grand Caribbean must be reassessed.

Nereis oligohalina (Rioja, 1946)

Neanthes oligohalina Rioja 1946: 207–210, Pl. 1, Figs 3–6, Pl. 2, Figs 13–19. Nereis oligohalina Hartman 1954: 414. Conde-Vela & Salazar-Vallejo 2015: 19–28, Figs 1A–O, 2 A–N, 6A, D, J–L, O, Q.

Type locality. Chachalacas sandbar, Veracruz, Gulf of Mexico.

Remarks. The species has been redescribed recently by Conde-Vela & Salazar-Vallejo (2015), and separating it from their related N. garwoodi, N. occidentalis and N. confusa, the latter from Pacific Mexican coasts.

Nereis panamensis Fauchald, 1977 Figure 13A–M

Nereis panamensis Fauchald 1977a: 29–31, Figs 6d–i.

Type locality. Galeta Island Reef, Panama, Caribbean Sea.

107

Type material. Caribbean Sea, Panama. Holotype USNM 53139, Galeta Island Reef, Panama, 5 October 1970, on coralline zone, Coll. STRI (Smithsonian Tropical Research Institute). Paratypes USNM 53140 (9), Galeta Island Reef, Colon, on Laurencia zone, Coll. STRI. Paratypes LACM-AHF 1136 (3), Galeta Island Reef, Colon (9°24'18"N 79°51'48.5"W), on Laurencia zone, 12 February 1971, Coll. STRI.

Additional material. Caribbean Sea, Panama. Topotypes USNM 73687 (2), Galeta Reef Flat, Panama, 18 May 1973, Coll. H.O. Brattstrom.

Description. Type material in regular condition. Holotype incomplete, posterior end slightly dried, 13 mm long, 1.4 mm wide, 46 chaetigers; tentacular cirri, one antenna and most dorsal cirri missing (Fig. 13A), pharynx dissected previously. Paratypes in regular condition, one dried, specimens 6–8 mm long, one with oocytes (Fig. 13B); tentacular cirri missing in most specimens. All specimens with body pale, pigmentation absent. Topotypes (USNM 73687) two specimens, largest one complete and used for description, 13 mm long, 1.1 mm wide, 53 chaetigers. Prostomium as long as wide, margin entire, with a slight medial groove; antennae lanceolate, passing the palps, as long as prostomium; eyes black, rounded, subequal, in trapezoidal arrangement (Figs 13A, C). Tentacular ring twice longer than first chaetiger; four pairs of tentacular cirri, longest one reaches chaetiger 4 (Figs 13A, C). Pharynx of holotype previously dissected. Jaws brown, 12 teeth, cutting edge completely dentate. Maxillary ring: I= 1 small cone, II= 7-8 in two rows, in arc, III= 11 cones in an ellipse, IV=17-19 cones in arc. Oral ring: V= 0, VI=8-10 cones in a round, VII-VIII= 14 cones in a single row, big cones alternating with small ones. Parapodial cirri pattern: Dorsal cirri longer than notopodial ligules throughout body, passing notopodial ligules throughout body, basally inserted in most- anterior chaetigers, displaced medially throughout body. Ventral cirri longer than neuropodial ligules in anterior segments, becoming subequal to neuropodial ligules toward posterior end, basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 13F), dorsal cirri basal, 1.5 times longer than notopodial ligules. Notopodial ligules digitiform, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, half as long as postchaetal ones; postchaetal lobes digitiform, as long as wide; neuropodial ligules twice longer and wider than

108 postchaetal lobes. Ventral cirri basal, 1.5 times longer than neuropodial ligules, 3/4 times as long as dorsal cirri. In anterior chaetigers (Fig. 13G), dorsal cirri medial, 1.6 times longer than notopodial ligules. Notopodial ligules wide at the base, narrow at the tip, subequal to notoacicular lobes; notoacicular lobes digitiform, three times longer than neuroacicular ones. Neuroacicular lobes subconical, blunt tip; postchaetal lobes subconical, blunt tip, twice longer than neuroacicular ones; neuropodial ligules digitiform, as long as postchaetal lobes. Ventral cirri basal, slightly longer than neuropodial ligules, half as long as dorsal cirri. In middle and posterior chaetigers (Figs 13H, I), dorsal cirri medial, 1.5 times longer than notopodial ligules. Notopodial ligules subconical, blunt tip, subequal to notoacicular lobes; notoacicular lobes digitiform, three times longer than neuroacicular ones. Neuroacicular lobes rounded; postchaetal lobes subconical, blunt tip, twice longer than neuroacicular ones; neuropodial ligules digitiform, as long as postchaetal lobes. Ventral cirri basal, subequal to neuropodial ligules, half as long as dorsal cirri. Notochaetae homogomph spinigers in anterior chaetigers, homogomph falcigers since chaetiger 19. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles. Notopodial homogomph spinigers pectinate, fine teeth, decreasing their length distally. Neuropodial homogomph spinigers as notopodial ones (Fig. 13J); heterogomph spinigers pectinate, fine teeth, decreasing their length distally. Homogomph falciger with bifid appearance, a distal tooth very stout, rounded tip and one medial tooth, acute tip; 2–3 small teeth, sometimes covered by the shaft (Fig. 13K). Heterogomph falciger pectinate, fine teeth, supra-acicular (Fig. 13M) slightly stouter than sub-acicular ones (Fig. 13L); blades with falcate tips, 1/2 of length of margin edentulate (Figs 13L, M). Pygidium tripartite, as long as last two segments (Fig. 13D); one anal cirri remaining, cirriform, as long as last five segments (Fig. 13E).

Variation. Posterodorsal tentacular cirri reach chaetigers 3–5 in paratypes (USNM 53140). Topotype specimen with paragnaths as follow. Maxillary ring: I= 1 cone, II= 12-13 cones, small and big, in arc, III= 10 cones in ellipse, VI= 10-11 in a round. Oral ring: V= 0, VI=8-9 cones in round, VII-VIII= 10 in a single row, big cones alternating with small ones.

109

Remarks. The posterior end of the holotype is lost. The original description has some important discrepancies with the type material examined from USNM, but is agree with the paratypes deposited in LACM-AHF. Some important differences are found among type specimens. Holotype has paragnaths on areas I and VI and are clearly arranged in round, while in paratypes from AHF they are arranged in a row. Fauchald (1977a) described the notopodia without superior lobes (i.e. notopodial ligules), although they are present in all type series. However, the type material from USNM has notopodial ligules throughout body, whereas in specimens from AHF, they are restricted to anterior parapodia, being shorter than notoacicular lobe and reduced to a very small lobe toward posterior end Further, he described the neuropodia as a bifid structure, with the superior division with pre- and postchaetal lobes (i.e. superior and inferior lobes), but this feature was not observed in specimens examined. On the other hand, de León-González et al. (1999), following the original description, identified material from the Caribbean Sea and Gulf of Mexico as N. panamensis; however, description provided by the cited authors differ with the current one of the holotype provided here, even differ in some features from the paratypes of LACM-AHF. The absence or evident reduction of notopodial ligules toward posterior end cannot be regarded as merely populational variation, since it is a feature to distinguish species (Santos 2007). A new species could be described, after the revision of additional material to assess if these features are not populational variations. species deserve a more detailed analysis to determine if type materials from AHF are, in fact, a distinct species.

Nereis riisei Grube & Ørsted in Grube, 1858

Nereis riisei Grube 1858: 162–163.

Type locality. St. Croix, Virgin Islands, Caribbean Sea.

Remarks. Salazar-Vallejo & Eibye-Jacobsen (2012) indicated Puntarenas, Costa Rica (Pacific Ocean) as the type locality of this species; however, St. Croix is the type locality as indicated

110 both in the description (Grube 1858: 163) and the index for the series (Grube 1859: 119). The label in the syntype material deposited in the Zoological Museum, University of Copenhagen (ZMUC–POL–1487) pointed the type locality is St. Croix and collected by Kröyer (Eibye- Jacobsen, pers. comm.). The species has a largely confused taxonomic history. Probably the confusion of type localities has caused the species to be regarded as having an amphiamerican status. The first amphiamerican report was made by Monro (1933a), synonymizing several species with N. riisei, later Hartman followed his conclusion in subsequent works (Hartman 1938a, 1940, 1944, 1954). This decision has been widely followed, except for some species as N. glandulata Hoagland, 1919 (e.g. Treadwell 1924, 1928). The species deserves a complete revision with materials from several localities and type materials of each species involved. Here some comments of the current synonymized species are added. The material examined and described here resembles more than one species referred to N. riisei, and they were referred as N. cf. riisei 1 and 2.

Species names referred to Nereis riisei

Nereis fasciata (Schmarda, 1861)

Heteronereis fasciata Schmarda 1861: 100–101, Pl. 31, Fig. 241, text Figs a, b, c, d, e, f, K. Augener 1925: 25–26. Hartman 1959b: 242.

Type locality. Jamaica.

Remarks. As the generic placement suggests, this species was described with epitokes. Augener (1925) revised the type material, and based upon the paragnaths on areas VII-VIII and the chaetae, determined that is identical to N. riisei. Diagnostic features include jaws with five teeth, areas VII-VIII with few paragnaths in one row, neuropodial heterogomph falcigers with short and long blades, natatory region starts at chaetigers 22–23. Several specimens of N. fasciata (v900 and v901), supposed to be the type material observed by Augener, are deposited at the Natural

111

History Museum of Vienna (H. Sattmann, pers. comm.); a further revision could confirm their belonging to N. riisei, once it is redefined.

Nereis bicruciata Augener, 1906

Nereis bicruciata Augener 1906: 151–153, Pl. 5, Figs 102–104, Pl. 6, Fig. 105.

Type locality. St. Croix, Virgin Islands, Caribbean Sea, at 210 m depth.

Remarks. Augener described this species based upon a one incomplete specimen from St. Croix. Hartman (1938a) examined the type specimen and determined that this species, together with N. nigripes Ehlers, are synonyms of N. riisei, and she identified some differences in dentition of jaws and number of paragnaths. Diagnostic features include tentacular cirri reaching chaetigers 10-11, dorsal cirri longer than notopodial ligules in anterior chaetigers, areas VII-VIII with six paragnaths in a row. Also, N. bicruciata was collected at 210 m depth, whereas N. riisei seems to be a shallow species.

Nereis glandulata Hoagland, 1919

Nereis glandulata Hoagland 1919: 575, Pl. 30, Figs 1–6.

Type locality. Guanica Harbor, Puerto Rico.

Remarks. Synonymyzed with N. riisei by Hartman (1944). The species could be effectively the same for its proximity with the type locality. Diagnostic features include eyes large and anterior and posterior pairs overlapping, tentacular cirri reach chaetiger 2, jaws with four teeth, area I with two paragnaths in a vertical row, areas VII-VIII with five paragnaths in one row, dorsal cirri longer than notopodial ligules in anterior chaetigers, falciger with long blade. Hoagland (1919, Pl. 30, fig. 5) reported heterogomph spinigers in notopodial and neuropodial, supra-acicular fascicles; their presence should be corroborated, but it seems to be a confusion.

112

Nereis varia Treadwell, 1941

Nereis (Neanthes) paucidentata Treadwell 1939: 6, Fig. 25. Nereis (Neanthes) varia Treadwell 1941: 3 (replacement name) Nereis riisei Hartman 1956: 279-280 (partim, non Grube & Ørsted in Grube, 1858)

Type locality. Charlestown, Massachussets.

Remarks. Type material of this species was examined by Hartman (1956). Diagnostic features include tentacular cirri reach chaetiger 4, dorsal cirri shorter than notopodial ligules throughout body, area I with paragnaths, areas VII-VIII with two rows of paragnaths. The areas VII-VII with two rows of paragnaths, if present, would be enough to separate it from N. riisei species group. Treadwell (1939) noted the dark pigmentation pattern on the dorsum and parapodia of each chaetiger, becoming darker toward posterior end. The shape of notopodial homogomph falcigers was not noticed.

Nereis nigripes Ehlers, 1868

Nereis nigripes Ehlers 1868: 508–510.

Type locality. Florida and Haiti.

Remarks. Augener (1925) regarded N. nigripes sensu Ehlers and Fauvel as identical to N. riisei. Monro (1933a) synonymized N. nigripes with N. riisei without further explanation. Hartman (1938a), based upon the decision by the latter author, syonymized this species and N. bicruciata with N. riisei. The database of MCZ indicates that the specimen from Haiti (ANNa-54) is located in that collection.

Nereis lata Hansen, 1882

Nereis lata Hansen 1882: 11. Augener 1934: 130.

113

Type locality. Rio de Janeiro, Brazil.

Remarks. Augener (1934) concluded that this species must be regarded as synonym of N. riisei. Diagnostic features include ligules and lobes with dark glands in middle and posterior chaetigers, areas VII-VIII with 4 paragnaths.

Nereis cf. riisei 1 Figure 14A–K

Material examined. Gulf of Mexico, United States. ECOSUR P0000 (1), R/V Gerda, Cruise 6717, St. 887 (25°05'N 86°28'W), E Gulf of Mexico, 97 m depth, 9 September 1967.Caribbean Sea, Mexico. ECOSUR P0000 (1), R/V Gerda, Cruise 6717, St. 898 (21°04’N 86°19W), E Arrowsmith Bank, 352 m depth, 10 September 1967. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7028, St. 1286 (21°06'N 86°28'W), Arrowsmith Bank, 280 m depth, 23 August 1970. Jamaica. ECOSUR P0000 (2), R/V Pillsbury, Cruise 7006, St. 1223 (17°47'N 77°41'W), S Jamaica, 28 m depth, 6 July 1970. Dominican Republic. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7006, St. 1271 (17°41'N 71°41'W), S Cabo Falso, 18 m depth, 18 July 1970. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7006, St.1272 (17°52'N 71°41'W), Bahia de las Aguilas, 24 m depth, 18 July 1970. Honduras. ECOSUR P0000 (1), R/V Pillsbury, Cruise, Sta. 1362 (15°36’N 88°04’W), Omoa Bay, Honduras, 24 m depth, 1 February 1971. Panama. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6607, St. 321 (09°21'N 79°57'W), Manzanillo Bay, Panama, 0 m depth, 5 July 1966. Saint Kitts and Nevis. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6907, St. 960 (17°27'N 62°54'W), NW St. Christopher, St. Kitts and Nevis, 32 m depth, 19 July 1969. Antigua and Barbuda. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6907, St. 967 (17°16'N 62°02'W), NW Antigua, 22 m depth, 20 July 1969. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6806, St. 968 (17°18'N 61°52'W), N Antigua, Antigua And Barbuda, Lesser Antilles, 18 m depth, 20 July 1969. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6802, St. 975 (17°29'N 61°55'W), SW Barbuda, Lesser Antilles, 29 m depth, 21 July 1969. Colombia. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6607, St. 349 (08°32'N 77°02'W), N Gulf of Urabá, Colombia, 54 m depth, 11 July 1966. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6607, St. 372 (09°46'N 76°10'W), W San Bernardo Archipelago, Colombia, 91 m depth, 13 July 1966. Western Atlantic Ocean, French

114

Guiana. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6806, St. 650 (06°07'N 52°19'W), NE French Guiana, 87 m depth, 8 July 1968.

Description. Specimen complete, 65 mm long, 2.5 mm wide, 123 chaetigers. Body pale, without pigmentation, massive glands in notopodial and notoacicular lobes and neuropodial ligules throughout body (Figs 14B–F). Prostomium longer than wide, anterior margin entire; antennae cirriform, as long as prostomium; eyes reddish, subequal, anterior pair slightly reniform, posterior one rounded (Fig. 14A). Tentacular ring twice longer than first chaetiger; four pairs of tentacular cirri, longest reach chaetiger 7 (Fig. 14A). Pharynx dissected. Maxillar ring: I= 4 cones, II= 7-11 cones, III=9 cones, IV= 10-12 cones. Oral ring: V= 0, VI= 3-3 cones, VII-VIII= 3 cones in one row. Parapodial cirri pattern: dorsal cirri as long as notopodial ligules throughout body, becoming longer in posterior chaetigers, basally inserted in anterior chaetigers, displaced medially toward posterior end. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 14C), dorsal cirri as long notopodial ligules, slightly displaced medially. Notopodial ligules digitiform, as long as neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, as long as postchaetal lobes; postchaetal lobes subconical, blunt tip; neuropodial ligules subconical, twice longer than neuroacicular lobes. Ventral cirri basal, 3/4 as long as neuropodial ligules. In anterior chaetigers (Fig. 14D), dorsal cirri medial, as long as notopodial ligules. Notopodial ligules subconical, blunt tip, as long as notoacicular lobes; notoacicular lobes rounded, as long as neuroacicular ones. Neuroacicular lobes subconical to digitiform, as long as postchaetal ones; postchaetal lobes digitiform; neuropodial ligules digitiform, as long as neuroacicular ones. Ventral cirri basal, 3/4 as long as neuropodial ligules. In medial and posterior chaetigers (Figs 14E–G), dorsal cirri medial, as long as notopodial ligules, becoming 1.5 times longer than notopodial ones in posterior chaetigers. Notopodial ligules subconical, as long as notoacicular ones; notoacicular lobes subconical to digitiform, 1–2 times longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, as long as neuropodial ligules, becoming slightly shorter in posterior chaetigers; postchaetal lobes as long as

115 neuroacicular ones, inconspicuous in posterior chaetigers; neuropodial ligules digitiform. Ventral cirri basal, 3/4 as long as neuropodial ligules. Notochaetae homogomph spinigers and falcigers. Neurochaetae homogomph spiniger and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular fascicles. Notopodial homogomph spinigers as neuropodial ones; homogomph falcigers pectinate, fine teeth, 1/3 of its inner edge edentate, distal tooth stout and falcate (Fig. 14H). Neuropodial homogomph spinigers serrate, teeth decreasing in size toward distal end (Fig. I). Neuropodial heterogomph falcigers pectinate, fine teeth, distal tooth stout, with a distal tendon fused to margin; supra-acicular falcigers stouter and with less teeth than sub-acicular ones (Figs 14J, K). Pygidium tripartite; anal cirri incomplete, as long as last 5 chaetigers (Fig. 14B).

Remarks. N. sp. 1 resembles to N. glandulata in the shape of prostomium, the size and position of eyes and in the paragnaths number, but they differ in the parapodial topology, the size of ventral cirri and in the shape of neuropodial heterogomph falcigers in anterior chaetigers. On the other hand, N. sp. 1 rather resembles N. bicruciata. The size and position of eyes, reaching of tentacular cirri, the development of parapodia and the neuropodial heterogomph falcigers as figured by Augener are very similar to those found in N. sp. 1. However, they differ in the reaching of the tentacular cirri and in the number of paragnaths; Augener reported areas VII-VIII as bare, however in several species examined, paragnaths on these areas are very small and difficult to observe, requiring methyl-green staining or an oblique illumination to improve the contrast and observe them. Unfortunately, notopodial homogomph falcigers were not illustrated to ensure the identification.

Nereis cf. riisei 2 Figure 15A–K

Material examined. Northwestern Atlantic Ocean, Turks and Caicos. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7106, St. 1421 (21°36’N 71°01’W), NE Grand Turk, 83 m depth, 19July 1971. Caribbean Sea, Jamaica. ECOSUR P0000 (3), R/V Pillsbury, Cruise 7006, St. 1195 (17°24'N 76°02'W), SE Jamaica, 14 m depth, 3 July 1970. ECOSUR P0000 (3), R/V Pillsbury,

116

Cruise 7006, St. 1196 (17°28'N 75°57'W), NE Morant Cays, Jamaica, 26 m depth, 3 July 1970. ECOSUR P0000 (2), R/V Pillsbury, Cruise 7006, St. 1252 (17°09'N 78°57'W), W Pedro Bank, Jamaica, 26 m depth, 14 July 1970. Dominican Republic. ECOSUR P0000 (2), R/V Pillsbury, Cruise 7006, St. 1254 (?N ?W), (Cruise from Pedro Bank, Jamaica and Dominican Republic), 20 m depth, 14 July 1970. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7006, St. 1284 (17°35'N 71°25'W), Cape Beata, Dominican Republic, 20 m depth, 19 July 1970. Nicaragua. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7101, St. 1343 (13°58'N 82°04'W), E Nicaragua, 25 m depth, 30 January 1971. Antigua and Barbuda. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 956 (17°00'N 62°30'W), NW Redonda Island, 37 m depth, 19 July 1969. Grenada. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6907, St. 852 (11°53'N 61°53'W), SW Grenada, Lesser Antilles, 13 m depth, 3 July 1969. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6907, St. 853 (11°55'N 61°43'W), S Grenada, Lesser Antilles, 17 m depth, 3 July 1969.

Description. Specimen complete (ECOSUR Pilsb1196), in good condition, 30 mm long, 2 mm wide, 62 chaetigers. One specimen (ECOSUR Pilbs 1254) with dark brown pigmentation throughout body; pigmentation present on prostomium and in palpophores, tentacular ring completely covered (Fig. 15A). Dorsal pigmentation as one wide transverse band per segment along anterior body, breaking into several rectangular patches toward posterior chaetigers; glands present in notopodial and neuropodial ligules and in notoacicular lobes throughout body (Figs 15A, B, H–K). Prostomium as long as wide, anterior margin entire; antennae cirriform, as long as prostomium; eyes rounded, subequal, in trapezoidal arrangement (Figs 15A, C). Tentacular ring as long as first chaetiger; four pairs of tentacular cirri, longest reach chaetiger 8 (Figs 15A, C). Pharynx dissected; jaws brown, 10 teeth, cutting edge completely dentate. Maxillary ring: I= 1 cone, II=18-20 cones in arc, III= 15 in rectangle, IV= 14-16 cones in arc. Oral ring: V=0, VI=8- 10 cones in round, VII-VIII= 5 cones in one row. Parapodial cirri pattern: dorsal cirri longer than notopodial ligules in anterior and middle chaetigers, becoming as long as notopodial ligules in posterior ones, basally inserted in anterior- most chaetigers, displacing medially toward posterior end. Ventral cirri as long as neuropodial ligules and basally inserted throughout body.

117

First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 15H), dorsal cirri basal, 1.2 times longer than notopodial ligules. Notopodial and neuropodial ligules subconical, blunt tip, subequal, several times wider than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, half as long as postchaetal lobe; postchaetal lobes subconical, blunt tip, 1/3 as long as neuropodial ligules. Ventral cirri basal, as long as neuropodial ligules. In anterior chaetigers (Fig. 15I), dorsal cirri medial, 1.5 times longer than notopodial ligules. Notopodial ligules subconical, blunt tip, as long as notoacicular ones; notoacicular lobes rounded, 1.2 times longer than neuroacicular ones. Neuroacicular lobes rounded, as long a neuropodial ligules; postchaetal lobes as long as neuroacicular ones; neuropodial ligules digitiform to rounded. Ventral cirri as long as neuropodial ligules. In middle and posterior chaetigers (Figs 15J, K), dorsal cirri medial as long as notopodial ligules. Notopodial ligules subconical, as long as and twice wider than notoacicular lobes; notoacicular lobes subconical, blunt tip, 2–3 times longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip; postchaetal lobes as long as neuroacicular ones; neuropodial ligules digitiform, twice longer than neuroacicular lobes. Ventral cirri basal, as long as neuropodial ligules. Notochaetae homogomph spinigers and falcigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular ones. Notopodial homogomph spinigers pectinate, fine teeth, decreasing in size toward distal end; homogomph falcigers pectinate, fine teeth, half of inner edge dentate, distal tooth very stout (Fig. 15E). Neuropodial heterogomph falcigers pectinate, fine teeth; supra-acicular falcigers stouter than sub-acicular ones, blades of supracicular ones half as long as sub-acicular ones (Figs 15F, G). Pygidium tripartite; anal cirri subulate, as long as last 7chaetigers (Fig. 15D).

Remarks. The morphologic features match with N. riisei species group; however, current descriptions do not ensure their belonging to any species, or if it is a new species. As indicated above, N. sp. 1 resembles N. glandulata in the dark pigmentation on prostomium and tentacular ring, the number of paragnaths and the presence of dark glands in notopodial ligules and lobes

118 and in neuropodial ligule. However, they differ in the arrangement of eyes, the reaching of tentacular cirri, the parapodial topology and the shape of neuropodial heterogomph falcigers. On the other hand, N. sp 1 resembles N. fasciata in having neuropodial heterogomph falcigers with distinct length of blade, but parapodial topology cannot be compared. Nereis sp. 2 is easily recognized from N. sp. 1 in having notopodial homogomph falcigers with less teeth restricted to a half of the inner edge, and in having supra-acicular falcigers much stouter and with shorter blades than sub-acicular ones. Also, in N. sp. 2, neuropodial heterogomph falcigers does not have distal tendon or at least they are inconspicuous, feature present in falcigers of N. sp. 1.

Questionable generic affinity

Nereis allenae Pettibone, 1956 Figure 16A–D

Nereis allenae Pettibone 1956: 287–290, Figs 4b–f, 6a–h.

Type locality. Near Caballo Blanco Island, Parguera, Puerto Rico.

Type material. Caribbean Sea, Puerto Rico. Holotype USNM 27778, Parguera, 15 September 1955, Coll. M. J. Allen. Antigua and Barbuda. Paratypes USNM 27779 (4), English Harbor, Antigua Island, July 1918, Coll. N.N. Nutting. Barbados. Paratypes USNM 27780 (3), Pelican Island, June 1918, Coll. N.N. Nutting.

Description. Holotype female incomplete, 5.6 mm long, 0.8 mm wide, 52 chaetigers. Oocytes present, some parapodia previously dissected. Body pale, without pigmentation, some glands observed in upper parapodia. Paratypes (USNM 27779) three males, two complete, remaining incomplete, largest and best preserved used for description (Figs 16A, B). Male complete, body pale, 8.6 mm long, wide, 76 chaetigers, pigmentation absent. Prostomium as long as wide, anterior margin entire; antennae 1/3 long of prostomium, eventrally displaced; eyes rounded, reddish, subequal, anterior-most pair ventrally displaced. All tentacular cirri missing. Tentacular ring 1/2 as long as first chaetiger (Figs. 16A)

119

Dorsal cirri missed almost all in atokous region. Body divided into three regions. Pre-natatory region include chaetigers 1–14, natatory region 15-59, and the caudal region 60–76 (Fig. B). First two chaetigers type C; remaining biramous. First seven chaetigers with modified lobes and ligules: neuroacicular ones hypertrophied, proximally blunt, distally narrow (Fig. 16C). Neurochaetae include a very stout, ankylosed falcigers, pectinate, curved teeth, distal end in an obtuse angle (Fig. D). Two ventral papillae per segment present along natatory and caudal regions (Fig. 16B).

Remarks. Pettibone (1956) included atokes and epitokes specimens as type material; however, it is difficult to know if atoke specimens correspond with the epitokes ones because they were collected in different localities, indeed. Male paratypes (USNM 27779) have ankylosed, modified falcigers in chaetigers 1-7, and its parapodial lobes are modified along these chaetigers, but these features are missing in female holotype and this type of difference has not been reported between adult members of the same species. These two characters are unusual for Nereis and consequently, these males might be transferred to a new genus, whereas the correspondence among males, females and atokes must be assessed elsewhere.

Nereis caecoides Hartman, 1965 Figure 17A–O

Nereis caecoides Hartman 1965b: 85–86, Pl. 11, Figs c, d.

Type locality. Off Bermuda, 2000 m.

Type material. Northwestern Atlantic Ocean, Bermuda. Holotype of Nereis caecoides LACM-AHF 818, R/V Atlantis, Sta. Bermuda 5 (32°11'24"N 64°41'36"W), Bermuda slope, 2000 m depth, 2 August 1960, Coll. H.L. Sanders.

Description. Holotype complete, broken into four fragments, 18 mm long, 1 mm wide, 62 chaetigers (Fig. 17A). Body cylindrical, pale, without pigmentation.

120

Prostomium as long as wide, half as wide as tentacular ring; tentacular ring as long as prostomium, with an anterior constriction at the tentacular cirri region; eyes absent (Fig. 17B). Tentacular ring 1.5 times longer than first chaetiger; four pairs of tentacular cirri, three cirri remaining, longest of them reach chaetiger 1 (Fig. 17B). Pharynx previously dissected; jaws light brown, 12 teeth, rounded tips, cutting edge completely dentate. Arrangement of paragnaths on pharynx inconspicuous. A fragment from areas III or IV with a patch of polygonal, scale-like paragnaths, densely packed, sometimes overlapped (Figs 17E, F). Parapodial cirri pattern: Dorsal cirri shorter than notoacicular lobes throughout body, basally inserted in anterior chaetigers, displaced medially in middle chaetigers, not passing lobes throughout body. Ventral cirri basal and shorter than neuropodial ligules in anterior and middle chaetigers, as long as neuropodial ligules in posterior chaetigers. First two chaetigers type C, remaining biramous. In chaetigers 3-4 (Fig. 17K), dorsal cirri 3/4 as long as notopodial ligules. Neuropodial ligules digitiform, enlarged, longer than wide, appearing from chaetiger 5, as long as notoacicular lobe; Notoacicular lobes digitiform, twice longer than neuroacicular ones. Neuroacicular lobes digitiform, postchaetal lobes absent; neuropodial ligules digitiform, enlarged, as long as notoacicular lobes. Ventral cirri half as long as neuropodial ligules. In chaetigers 5–35 (Fig. 17L), dorsal cirri basal, 2/5 as long as notoacicular lobes. Notopodial ligules digitiform subconical, diminishing their length progressively; notoacicular lobes digitiform, enlarged, 1.2 times longer than neuropodial ones, papillae conspicuous. Neuroacicular lobes digitiform; postchaetal lobes absent; neuropodial ligules digitiform, enlarged, as long as notoacicular lobes. Ventral cirri basal, half as long as neuropodial ligules. From chaetiger 36 to end of body (Figs 17M, N), dorsal cirri medial, half as long as notoacicular lobes. Notoacicular lobes digitiform to subconical, enlarged, 1.5–2.0 times longer than neuroacicular lobes, papillae conspicuous; notopodial ligules absent from chaetiger 36. Neuroacicular lobes subconical, blunt tip, twice longer than neuropodial ligules; neuropodial ligules digitiform, enlarged, medially inserted in neuroacicular lobes. Ventral cirri basal, as long as neuropodial ligules. Notochaetae homogomph falcigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falciger in sub-acicular fascicles.

121

Notopodial homogomph falcigers (Fig. 17G) with edentate blades, half of blade embedded in shaft, not fully emerged in middle chaetigers (Fig. 17O). Neuropodial homogomph spinigers (Fig. 17H) pectinate, fine teeth, decreasing in size toward dirsal end. Neuropodial heterogomph falcigers pectinate, coarse teeth, distal tooth falcate, inner edge completely dentate; supra-acicular falcigers with blades longer than sub-acicular ones (Figs 17I, J). Pygidium tripartite; anal cirri subulate, as long as last five chaetigers (Fig. 17C).

Remarks. Nereis caecoides has remarkable features that even its generic affinities are problematic, although it lacks eyes and prostomial shape is also present among other deep-sea nereidids. In addition, the parapodial structures are narrow, and notoacicular and neuroacicular lobes are very prominent; the notopodial ligules are present in some anterior chaetigers only, as Hartman had indicated. Notopodial homogomph spinigers were not seen, even in a whole-mount specimen, and just one homogomph falciger per fascicle was observed. Hartman probably observed the embed falcigers in middle chaetigers and regarded them as spinigers with broken blades. Hartman did not detail the pharyngeal armature, and it seems she included the species in Nereis based upon the presence of notopodial homogomph falcigers only. In fact, there are no paragnaths if observed with a naked eye, and mounting pharynx fragments for examination under microscope is difficult. Therefore, a small portion of pharyngeal tissue (areas III or IV) was extracted, mounted, and then examined in a compound microscope. A patch of small, flat paragnaths was present; as indicated above, these paragnaths are polygonal, and in some cases, they have a projected arista. Flat paragnaths were observed in Micronereis halei, but they are different to those present in N. caecoides. In M. halei, they are round and rather isolated, whereas in N. caecoides they are polygonal and are contiguous. Here the term scale-like is proposed for this type of paragnaths. Fauchald (1977b) recognized that species attributed to Nereis from deep waters have a distinct parapodial morphology. Ligules and lobes generally are long and slender, perhaps a deep-sea adaptation (Fauchald 1972), and parapodial rami are very separate and divergent. In general, these parapodia resemble those present in a recently reinstated deep-sea genus, Typhlonereis Hansen, 1879 (Bakken 2003), and in another, currently regarded as a junior synonym Profundilycastis Hartmann-Schröder, 1965. However, N. caecoides and N. caymanensis differ

122 from both genera in having both paragnaths and homogomph falcigers. N. caecoides seems to be more similar to Profundilycastis profundus (Hartman, 1965), described in the same publication; further comments in the corresponding remarks for that genus.

Nereis caymanensis Fauchald, 1977 Figure 18A–G

Nereis caymanensis Fauchald 1977b: 160–162, Pl. II, Figs a–d.

Type locality. West Cayman Trench, Caribbean Sea, 6840–6850 m depth.

Type material. Paratype LACM-AHF 1206, Akademik Kurchatov, Sta. 1259, West Cayman Trench, 5800-6500 m depth, 23 March 1975.

Description. Paratype complete, 8 mm long, 0.8 mm wide, 30 chaetigers. Body pale, without pigmentation (Fig. 18A). Prostomium as long as wide; antennae cirriform, as long as prostomium; eyes minute, anterior pair visible only (Fig. 18B). Tentacular ring 1.5 times longer than first chaetiger; four pairs of tentacular cirri, longest reach chaetiger 4 (Fig. 18B). Pharynx previously dissected, damaged; jaws brown, five teeth, distal half of cutting edge dentate. Maxillary ring: I= 1 cone, II= 16-¿? cones, III= ¿?, IV=21-24 cones. Oral ring: V= 0, VI= 0-2 cones, VII-VIII= 8 cones. Parapodial cirri pattern: dorsal cirri longer than notopodial ligules throughout body, basally inserted in anterior-most chaetigers, displaced medially toward posterior chaetigers. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body. First two chaetigers type C, remaining biramous. In anterior chaetigers (Fig. 18C), dorsal cirri medial, 1.5 times longer than notopodial ligules. Notopodial ligules digitiform; notoacicular lobes digitiform, enlarged, twice longer than notopodial ligules and neuroacicular lobes, papillae conspicuous. Neuroacicular lobes subconical, blunt tip, postchaetal lobes inconspicuous; neuropodial ligules digitiform, half as long as neuroacicular lobes. Ventral cirri basal, as long as neuropodial ligules.

123

Notochaetae homogomph spinigers and falcigers. Neurochaetae homogomph spiniger and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular fascicles. Notopodial homogomph spinigers as neuropodial ones; homogomph falcigers pectinate, coarse teeth, inner edge dentate almost completely dentate, distal tooth stout (Fig. 18E). Neuropodial homogomph spinigers pectinate, fine teeth, decreasing in size toward distal end (Fig. 18D); heterogomph falcigers pectinate, fine teeth, 1/6 of inner margin edentate, distal tooth stout (Figs 18F, G).

Remarks. This description matches Fauchald (1977b) on parapodial morphology, but differs in other features. The number of paragnaths on areas VII-VIII, Fauchald reported about 30 paragnaths in several rows, whereas in the paratype only 8 were found. The pharynx was damaged after the previous dissection, and possibly some fragments from areas VI-VIII were detached. On the other hand, his figure (Fauchald 1977b, Pl. 2, Fig. C) illustrating the notopodial homogomph falciger shows a much shorter blade than the one observed in the paratype. However, he illustrated a parapodium from chaetiger 19, and probably the homogomph falciger belong to this chaetiger; in some cases, blades of chaetae reducing their length toward posterior chaetigers, it seems to be the case for N. caymanensis. This species resembles N. caecoides in parapodial features; however, the shape of the anterior end, the blades of notopodial homogomph falcigers, and the presence of conical paragnaths separate them.

Nereis goajirana Augener, 1933

Nereis goajirana Augener 1933a: 253–257, Figs 5a–c.

Type locality. Riohacha, La Guajira, Colombia.

Remarks. N. goajirana has notopodial accessory lobes, presumably postchaetal ones, and therefore does not fit within Nereis definition, but parapodial topology resembles some Neanthes

124 species provided with three notopodial structures. Diagnostic features include tentacular cirri reaching chaetiger 9 or 10; dorsal cirri shorter than notopodial ligules throughout body; paragnaths present in all areas except V. The species seems to be related with L. guadalupensis, see remarks section for this species.

Nereis heterodonta (Schmarda, 1861)

Mastigonereis heterodonta Schmarda 1861: 110–111, Pl. 21, Figs 252, text Figs A,a,B,b,K. Nereis heterodonta Hartman 1959b: 260.

Type locality. Boca Chica, Cartagena, Colombia; and Jamaica, in corals.

Remarks. Schmarda (1861) reported a foliose structure over dorsal surface of the body adjacent to parapodia, becoming closer to dorsal cirri toward posterior segments. The foliose structure has not been reported for any Nereis species neither other genera, therefore the inclusion by Hartman (1959b) into Nereis is doubtful. Other diagnostic features include jaws dentate, long tentacular cirri; dorsal cirri medial, slightly longer than notopodial ligules, lobes and ligules subequal, increasing in length toward posterior end, except neuropodial lobes being shorter than remaining structures. Schmarda's material is deposited in the Natural History Museum of Vienna; however, there is no material left for N. heterodonta (H. Sattmann, pers. comm.); therefore, topotypes are required to redescribe the species and confirm the presence of the foliose structures.

Species requiring further revision

Nereis anodonta Schmarda, 1861

Nereis anodonta Schmarda 1861: 105–106, text Figs A,a,B,b,K.

Type locality. South coasts of Jamaica.

125

Remarks. Schmarda (1861) noted that species has edentate jaws, however the author did not include critical features as pharyngeal armature or presence of notopodial homogomph falcigers to ensure their belonging to the genus. Other diagnostic features include tentacular cirri short and slender; dorsal cirri shorter than notopodial ligules throughout body; ventral cirri shorter than neuropodial ligules in anterior body, becoming subequal toward posterior end; lobes and ligules subequal in anterior chaetigers, notopodial ligules becoming longer than remaining structures toward posterior chaetigers. Schmarda's material is deposited at the Natural History Museum of Vienna; however, there is no material of N. anodonta (H. Sattmann, pers. comm.). Topotypic materials are required to redescribe the species.

Nereis caudipunctata Grube & Ørsted in Grube, 1858

Nereis (Heteronereis) caudipunctata Grube 1858: 166–167.

Type locality. St. Croix, Virgin Islands, Caribbean Sea, at 37 m depth.

Remarks. The description was based upon epitoke specimens. Salazar-Vallejo & Eibye-Jacobsen (2012) located the holotype (ZMUC-POL-1494). Diagnostic features include natatory region begin in chaetigers 14-15, antennae as long as prostomium, longest tentacular cirri reach chaetiger 8; dorsal cirri evidently modified along chatigers 4-7 , features in natatory parapodia very general. The transformation pattern resembles a Platynereis species but its transfer must be postponed to the revision of the type material.

Nereis krebsii Grube & Ørsted in Grube, 1858

Nereis krebsii Grube 1858: 161.

Type locality. St. Croix, Virgin Islands, Caribbean Sea.

126

Remarks. Augener (1858) did not find the type material of this species, but one specimen identified as N. krebsii from the Pacific Ocean; he discouraged the used of the name after the scarce original description. Recently Salazar-Vallejo & Eibye-Jacobsen (2012) located the holotype (ZMUC-POL-1482), and pointed out that Hartman (1959b) regarded it as indeterminable, but this status requires the study of the type materials. Probably the species belongs to another genus. Diagnostic features include tentacular cirri reaching chaetiger 4 and notopodial ligules and lobes are more prominent than neuropodial ones.

Nereis marginata Grube & Ørsted in Grube, 1858

Nereis marginata Grube 1858: 160–161.

Type locality. St. John, Virgin Islands, Caribbean Sea.

Remarks. Salazar-Vallejo & Eibye-Jacobsen (2012) located the holotype (ZMUC-POL-1483), and pointed that Hartman (1959b) regarded it as indeterminable. Diagnostic features include tentacular cirri reach chaetiger 4, notopodial ligules longer than notoacicular lobes, dorsal cirri shorter than notopodial ligules.

Nereis puncturata Grube & Ørsted in Grube, 1858

Nereis puncturata Grube 1858: 163.

Type localities. Valparaiso, Chile, Pacific Ocean; St. Croix, Virgin Islands, Caribbean Sea.

Remarks. The species was described with specimens from two distinct localities. Salazar-Vallejo & Eibye-Jacobsen (2012) located the type material in two museums: specimen from St. Croix is deposited in the Zoological Museum, University of Copenhagen (ZMUC-POL-1486), while specimens from Valparaiso are deposited in the Zoological Museum, Berlin (ZMB 4040). Grube did not include type localities in the description, but it is in the index for all species (Grube

127

1859). The species requires the proposition of lectotypes to separate the Atlantic from the Pacific forms.

Questionable records

Nereis callaona Grube & Kröyer in Grube, 1858

Nereis (Nereilepas) callaona Grube 1858: 165–166.

Type locality. Callao, Peru, Pacific Ocean.

Remarks. The name was introduced by Fauchald (1977) from Galeta Reef (Atlantic) and Paitilla Beach (Pacific), Panama, but he recognized differences in blades of chaetae. The species has an amphiamerican status but it is doubtful and deserves a further study to solve its affinities.

Nereis falcaria (Willey, 1905)

Ceratonereis falcaria Willey 1905: 272, Pl. 4, Fig. 89.

Type locality. Southwest Cheval Paar, Sri Lanka, Indian Ocean.

Remarks. The names N. falcaria (Willey, 1905), N. kauderni Fauvel, 1921 and N. jacksoni Kinberg, 1865 have been reported from the Gran Caribbean and other regions as Nereis species with notopodial homogomph falcigers with strongly bifid ends, often disregarding parapodial morphology. The regional species belonging of this group are N. panamensis and N. lanai, records from the region might belong to either species.

Nereis falsa de Quatrefages, 1865

Nereis falsa de Quatrefages 1865: 505.

128

Type locality. Crimea, Black Sea, Mediterranean Sea.

Remarks. The name was introduced in several reports (Dean 2012; Salazar-Vallejo 1996). This species, currently regarded as cosmopolitan, has a confused, problematic taxonomic history, including the fact that it is not listed for the Black Sea. These problems must be solved to clarify the taxonomic status of synonymized species, and the possible proposal of new species for the Grand Caribbean and others regions.

Nereis jacksoni Kinberg, 1865

Nereis jacksoni Kinberg 1866: 169.

Type locality. Port Jackson (Sydney Harbour), Australia.

Remarks. The name was introduced for Cuban coasts (Salazar-Vallejo 1996). See remarks section for N. falcaria above. Salazar-Vallejo & Jiménez-Cueto (1997) doubtfully identified two heterenereidids as N. jacksoni and provided differences in paragnaths number between specimens from Caribbean Sea and from Australia. After the revision of specimens, the belonging to Nereis group with bidentate notopodial homogomph falcigers is confirmed, but they cannot be identified with certainty. Probably the heteronereis belong to N. panamensis since they have notopodial ligules as long as notoacicular lobes in chaetigers from pre-natatory region.

Nereis kauderni Fauvel, 1921

Nereis kauderni Fauvel 1921: 8–11, Pl. 1, Figs 1–7.

Type locality. Majunga (Mahajanga), Madagascar.

Remarks. The name (current synonym of N. falcaria) was introduced by (Fauvel 1953). See remarks section for N. falcaria above.

129

Nereis lamellosa Ehlers, 1868

Nereis lamellosa Ehlers 1868: 564–567, Pl. 22, Figs 10–17.

Type locality. Lagosta (Lastovo, Croatia), Adriatic Sea.

Remarks. The specimens identified under this name seem to be an Alitta species, and this has often been considered as a synonym of A. succinea due to their similar morphology (Day 1973). However, the specimens reported by Taylor (1984) have notopodial homogomph falcigers, hence cannot be assigned into Alitta. The species deserves a redescription with types and topotypic material to clarify their current taxonomy and to separate it from other related species.

Nereis pelagica Linnaeus, 1758

Nereis pelagica Linnaeus 1758: 654; 1761: 508–509. Chambers & Garwood 1992: 38–39, Figs 1b, 3, 4, 5f, 6a-b, 17a-c, 18a-b, 26c, 31, 45a–l, 57, 68a–d.

Type locality. Mare Hallandiam (Kattegat Sea), Atlantic Ocean.

Remarks. There is no redescription available of this species; meanwhile, the complete description of Chambers & Garwood (1992) from the United Kingdom coasts is widely adopted as the true N. pelagica. Possibly records belong to several species due to the high morphological variability attributed to this species. Because this is the type species for the genus, it deserves an exhaustive clarification of all species involved.

Nereis rigida Grube & Ørsted in Grube, 1858

Nereis rigida Grube 1858: 162.

Type locality. Puntarenas, Costa Rica.

130

Remarks. The name was introduced by Augener (1906) when he tried to redescribe this species based upon Caribbean specimens (Dominica); however, he did not offer any explanation about the conspecificity. The repository of the type material is unknown (Salazar-Vallejo & Eibye- Jacobsen 2012), the species require a redescription with topotypes, although it is recognizable in some identification keys (Conde-Vela & Salazar-Vallejo 2015; González-Escalante & Salazar- Vallejo 2003; Ramírez-Hernández et al. 2015), and it must be restricted to Eastern Pacific coasts.

Nereis trifasciata Grube, 1878

Nereis (Lycoris) trifasciata Grube 1878: 74–75.

Type locality. Bohol, Philippines.

Remarks. Ehlers (1901) introduced the name for the fauna of Juan Fernandez Archipelago, Chile. Hartman (1964) made a distinction from the original N. trifasciata and the one described by Ehlers, and she considered that N. trifasciata sensu Ehlers belongs to Neanthes, whereas the first one belongs to Nereis because it has notopodial homogomph falcigers. However, in her catalogue (1965:38) she considered ‘Ehlers´s species’ in Nereis as well, even as an independent species. These species are different, and Ehlers´ specimens deserve to be proposed as a distinct species. The name was introduced by Monro (1939) to the Grand Caribbean fauna with a supposed juvenile from Gorda Cay, Honduras; he based the identification in Fauvel (1923).

Nereis zonata Malmgren, 1867

Nereis zonata Malmgren 1867: 46–47, Pl. 5, Figs 34, 34A–D.

Type locality. Spitsbergen, Svalbard, Arctic Ocean.

Remarks. The name was introduced by Fauvel (1923a) and followed by Liñero-Arana & Reyes- Vásquez (1979). The species deserves a redescription with types and topotypic material to clarify

131 their current taxonomy and other related species, but its presence in the Grand Caribbean is doubtful.

Nereis victoriana Augener, 1918

Nereis victoriana Augener 1918: 180–184, Pl. 2, Figs 29, 30; Pl. 3, Figs 72, 73; text Fig. 13.

Type locality. Limbe (Victoria), Cameroon.

Remarks. The name was introduced by Augener (1922) in a species list without further references. González-Escalante & Salazar-Vallejo (2003) noted the relationship among N. victoriana, N. callaona and N. falsa, but hesitated about their presence in the Grand Caribbean.

Perinereis Kinberg, 1865

Type species. Nereis novaehollandiae Kinberg, 1865, by subsequent designation (Hartman 1948).

Diagnosis (after Bakken & Wilson 2005). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary ring with conical paragnaths in areas I, II, III; area IV with or without paragnaths. Oral ring with conical and smooth-bar paragnaths in areas VI–VIII, area V sometimes without paragnaths. First two chaetigers type C. Notopodial ligules with different development types, sometimes pennant-like in posterior chaetigers. Prechaetal notopodial lobe present or absent; neuropodial postchaetal lobe present throughout body. Notochaetae homogomph spinigers; neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph falcigers and spinigers in sub-acicular fascicles.

Remarks. The species belonging to this genus are very similar to each other and are difficult to separate, indeed. However, after a detailed examination there are slight and consistent

132 differences. The number of bar paragnaths on area VI has been widely used for separating species groups: with one bar on each side, with two bars, and with three or more bars. Also, the use of paragnaths number on areas I and V has been used as diagnostic features, but rejected in some occasions. Hylleberg et al. (1986) found variation in these areas in their material identified as P. striolata (Grube, 1878) and therefore they dismissed their use. Based on this and other ‘minor’ variants as colour patterns, length of tentacular cirri, and size of ligules, they synonymized five species; also, the authors argued that P. striolata is a “very plastic species” (Hylleberg et al. 1986), although they used Area I for separating it from P. cultrifera (Grube, 1840). However, other studies corroborate their utility for separating similar species. For example, Fauvel (1932) made an identification key for the genus, practically based on paragnaths number and type; also, he included additional keys to identify the different ‘varieties’ of P. nuntia (Savigny in Lamarck, 1818) and P. cultrifera based on areas I and V. Statistical analyses of paragnaths number in Perinereis have been performed. Hutchings et al. (1991) made a PCA analysis of paragnaths number for Australian Perinereis species, and found areas V and VII-VIII useful for separating three species, and areas I, II and IV for other two; also the identification key was based mostly on paragnaths number and type. Wilson & Glasby (1993) used a CVA analysis for discriminating some samples referred to P. nuntia species group, and found some cases of non-normality in counts on areas I and V, but they were not discarded because they did not affect their conclusions. Moreover, Glasby & Hsieh (2006) found a very low correlation between body size and paragnaths number in areas I, V and VII-VIII, regarding both as size-independent together with dorsal cirrus-notopodial ligule ratios. Recently, Coutihno et al. (2015) demonstrated that paragnaths number, especially in areas I and V, are useful for separating two sympatric species: P. anderssoni (Kinberg, 1865) and P. ponteni (Kinberg, 1865). Additionally to paragnaths number, there are other important parapodial features, mainly the relative size of notopodial ligule, the dorsal cirri and the ratio of length of the last two. The material examined in this study showed slight but consistent differences and resulted in the recognition of some new species and another one is reinstated, but they are not considered as cryptic species or regional variants.

133

Perinereis bairdii (Webster, 1884) reinst. Figures 19A–F, 20A–N

Nereis bairdii Webster 1884: 312–313, Pl. 8, Figs 22a–28. Nereis (Perinereis) melanocephala McIntosh 1885: 216–219, Pl. 34, Figs 14–17, Pl. 16a, Figs 8– 9. Perinereis bairdii Monro 1933: 41 (partim). Perinereis anderssoni Salazar-Vallejo & Jiménez-Cueto 1997 (non Kinberg, 1865). Perinereis floridana Salazar-Vallejo & Jiménez-Cueto 1997: 367, Figs 8, 32, 33 (partim, non Ehlers, 1868).

Type locality. Bermuda.

Type material. Northwestern Atlantic Ocean, Bermuda. Syntypes of Nereis bairdii USNM 4786 (~7), Bermuda, 1876, Coll. G.B. Goode.

Additional material. Caribbean Sea, Mexico. LACM-AHF 7367 (3), Isla Contoy, Quintana Roo (21°27'36.85"N, 86°47'10.40"W), on coralline turf, mostly Amphiroa and Gelidium, 0.6 m depth, 2 March 2001, Coll. L. Harris. LACM-AHF 7368 (1), Isla Contoy, Quintana Roo (21°27'36.85"N, 86°47'10.40"W), on Halimeda, 1 m depth, 2 March 2001, Coll. L. Harris. ECOSUR P1112 (1), Isla Contoy, Quintana Roo, Mexico (21°30'10.75"N 86°48'0.95"W), 11 October 1993, Coll. S.I. Salazar-Vallejo. ECOSUR P0000 (1), Corona Beach, Cozumel Island, Quintana Roo, Mexico (20°22'18.42"N 87° 1'27.24"W), 25 February 1999, in sabelids’ reef, 2 m depth, Coll. S.I. Salazar-Vallejo, L.F. Carrera-Parra. ECOSUR P1109 (2), Yamach, Quintana Roo, Mexico (19°55'58.07'' N 87°26'40.64''W), 28 February 1986, Coll. L. Aguilar, J. Fernández. ECOSUR P1110 (2), Hualalpich, Ascensión Bay, Quintana Roo, Mexico (19°53'17.12"N 87°25'48.00"W), 19 June 1986, Coll. L. Aguilar, A. Gómez. ECOSUR P1115 (1), Playa Aventuras, Quintana Roo, Mexico (20°21'47.2''N 87°19'53.1''W), 18 May 1986. ECOSUR P0000 (2), Mahahual, Quintana Roo, Mexico (18°43'27.06"N 87°42'1.53"W), 0-1 m depth, in corals, 5 June 1998. ECOSUR P0000 (2), Mahahual, Quintana Roo, Mexico (18°43'27.06"N 87°42'1.53"W), in dead corals, 18 May 2002. ECOSUR P0000 (1), Mahahual, Quintana Roo,

134

Mexico (18°43'27.06"N 87°42'1.53"W), 1 December 2000. ECOSUR P0000 (1), Mahahual, Quintana Roo, Mexico (18°43'29.08"N 87°42'0.99"W), in rocky substrate, 18 March 2001. Jamaica. ECOSUR P0000 (1), R/V Pillsbury, Cruise 7006, St. 1195 (17°24'N 76°02'W), SE Jamaica, 14 m depth, 3 July 1970. Panama. ECOSUR P0000 (1), Club Náutico, Colón, 3 June 2002, Coll. S.I. Salazar-Vallejo. Venezuela. ECOSUR P0000 (1), Turpialito, Cumaná, on Millepora, 15 m depth, 22 February 2002, Coll. I. Liñero-Arana. ECOSUR P0000 (1), Turpialito, Cumaná, on rocky substrate, 2 m depth, 25 February 2002, Coll. S.I. Salazar-Vallejo. ECOSUR P0000 (3), Mochima, Cumaná, on Millepora, 1 March 2002, Coll. M. Liñero. British Virgin Islands. LACM-AHF 7369 (17), Long Point, Guana Island (18°29’0”N 64°34’58”W), on large barnacles, 13 July 2000, Coll. T. Zimmerman, G. Hendler, J. Martin, R. Ware.

Description. Type material includes ten syntypes, one dried, all complete, some belong to P. floridana, one was dissected for description. Body pale, without pigmentation (Figs 19A, B), 26 mm long, 1.7 mm wide, 64 chaetigers. From non-type material, one selected for dissection and comparing with dissected syntype, and another with pharynx everted to indicate paragnaths arrangement and jaws. Specimen dissected complete, 24 mm long, 1.8 mm wide, 64 chaetigers. Body pale, pigmentation brown, slightly faint; four thin, parallel lines arising from bases of antennae, extending towards eyes, and two inner, oblique lines fused into one line extending from back. Tentacular ring with four rectangular spots and one central. First chaetiger variegated, second and remaining ones with pigmentation as in tentacular ring, spots becoming sharper and thinner toward posterior body end (Figs 20A, B). Prostomium longer than wide, anterior margin entire; antennae subulate, half as long as prostomium, not passing palps; eyes rounded, subequal, in rectangular arrangement, (Figs 19A, 20A). Tentacular ring twice longer than first chaetiger; four pairs of tentacular cirri, longest tentacular cirri reach chaetiger 2 up to 4 (Figs 19A, 20A). Pharynx partially everted in syntype specimen (oral ring visible only), fully everted in non-type one; jaws brown with 6 teeth, cutting edge completely dentate (Fig. 20J). Maxillary ring: I= 7 cones in round, II= 20-23 cones in arc, III= 31 cones in ellipse, IV= 38-45 cones in round. Oral ring: V= 0, VI= 1-1 smooth bars, VII-VIII= 23 (32 in syntype) in two rows, anterior one alternating, forming a jagged line, posterior one in transverse line (Figs 19A, B; 20H, I).

135

Parapodial cirri pattern: Dorsal cirri shorter than notopodial ligules throughout body, basally inserted in anterior region, displaced medially in middle chaetigers, subdistal in posterior chaetigers, passing the ligules throughout body. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Figs 19C, 20C), dorsal cirri basal, slightly shorter than notopodial ligules. Notopodial and neuropodial ligules globose, subequal, three times longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, slightly longer than postchaetal ones; postchaetal lobes rounded. Ventral cirri basal, slightly shorter than neuropodial ligules. In anterior chaetigers (Figs 19D, 20D), dorsal cirri medial, slightly shorter than notopodial ligules. Notopodial ligules subconical, blunt tip, wider than long, twice longer than notoacicular lobes; notoacicular lobes digitiform, twice longer than neuroacicular ones, as long as neuropodial ligules. Neuroacicular lobes subconical, blunt tip, distally bilobed, slightly longer than postchaetal ones; postchaetal lobes rounded; neuropodial ligules digitiform. Ventral cirri basal, slightly shorter than neuropodial ligules. In middle and posterior chaetigers (Figs 19E; 20E, F), dorsal cirri medial, 3/4 as long as notopodial ligules. Notopodial ligules subconical, blunt tip, twice longer than notoacicular lobes; notoacicular lobes subconical, blunt tip, three times longer than neuroacicular lobes, slightly longer than neuropodial ligules. Neuroacicular lobes rounded, distally bilobed, slightly longer than postchaetal lobes; postchaetal lobes rounded; neuropodial ligules digitiform, subequal to neuroacicular lobes. Ventral cirri basal, slightly shorter than neuropodial ligules. In posterior-most chaetigers (Figs 19F, 20G), dorsal cirri subdistal, half as long as notopodial ligules. Notopodial ligules pennant-like, four times longer than notoacicular lobes; notoacicular lobes digitiform, twice longer than neuroacicular ones. Neuroacicular lobes rounded, slightly longer than postchaetal ones; postchaetal lobes rounded; neuropodial ligules digitiform, subequal to or slightly longer than neuropodial ligules. Ventral cirri basal, slightly longer than neuropodial ligules. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers and heterogomph falciger in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular ones. Notopodial homogomph spinigers pectinate, fine teeth, decreasing in size toward distal end (Fig. 20M). Neuropodial homogomph spinigers as notopodial ones; heterogomph spinigers pectinate,

136 fine teeth, decreasing in size toward distal end (Fig. 20N). Neuropodial heterogomph falcigers falcate, pectinate, two-thirds of inner margin edentulate, distal tooth stout, with a fine tendon fused to inner margin (Figs 20K, L). Pygidium crenulated; anal cirri subulate, as long as last four segments (Fig. 20B).

Variation. The pigmentation often is faded-off, it is likely darker in living specimens. The parapodial proportions vary little; in posterior chaetigers, the notopodial ligules shape can vary, in some occasions the tip of dorsal cirri extend beyond the tip of notopodial ligules and sometimes do not extend beyond, however proportions remain constant.

Remarks. The syntypes of Nereis bairdii have two distinct morphological patterns: one form corresponds to larger specimens having long posterodorsal tentacular cirri, reaching chaetiger 8, while in the other form specimens are shorter and have short tentacular cirri, reaching chaetiger 2. Webster illustrated the head (pl. 8, fig. 22) of a specimen with long tentacular cirri and its pharynx withdrawn, and other with pharynx everted to show pharynx armature and another specimen for parapodia. After revision, pharyngeal armature and parapodia of specimens with long tentacular cirri match with description of P. floridana, however specimens were not separate, and must be separated from the type series. Treadwell (1901, 1939) recognized two varieties of this species based upon pigmentation and development of notopodial ligules in posterior parapodia. The first variety has darker brown pigment in anterior region, with white spots on brown bands of prostomium and on segments (matching P. bairdii), and the notopodial ligules in posterior chaetigers are well developed (longer and wider than remaining structures); the second variety has brown pigment but less extended, and lacks white spots on prostomium and segments, notopodial ligules less developed and with tentacular cirri longer than in first variety (might be P. floridana or other species). Hartman (1944) synonymized N. bairdii and Nereis (Perinereis) melanocephala McIntosh, 1885 with P. anderssoni Kinberg, 1865. However, there are some differences that can be used to separate them. In P. bairdii, the dorsal and neuropodial ligules in first two chaetigers are globose, while in P. anderssoni they are subconical. Also, in P. bairdii the dorsal cirri are shorter than notopodial ligules throughout body, whereas in P. anderssoni they are longer than notopodial ligules in anterior chaetigers and becoming shorter toward posterior ones. Furthermore, in P.

137 anderssoni the neuropodial heterogomph falcigers are stouter and with inner margin less dentate than in P. bairdii. Specimens identified as P. anderssoni and a part of the material identified as P. floridana by Salazar-Vallejo & Jiménez-Cueto (1997) belong to P. anderssoni; other records from the Caribbean Sea might belong to P. bairdii.

Perinereis cariacoensis Liñero-Arana, 1983

Perinereis cariacoessis (sic) Liñero-Arana 1983: 5–6, Figs 4a–e.

Type locality. Ensenada Honda, Gulf of Cariaco, Venezuela.

Remarks. Liñero-Arana (1983) proposed this species from Venezuelan coasts based on one specimen; therefore, intraspecific variation was not evaluated. The author noted differences between P. cariacoensis and P. mochimaensis, however de León-González & Goethel (2013) considered that the features included are not sufficient to ensure their validity, requiring a revision. Further, P. cariacoensis was described based upon one complete specimen, whereas P. mochimaensis with incomplete one. The latter has a notopodial ligule very developed in posterior segments, even if not are the posterior-most, while in P. cariacoensis this feature was not observed. The paragnaths number among them is very close, however the areas V and VI cause confusion. The genus have a high incidence of sympatric species, so here P. cariacoensis is considered valid, but concurring with de León-González & Goethel (2013), both species require a revision with more specimens to discard possible malformations.

Perinereis cariboea de León-González & Solís-Weiss, 1998 Figure 18H–N

Perinereis cariboea de León-González & Solís-Weiss 1998: 677–680, Figs 3A–E.

Type locality. Punta Pájaros, Ascensión Bay, Mexico.

138

Material examined. Caribbean Sea, Mexico. ECOSUR P0000 (1), Cozumel Island (20°22'18.02"N 87° 1'27.16"W), Quintana Roo, Mexico, 25 February 1999.

Description. Specimen complete, 9 mm long, 1.1 mm wide, 34 chaetigers. Body brown, slightly faint. Palps with a spot on inner, basal margins; four thin, parallel lines arising from bases of antennae, extending towards eyes, and two oblique lines arising from the base of the latter, posteriorly forming a triangular spot; Tentacular ring pigmented medially. Anterior chaetigers with rectangular pigment spot on middle area of chaetigers; first chaetiger variegated; second and subsequent segments with two rounded spots, on posterolateral segment margins, four rectangular spots in rectangle, and another one central per segment, becomes T-shaped, discoloring posteriorly (Fig. 18H). Prostomium longer than wide; antennae cirriform, four times longer than wide, passing palps; eyes rounded, subequal, in rectangle (Fig. 18H). Tentacular ring twice longer than first chaetiger; four pairs of tentacular cirri, longest one reaches chaetiger 4 (Fig. 18H). Pharynx dissected, damaged; jaws with 10 teeth, cutting edge completely dentate (Fig. 18N). Maxillary ring: I= 2 cones in line, II= 8-6 cones in arc, III and IV damaged. Oral ring: V= 0, VI= 2-2 smooth bars, VII-VIII= 11 cones in two rows, anterior row alternating, forming a jagged line, posterior series transverse, linear. Parapodial cirri pattern: Dorsal cirri slightly shorter than notopodial ligules in anterior region, becoming shorter toward posterior end, passing notopodial ligules throughout body, basally inserted in anterior chaetigers, displaced medially in middle body, and subdistal in posterior chaetigers. Ventral cirri shorter than neuropodial ligules throughout body, basally inserted in anterior region, progressively distal throughout body. First two chaetigers type C, remaining biramous. In anterior chaetigers (Fig. 18K), dorsal cirri medial, slightly shorter than neuropodial ligules. Notopodial ligules subconical, blunt tip, wider than long, slightly longer than notoacicular lobes; notoacicular lobes subconical, blunt tip, as long as wide, slightly longer than neuroacicular ones. Neuroacicular lobes subconical, blunt tip, wider than long; postchaetal lobes rounded, longer than neuroacicular ones; neuropodial ligules digitiform, subequal to neuroacicular lobes. Ventral cirri basal, shorter than neuropodial ligule, slightly shorter than dorsal ones.

139

In middle chaetigers (Fig. 18L), dorsal cirri medial, shorter than notopodial ligules. Notopodial ligules subconical, blunt tip, as long as wide, subequal to notoacicular lobes; notoacicular lobes subconical, blunt tip, as long as wide, slightly longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, wider than long; postchaetal lobes rounded, longer than neuroacicular ones; neuropodial ligules digitiform, subequal to neuroacicular lobes. Ventral cirri shorter than neuropodial ligules, slightly shorter than dorsal ones. In posterior chaetigers (Fig. 18M), dorsal cirri subdistal, half as long as notoacicular lobes. Notopodial ligules wider than long, subconical, dorsal surface convex, ventral one flat, 2–3 times longer than notoacicular lobes; notoacicular lobes subconical, blunt tip, slightly longer than wide, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, wider than long; postchaetal lobes rounded, subequal to neuroacicular ones; neuropodial ligules digitiform, shorter than neuroacicular lobes. Ventral cirri basal, shorter than neuroacicular lobes, slightly shorter than dorsal ones. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles. Notopodial homogomph spinigers pectinate, teeth decreasing in size towards end. Neuropodial homogomph spinigers as notopodial ones; heterogomph spinigers pectinate, teeth decreasing in size towards end. Heterogomph falcigers pectinate, distal tooth stout (Fig. 18J), supra-acicular falcigers broader than sub-acicular ones. Pygidium crenulated; anal cirri subulate, as long as last five segments (Fig. 18I).

Remarks. The species resembles P. paradisea sp. n. (see below), especially in the pigmentation pattern, but there are some differences that separate them. In P. cariboea, there are fewer paragnaths on maxillary ring than in P. paradisea sp. n., mainly in area I; the arrangement of paragnaths on area VII-VIII also differs, in P. cariboea there are two rows, while P. paradisea has a single row. Moreover, in P. cariboea notopodial ligules become up to two times longer than notoacicular lobes towards posterior end, and dorsal cirri are subulate, whereas in P. paradisea sp. n. notopodial ligules become up to four times longer than notoacicular lobes, and dorsal cirri are subclavate. Additionally, heterogomph falcigers in P. cariboea have more teeth than those found in P. paradisea sp. n.

140

Perinereis floridana (Ehlers, 1868) Figure 19G–J, 21A–L

Nereis floridana Ehlers 1868: 503–506. Perinereis floridana Monro 1933b: 256. Nereis bairdii Webster 1884: 312, Pl. 8, Fig. 22 (partim)

Type locality. Captiva Island (Key), Florida.

Type material. Northwestern Atlantic Ocean, Bermuda. Syntypes of Nereis bairdii USNM 4786 (+2), Bermuda, 1876, Coll. G.B. Goode. Additional material. Gulf of Mexico, United States. LACM-AHF 7377 (1), Lemon Bay, Charlotte County, Florida (26°54’50”N 82°20’26”W), 11 January 1938, Coll. O. Hartman. LACM-AHF 7376 (3), Lemon Bay Flats, Charlotte County, Florida (26°54’50”N 82°20’26”W), 14 January 1938, Coll. O. Hartman. Caribbean Sea, Mexico. ECOSUR P1100 (16), Isla Cerritos, Yucatan, 18 January 1991, Coll. Sergio I. Salazar-Vallejo. ECOSUR P1112 (1), Isla Contoy, Quintana Roo, Mexico, 11 October 1993, Coll. Sergio I. Salazar-Vallejo. ECOSUR P0000 (2), Holbox, Quintana Roo, Mexico, in rocks, 4 May 2000, coll. C. Campos-Vázquez , C. Llanes-Baeza. ECOSUR P0000 (1), Mahahual, Quintana Roo, Mexico, 6 April 2012, Col. S.I. Salazar-Vallejo, V.M. Conde-Vela.

Description. Specimens from Isla Cerritos complete, pigmentation faint. Best complete, preserved specimen, 52 mm long, 2.8 mm wide, 107 chaetigers, tapered posteriorly. Reminiscence of pale lines present, fingerprint-like pattern present from chaetiger 8. Syntypes of P. bairdii complete, in good conditions, one with pharynx everted; body pale, without pigmentation (Figs 19G, H). Prostomium longer than wide, medial groove present; antennae cirriform, half as long as prostomium, not passing the palps; eyes rounded, subequal, in rectangle, lateral pairs very distant each other, on lateral margins of prostomium (Fig. 21A). Tentacular ring 1.5 times longer than first chaetiger; tentacular cirri with cirrophore conspicuous, longest one reaches chaetiger 7 (Fig. A); antennae and tentacular cirri slightly moniliform.

141

Pharynx dissected, everted in syntypes of P. bairdii (Figs 19H–J); jaws brown with 8 teeth, rounded tips, first two-thirds of cutting edge dentate (Fig. H). Maxillary ring: I= 2 cones in vertical line, II= 7-9 cones in arc, III= 13 cones in rectangle, IV= 16-17 cones in arc (Fig. 19I). Oral ring: V= 1 cone, VI= 1-1 smooth bars, VII-VIII= 18 in two irregular rows, anterior row in line, posterior row alternating, forming a jagged line (Fig. 19J) Parapodial cirri pattern: Dorsal cirri slightly shorter than notopodial ligules in anterior region, becoming shorter toward posterior end, not passing notopodial ligules throughout body, basally inserted in anterior chaetigers, displaced medially throughout body. Ventral cirri shorter than neuropodial ligules throughout body, basally inserted in anterior region, progressively distal throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 21C), dorsal cirri basal, slightly shorter than notopodial ligules. Notopodial ligules digitiform, as long as neuropodial ligules; neuroacicular lobes subconical, half as long as neuropodial ligules; postchaetal lobes absent; neuropodial ligules digitiform. Ventral cirri basal, shorter than neuropodial ligules. In anterior chaetigers (Fig. 21D), dorsal cirri basal, slightly shorter than notopodial ligules. Notopodial ligules subconical, blunt tip, as long as wide, slightly longer than notoacicular lobes; notoacicular lobes subconical, blunt tip, slightly longer than neuroacicular lobes, papillae conspicuous. Neuroacicular lobes subconical, blunt tip, slightly shorter than neuropodial ligules, distally bilobed; neuropodial ligules digitiform, subequal to neuroacicular lobes. Ventral cirri basal, shorter than neuropodial ligules. In middle and posterior chaetigers (Figs 21E, F), dorsal cirri medial, half as long as notopodial ligules. Notopodial ligules subconical, blunt tip, slightly longer than wide; notoacicular lobes subconical, blunt tip, shorter than notopodial ligules, papillae inconspicuous. Neuroacicular lobe subconical, blunt tip, distally bilobed; neuropodial ligules digitiform, slightly longer than neuroacicular lobse. Ventral cirri basal, shorter than neuropodial ligules. In posterior-most chaetigers (Fig. 21G), dorsal cirri half as long as notopodial ligules. Notopodial ligules pennant-like, twice longer than wide, twice longer than notoacicular lobes; notoacicular lobes digitiform, twice longer than neuroacicular lobes, papillae inconspicuous. Neuroacicular lobes subconical, blunt tip, distally bilobed; neuropodial ligules digitiform, slightly longer than neuroacicular lobes. Ventral cirri shorter than neuropodial ligules, both cirri with similar width.

142

Notochaetae homogomph spinigers; neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles. Notopodial homogomph spinigers serrate, teeth decreasing in size toward end (Fig. 21L). Neuropodial homogomph spinigers as notopodial ones; heterogomph spinigers serrate, teeth decreasing in size toward end (Fig. 21K). Heterogomph falcigers pectinate, teeth fine, several times longer than wide, distal tooth stout, supra-acicular falcigers with distal tooth stouter than in sub-acicular ones (Figs 21I, J). Pygidium crenulated; anal cirri subulate, as long as last five segments (Fig. 21B).

Variation. Specimens from Holbox have two paragnaths on areas I and V, and in specimen from Mahahual, dorsal cirri pass slightly notopodial ligule in posterior chaetigers. These differences are not regarded as diagnostic but part of interpopulation variability.

Remarks. de León-González & Solís-Weiss (1998) reported this species from Atlantic Mexican localities; here specimens are from two other localities within the distribution range for the species. Some distinct features of this species are the marginal position of the small eyes, the large tentacular cirri and the two or three paragnaths in a vertical line on area V, not present in other species from the Grand Caribbean. These features allow recognize that some syntypes of Nereis bairdii with long tentacular cirri belong to P. floridana; further comments see remarks of P. bardii above. In specimens examined, neuropodial postchaetal lobes were not clearly observed, and the laterally bilobed tips of neuroacicular lobes give the appearance of a second lobe. Langerhans (1880) made a redescription of the species but his specimens were from Madeira, and he noticed its resemblance to Perinereis ponteni, P. obsfucata, P. striolata and P. cultrifera, but he decided to identify them as P. floridana. Hartman (1938a) considered this as a redescription despite not based upon type or topotype specimens. In specimens from Madeira, notopodial ligules in posterior chaetigers are 1.5 times longer than notoacicular lobe, whereas in P. floridana notopodial ligules are 2.5 times longer. In addition, in specimens from Madeira, ventral cirri are apparently shorter than dorsal cirri throughout body, whereas in P. floridana they are subequal to each other. Moreover, in specimens from Madeira the anterodorsal tentacular cirri are slightly

143 shorter than posterodorsal, whereas in P. floridana the former are half as long as the latter. Monro (1933b) disagreed with Fauvel (1923: 352, 1932: 104-105), and separated this species from P. cultrifera, arguing differences in number and arrangement of paragnaths and insertion of dorsal cirri.

Perinereis mochimaensis Liñero-Arana, 1983

Perinereis mochimaensis Liñero-Arana 1983: 4–5, Figs 3a–f.

Type locality. Isla Larga, Mochima Bay, Venezuela.

Remarks. The report of this species from Mexican Caribbean (Salazar-Vallejo & Jiménez-Cueto 1997) belongs to P. paradisea sp. n.; P. mochimaensis has not been found in other locality, see remarks section of P. cariacoensis above for further commentary.

Perinereis rookery de León-González & Goethel, 2013

Perinereis rookery de León-González & Goethel 2013: 3–7, Figs 1A–I, 2A–B.

Type locality. Rookery Bay, Naples, Florida, at 1.2-1.5 m depth.

Remarks. The species seems to be associated with oyster reefs only. It is related with P. cariboea and P. paradisea sp. n. by having two bars on areas VI, but it can recognized from the other two by the presence of paragnaths on area V, notopodial ligules with similar size along body, and chaetae with a prominent distal tooth, occupying up to half of length of blades.

Distribution. Florida, United States.

144

Perinereis paradisea sp. n. Figure 22A–P

Perinereis mochimaensis Salazar-Vallejo & Jiménez-Cueto 1997: 368-369, Figs 9, 10, 34, 43 (non Liñero-Arana, 1983).

Type material. Caribbean Sea, Mexico. Holotype ECOSUR 0000 and paratype ECOSUR 0000 (1), Paradise Beach, Cozumel Island (20°24'3.19"N 87° 1'5.21"W), Quintana Roo, Mexico, 5 June 1995, Coll. S.I. Salazar-Vallejo. Paratype ECOSUR 0000 (1), Mahahual (18°43'28.14"N 87°42'3.72"W), 18 March 2001, in rocky substrate, Coll. S.I. Salazar-Vallejo.

Additional material. Caribbean Sea, Mexico. ECOSUR P1171 (3), Banco Chinchorro, Quintana Roo, Mexico (18°35'51''N 87°20'20.5''W), 3 October 1983. ECOSUR P1172 (1), El Placer, Quintana Roo, Mexico (18°53'21.36"N 87°38'25.61"W), 26 July 1990. ECOSUR P1113 (1), Punta Nizuc, Quintana Roo, Mexico (21° 02' 11.7" N86° 46' 44.2" W), 4 m, 1 September 1997. ECOSUR P1220 (1), Buenavista, Quintana Roo, Mexico (18°30'39.57"N 87°45'28.05"W), 27 September 1996, Coll. S.I. Salazar-Vallejo, L.F. Carrera-Parra. ECOSUR P0000 (2), Cozumel Island, Quintana Roo, Mexico (20°30'32.78"N 86°57'9.60"W), in rocky substrate, 24 April 2001. Coll. S.I. Salazar-Vallejo. ECOSUR P0000 (2), Xcacel, Quintana Roo, Mexico (20°20'17.2'' N 87°20'50.6'W'), 5 June 1995, Coll. L.F. Carrera-Parra, S.I. Salazar-Vallejo. ECOSUR P0000 (2), Punta Herradura, Quintana Roo, Mexico (18°32'26'' N 87°44'30''W), 4 m depth, 28 October 1997, Coll. S.I. Salazar-Vallejo, L.F. Carrera-Parra. ECOSUR P0000(3), Xcalak, Quintana Roo, Mexico (18°15'56''N 87°49'45''W), 2 m depth, 300 m from coast, in corals, 31 August 2004, Coll. L.F. Carrera-Parra. Jamaica. ECOSUR P0000 (2), R/V Pillsbury, Cruise 7006, St. 1195 (17°24'N 76°02'W), SE Jamaica, 14 m depth, 3 July 1970.

Type locality. Paradise Beach, Cozumel Island, Quintana Roo, Mexico.

Etymology. The Latin word Paradisea resembles the name of the type locality, Paradise Beach, Cozumel, Mexico.

145

Description. Holotype complete, 38 mm long, 1.8 mm wide, 69 chaetigers, tapered posteriorly. Body pale, pigmentation very faint, only visible in tentacular ring. Paratype (ECOSUR 0000) with brown pigmentation; palps with a spot on inner basal margins; four thin, parallel lines arising from bases of antennae, extending towards eyes; Tentacular ring pigmented medially; first chaetiger variegated; second and subsequent ones with two rounded spots, one of each lateral margin of segment, four rectangular spots in rectangle and one central in central region of segment, discoloring posteriorly (Fig. 22C). Prostomium longer than wide; antennae lanceolate, twice longer than wide, not pass the palps; eyes rounded, subequal, in rectangle (Fig. 22A). Tentacular ring 1.5 times longer than first chaetiger; tentacular cirri with cirrophore conspicuous, longest reach chaetiger 2 (Fig. 22A). Pharynx everted, jaws brown with 6 teeth, distal tooth very incurved (Fig. 22D). Maxillary ring: I= 7 cones in round, II= 13-15 cones in arc, III= 12 cones in rectangle, IV= 17-16 cones in round. Oral ring: V= 0, VI= 2-2 smooth bars, VII-VIII= 13 cones in a complete row, slightly alternating, forming a jagged line; and 2 additional paragnaths, displaced posteriorly and laterally (Figs 22E, F); more visible and numerous in other specimens (Fig. 22P). Parapodial cirri pattern: Dorsal cirri slightly longer than notopodial ligule in anterior-most chaetigers, becoming shorter than notopodial ligules throughout body, basally inserted in anterior region, displaced medially in medial chaetigers, subdistal in posterior chaetigers, slightly passing notopodial ligules throughout body. Ventral cirri shorter than neuropodial ligules throughout body, basally inserted in anterior region, progressively distant throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 22G), dorsal cirri basal, subpyriform, subequal to dorsal ligules. Dorsal and neuropodial ligules subconical, blunt tip, longer than wide, subequal, 3 times longer than neuroacicular lobes. Neuroacicular lobes rounded. Ventral cirri subpyriform, shorter than neuropodial ligules. In anterior and middle chaetigers (Figs 22H, I), dorsal cirri subclavate, medial, half as long as notopodial ligules. Notopodial ligules subconical, subequal or slightly longer than notoacicular lobes; notoacicular lobes subconical, longer than wide, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, wider than long; postchaetal lobes rounded, subequal to neuroacicular ones; neuropodial ligules digitiform, twice longer than neuroacicular lobes. Ventral cirri subclavate, shorter than neuropodial ligules.

146

In posterior chaetigers (Figs 22J, K), dorsal cirri subclavate, subdistal, half as long, or one-third as long as notopodial ligules. Notopodial ligules pennant-like, longer than wide, 3–4 times longer than notoacicular lobes; notoacicular lobes subconical, blunt tip, 3 times longer than neuroacicular ones. Neuroacicular lobes rounded, wider than long; postchaetal lobes rounded, subequal to neuroacicular ones; neuropodial ligules digitiform, twice longer than neuroacicular lobes. Ventral cirri subclavate, shorter than neuropodial ligules. Notochaetae homogomph spinigers throughout body; neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular fascicles. Notopodial homogomph spiniger pectinate, finely denticulate, teeth decreasing in size distally (Fig. 22O). Neuropodial homogomph spiniger as notopodial ones, blade longer; heterogomph spiniger pectinate, finely denticulate, teeth decreasing in size distally (Fig. 22N). Heterogomph falcigers falcate, pectinate, teeth 6 times longer than wide, distal tooth stout; supra-acicular falcigers stouter than sub-acicular ones (Figs 22L, M). Pygidium crenulated; anal cirri subulate, as long as last three segments (Fig. 22B).

Variation. Specimens from Punta Herradura have stronger pigmentation than type material. The round glandular mass observed in holotype was not confirmed in other specimens, but some appear in inner margin and tip of notopodial ligules. Specimen from Xcalak shows the incomplete paragnath row (Fig. 22P), with four additional paragnaths.

Remarks. This species resembles P. mochimaensis, but there are some pharyngeal differences between them. In P. paradisea sp. n., area V lacks paragnaths and area VII-VIII have one complete row and rarely some additional paragnaths; whereas in P. mochimaensis, area V have two paragnaths and area VII-VIII have two complete rows. Unfortunately, there is no material available of P. mochimaensis for direct comparison; however, the differences in pharyngeal areas are very conservative in specimens of P. paradisea sp. n., and are considered sufficient to separate them. On the other hand, P. paradisea sp. n. resembles to P. cariboea, but they differ in some features. In P. paradisea sp. n., areas VII-VIII have paragnaths in one row and rarely with few additional paragnaths in a second discontinuous row, whereas in P. cariboea they are in two complete rows.

147

Moreover, in P. paradisea sp. n., the notopodial ligules are 4–5 times longer than notoacicular lobes in posterior chaetigers, while in P. cariboea they are 1.5 to 2 times longer than notoacicular lobes. Finally, in P. paradisea, the notoacicular lobes are 2–3 longer than neuropodial ones throughout body, while in P. cariboea they are subequal.

Distribution. Mexican Caribbean.

Questionable records

Perinereis anderssoni Kinberg, 1865 Figure 23A–K

Perinereis anderssoni Kinberg 1866: 175; de León-González & Solís-Weiss 1998: 675, Figs 1A– G (partim). Nereis minor Hansen 1882: 12, Pl. 4, Figs 8–12.

Type locality. Rio de Janeiro, Brazil.

Material Examined. Brazil. ECOSUR P0000 (7), Praia Vermelha, Rio de Janeiro, 16 August 2012, Coll. M. Coutihno, V. Schawn.

Description. Specimens complete, dorsally pigmented, brown. Largest specimen 23 mm long, 1.9 mm wide, 67 chaetigers, tapered posteriorly. Pigmentation reddish brown in palps, prostomium and chaetigers, more intense in anterior region, progressively discoloring posteriorly. Pale lines present (Fig. 23A), fingerprint-like pattern from chaetiger 13. Prostomium longer than wide, pigmented; antennae cirriform, four times longer than wide, not passing palps; eyes rounded, subequal, in trapezoidal arrangement, posterior pair covered by anterior margin of tentacular ring (Fig. 23A). Tentacular ring 1.5 times longer than first chaetiger; tentacular cirri with cirrophore conspicuous, longest tentacular cirri reach chaetiger 1; antennae and tentacular cirri moniliform, basal segment largest, others decrease in size progressively (Fig. 23A).

148

Pharynx dissected, jaws brown with 10 rounded teeth, distal one very incurved (Fig. 23D). Maxillary ring: I= 5 cones in triangle, II= 11-12 cones in arc, III= 19 cones in rectangle, IV= 19- 33 cones in arc. Oral ring: V= 3 cones in triangle, VI= 1-1 smooth bars, VII-VIII= 39 cones in two jagged rows. Parapodial cirri pattern: Dorsal cirri longer than notopodial ligules in anterior region, becoming shorter toward posterior end, passing notopodial ligules throughout body, basally inserted to dorsal ligules in most-anterior segments, displaced medially in anterior and medial chaetigers, subdistal in posterior chaetigers. Ventral cirri shorter than neuropodial ligules in anterior segments, subequal throughout body, basally inserted in anterior region, progressively distant throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 23E), dorsal cirri 1.5 times longer than dorsal ligules. Dorsal ligules subconical, blunt tip, twice longer than neuroacicular lobes. Neuroacicular lobes rounded, papillae conspicuous; postchaetal lobes rounded, subequal to each other; neuropodial ligules digitiform, subequal to dorsal ligules. Ventral cirri basal, slightly longer than neuropodial ligules. In anterior chaetigers (Fig. 23F), dorsal cirri medial, slightly longer than notopodial ligules. Notopodial ligules subconical, blunt tip, wider than long, subequal to notoacicular lobes; notoacicular lobse digitiform, papillae inconspicuous. Neuroacicular lobes rounded, slightly longer than neuropodial ligules; postchaetal lobes rounded, subequal to neuroacicular one; neuropodial ligules digitiform. Ventral cirri basal, subequal to neuropodial ligules. In middle and posterior chaetigers (Fig. 23G, H), dorsal cirri medial, as long as notopodial ligules. Notopodial ligules subconical, blunt tip, longer than wide, subequal to notoacicular lobse in middle region, longer than in posterior ones; notoacicular lobes subconical, blunt tip, longer than wide, 2–3 times longer than notoacicular lobes, papillae inconspicuous. Neuroacicular lobes rounded; postchaetal lobes rounded, subequal to neuroacicular ones; neuropodial ligules digitiform, twice longer than neuroacicular lobes in middle region, shorter than in posterior one. Ventral cirri basal, as long as neuropodial ligules. In posterior-most chaetigers (Fig. 23I), dorsal cirri subdistal, 0.6 length of notopodial ligules. Notopodial ligules subconical, blunt tip, wider than long, dorsal surface convex, ventral flat, 2–3 times longer than notoacicular lobes; notoacicular lobes subconical, twice longer than neuroacicular lobes, papillae inconspicuous. Neuroacicular lobes rounded; postchaetal lobes

149 rounded, subequal to neuroacicular ones; neuropodial ligules digitiform, subequal to neuroacicular lobes. Ventral cirri basal, subequal to neuropodial ligules. Notochaetae homogomph spinigers throughout body; neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular fascicles. Notopodial homogomph spiniger pectinate, finely denticulate, teeth decreasing in size distally. Neuropodial homogomph spiniger as notopodial ones; heterogomph spiniger pectinate, finely denticulate, teeth decreasing in size distally. Heterogomph falcigers pectinate, teeth fine, several times longer than wide, distal tooth stout; supra-acicular falcigers with distal tooth stouter than those present in sub-acicular ones, with a small bulge in cutting edge (Figs 23J, K). Pygidium crenulated; anal cirri subulate, as long as last three segments (Fig. 23B).

Variation. The specimens studied have little intraspecific variation. The moniliform aspect of dorsal cirri is absent in some specimens.

Remarks. de León-González & Solís-Weiss (1998) redescribed this species based on the best preserved syntype, however they regarded it as amphiamerican. de León-González & Solís-Weiss (1998) included some specimens from Juan Fernandez (Pacific Ocean) and identified them as P. anderssoni; however, they are probably P. pseudocamiguina Augener, 1922, after Hartman (1948) recognized the similarity with P. anderssoni. However, in P. pseudocamiguina the notopodial ligules in posterior chaetigers are longer than wide, and dorsal cirri are as long as notoacicular lobe, whereas in P. anderssoni notopodial ligules are wider than long, with convex dorsum, and dorsal cirri are longer than notoacicular lobe. In order to improve the available redescription, here a second description is provided, including some new illustrations. Moniliform aspect of tentacular cirri and antennae, and fingerprint-like patterns have not been reported anteriorly. After examination of tentacular cirri, the constrictions are only superficial, and not true joints. Pigmentation is similar to N. oligohalina (Rioja) from the Gulf of Mexico (Conde-Vela & Salazar-Vallejo 2015), interestingly both species were collected from Rhizophora mangle roots.

150

Hartman (1948) considered N. bairdii Webster, 1884 and N. minor Hansen, 1882 as junior synonyms to P. anderssoni, after Augener (1934) considered the former as a junior synonym of the latter. However, P. bairdii is herein reinstated but the synonymy with N. minor is retained until further revision of type material. Liñero-Arana & Reyes-Vásquez (1979) reported P. anderssoni from Venezuela; however, there are very conspicuous differences between their materials and P. anderssoni, especially in the larger development of notopodial ligule and neuroacicular lobe, the former several times longer than parapodium itself; also, the number of paragnaths is much higher on areas II, III and IV, and smaller in areas V and VII-VIII . Therefore, Venezuelan specimens belong to another, undescribed species.

Perinereis elenacasoae Rioja, 1947

Perinereis elenacasoi Rioja 1947: 531, Figs 8 –17; Salazar-Vallejo 1989: 50. Perinereis elenacasoae de León-González & Solís-Weiss 1998: 681 –681, Figs 4A –E, 5A-D.

Type locality. Mazatlán, Mexico, Pacific Ocean.

Remarks. de León-González & Solís-Weiss (1998) designated a neotype for this species and their redescription was based on it. However, they included specimens from several localities on both Pacific and Atlantic coasts of Mexico, regarding it as amphiamerican. Localities of the specimens from Mexican Atlantic coasts include Veracruz and Campeche (Gulf of Mexico) and Quintana Roo (Caribbean Sea); these and other records would belong to P. bairdii.

Perinereis vancaurica (Ehlers, 1868)

Nereis languida Grube 1868: 15 –16, Pl. II, Fig. 1. Nereis vancaurica Ehlers 1868: xx (replacement name for N. languida Grube, preoccupied by N. languida Kinberg, 1865).

Type locality. Vancauri (Nicobar Islands), Andaman Sea, Indian Ocean.

151

Material examined. Malaysia. ECOSUR P2248 (1), Kuala Selangor, Malaysia, Indian Ocean, date unknown, Coll. P. Fauvel.

Remarks. Kinberg (1866) preoccupied the specific name languida for a species from Port Jackson, Australia, and Ehlers proposed the replacement name vancaurica. Accordingly with de León-González & Goethel (2013), the descriptions provided by Wesenberg-Lund (1958) and Fauvel (1919, 1923a) from Bonaire and French Guiana, respectively, does not allow to ensure the identification. However, the revision of one specimen from Malaysia allows us to highlight some differences. The description by Wesenberg-Lund (1958) is scarce, but she illustrated a parapodium from middle body. Notopodial ligule and lobe are subequal to each other, neuroacicular lobe is large with postchaetal lobes very prominent, and neuropodial ligule is digitiform; also dorsal and ventral cirri have similar length, but dorsal ones are slightly wider. Conversely, in specimen from Malaysia, the notopodial ligule and lobe are subequal and twice longer than neuroacicular lobe, both neuropodial and postchaetal lobes are subequal to each other, and neuropodial ligule is longer than neuroacicular lobe; also, the dorsal cirrus is twice longer than ventral one. Unfortunately, material from the Grand Caribbean identified as this species was not available, but we anticipate that regional specimens will show some differences to separate them from P. vancaurica.

Platynereis Kinberg, 1865

Type species. Platynereis magalhaensis Kinberg, 1865, by subsequent designation (Hartman 1948).

Diagnosis (after Bakken & Wilson 2005). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary ring with pectinate-bar paragnaths in areas III- IV; missing in areas I-II. Oral ring with pectinate-bar paragnaths, sometimes arising from plake- like basements in areas VI–VIII; area V without paragnaths. First two chaetigers type C. Notopodial ligules similar in size throughout body. Prechaetal notopodial lobe present or absent;

152 neuropodial postchaetal lobe present throughout body. Notochaetae homogomph spinigers and homogomph falcigers; neurochaetae homogomph spinigers and heterogomph falcigers in supra- acicular fascicles, heterogomph falcigers and spinigers in sub-acicular fascicles.

Remarks. Notopodial homogomph falcigers are missing in two species, P australis (Schmarda) and P. hutchingsae de León-González & Solís-Weiss, but in the former homogomph falcigers were found in juveniles and in regenerating posterior ends (Read 2007), the latter imply that loss of homogomph falcigers occur during ontogeny. The species from the Grand Caribbean region are similar among them. There are several doubts about what species must be reinstated from the synonymy with P. dumerilii, mainly by the wide occurrence of homogomph falcigers with a distal tooth or mucron, being not a specific feature. Read (2007) recognized at least three species groups into Platynereis based upon features of notopodial falcigers. In the Grand Caribbean region, the group with bidentate falcigers is present; it can be separated into two subgroups, one with notopodial falcigers present in adults as P. dumerilii and P. magalhaensis, and other lacking notopodial falcigers in adults as P. australis and related species. Platynereis dumerilii Audouin & Milne-Edwards is the only species widely recorded. There are some names available from the Grand Caribbean that could be formally reinstated after a revision of type materials of species involved, mainly Treadwell's species. Meanwhile, here they are considered as valid because they 1) have enough features to distinghish them, and 2) these features are present in Grand Caribbean specimens. Because of this, these species are informally reinstated, pending the study of their type materials, and this explains why ‘inf. reinst.’ was added to emphasize this.

Platynereis antillensis (McIntosh, 1885) inf. reinst. Figure 24A–G

Nereis antillensis McIntosh 1885: 224–225, Pl. 35, Figs 7–9, Pl. 16a, Figs 14–16. Platynereis dumerilii Hartman 1944: 17 (partim, non Audouin & Milne-Edwards, 1834)

Type locality. Off St. Thomas, Virgin Islands, Caribbean Sea.

153

Material examined. Northwestern Atlantic Ocean, Bermuda. ECOSUR P0000 (3), R/V Pillsbury, Cruise 6406, St. 153 (30°18'N 67°00'W), SW Bermuda, 1400 m depth, 4 August 1964. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6406, St. 152 (32°37'N 64°32'W), S Bowditch Seamount, 1100 m depth, 4 August 1964. Sargasso Sea. ECOSUR P0000 (3), R/V Pillsbury, Cruise 6406, St. 160 (29°28'N 70°05'W), NW Atlantic Ocean, 15 m depth, 6 August 1964. United States. ECOSUR P0000 (4), R/V Pillsbury, Cruise 6406, St. 106 (31°10'N 79°19'W), Blake Plateau, off Georgia, 507 m depth, 27 July 1964. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6406, St. 118 (31°52'N 78°16'W), Richardson Hills, off South Carolina, 1 m depth, 29 July 1964. ECOSUR P0000 (1), R/V Gerda, Cruise, St. 279 (25°33’N 79°50’W), Miami Terrace, 252 m depth, 31 March 1964. Bahamas. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6406, St. 169 (27°30'N 76°27'W), Jacksonville Canyon, 1 m depth, 8 August 1964. ECOSUR P0000 (13), R/V Pillsbury, Cruise 6406, St. 171 (27°28'N 76°33'W), Jacksonville Canyon, Bahamas, 1 m depth, 8 August 1964. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6406, St. 195 (27°28’N 78°19’W), N Abacos, 20 m depth, 11 August 1964. Caribbean Sea, Mexico. ECOSUR P1198 (16), Bajo Pepito, Isla Mujeres, Quintana Roo, 1 March 1997, Coll. JE, MD, LY. ECOSUR P0000 (3), R/V Pillsbury, Cruise 670, St. 580 (21°05'N 86°23'W), Arrowsmith Bank, Quintana Roo, ? depth, 22 May 1967. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6703, St. 583 (21°10’N 86°18’W), NE Arrowsmith Bank, Quintana Roo, 27m depth, 23 May 1967. Haiti. ECOSUR P0000 (3), R/V Pillsbury, Cruise 7006, St. 1267 (17°53’N 71°59’W), S Haiti, 0 m depth, 17 July 1970. Saint Kitts and Nevis. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 961 (17°27'N 62°51'W), NW St. Kitts, St. Kitts and Nevis, Lesser Antilles, 11 m depth, 19 July 1969. Colombia. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6504, St. 382 (10°21'N 75°59'W), W Cartagena, Colombia, ? m depth, 14 July 1966. Grenade. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 853 (11°55'N 61°43'W), S Grenada, Lesser Antilles, 17 m depth, 3 July 1969. Guadeloupe. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6907, St. 937 (16°10'N 61°18'W), S St. Francois, 11 m depth, 16 July 1969.

Description. Specimens complete or incomplete; one complete used for decription (ECOSUR P0000); Body pale, 15 mm long, 2 mm wide, 72 chaetigers. (Figs 24A, B). Prostomium cordiform, as long as wide, anterior margin entire; antennae cirriform, as long as prostomium; eyes round, reddish, anterior pair slightly larger than posterior one, in trapezoidal

154 arrangement (Fig. 24A). Tentacular ring 1.5 times longer than first chaetiger; four pairs of tentacular cirri, longest reach chaetiger 13 (Fig. 24A). Pharynx dissected; jaws brown, 10 teeth, cutting edge completely dentate. Maxillary ring: I= 0, II= 0, III= 0, IV= 4-? pectinate rows in one patch. Oral ring: V= 0, VI= 3 pectinate rows in one patch, VII-VIII= 5 patches of three pectinate rows each. Parapodial cirri pattern: dorsal cirri longe than notopodial ligules throughout body, basally inserted in anterior-most chaetigers, displaced medially toward posterior chaetigers. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 24C), dorsal cirri basal, twice longer than notopodial ligules. Notopodial ligules subconical, blunt tip, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, half as long as postchaetal lobes; postchaetal lobes digitiform, slightly longer than neuropodial ligules; neuropodial ligules digitiform. Ventral cirri basal, 3/4 as long as neuropodial ligules. In anterior chaetigers (Fig. 24D), dorsal cirri medial, 2.5 times longer than notopodial ligules. Neuropodial ligules subconical, blunt tip, as long as notoacicular lobes; notoacicular lobes subconical, blunt tip, twice longer than neuroacicular ones. Neuroacicular lobes digitiform, half as long as postchaetal lobes; postchaetal lobes digitiform, slightly longer than neuropodial ligules; neuropodial ligules digitiform. Ventral cirri basal, as long as neuropodial ligules. In middle chaetigers (Fig. 24E), dorsal cirri medial, 1.2 times longer than notopodial ligules. Notopodial ligules subconical, blunt tip, as long as notoacicular lobes; notoacicular lobes subconical, blunt tip, twice longer than neuroacicular ones. Neuroacicular lobes digitiform, as long as postchaetal lobes and neuropodial ligules; postchaetal lobes digitiform; neuropodial ligules digitiform. Ventral cirri basal, 3/4 as long as neuropodial ligules. In posterior chaetigers (Fig. 24F), dorsal cirri medial, 1.2 times longer than notopodial ligules. Notopodial ligules subconical, blunt tip, enlarged, 1.5 times longer than notoacicular ones; notoacicular lobes digitiform, blunt tip, enlarged, 3 times longer than neuroacicular ones. Neuroacicular lobes digitiform; postchaetal lobes inconspicuous; neuropodial ligules digitiform, twice longer than neuroacicular lobes. Ventral cirri basal, 3/4 as long as neuropodial ligules. Notochaetae homogomph spinigers and falcigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular ones.

155

Notopodial homogomph spinigers pectinate, fine teeth, decreasing toward distal end; homogomph falcigers stout, pectinate, 2–4 teeth restricted to basal potion of the blades, with a tendon fused to inner margin, with a mucron (Fig. 24G). Pygidium tripartite; anal cirri subulate, as long as last 13 chaetigers (Fig. 24B).

Remarks. Hartman (1944) synonymized P. antillensis with P. dumerilii (Audouin & Milne- Edwards, 1834) without any explanation. The specimens examined match the description by McIntosh, especially in the paragnaths arrangement and parapodial proportions. Further, P. antillensis resembles two other species: P. trimaculosa (Treadwell, 1940) and P. lutea Treadwell, 1928; it is probable that P. trimaculosa becomes a junior synonym of P. antillensis, but P. lutea is distinct. The most reliable features to separate these species, with current information, are in paragnaths arrangement. In P. antillensis, areas I, II, III and V lack paragnaths, whereas in P. trimaculosa paragnaths are absent on areas I, II and V, but parapodial proportions are similar in 10th and 26th chaetigers as described by McIntosh. In the P. antillensis specimens the notopodial ligules are longer than notoacicular lobes in middle and posterior chaetigers; whereas in P. trimaculosa, they are subequal. Moreover, in P. antillensis, dorsal cirri from posterior chaetigers are longer than notopodial ligules, whereas in P. trimaculosa they are subequal. Additional differences might be noticed after revision of types.

Distribution. Northwestern Atlantic Ocean, Caribbean Sea.

Platynereis hutchingsae de León-González, Solís-Weiss & Valadez-Rocha, 2001 Figure 25A–I

Platynereis hutchingsae de León-González et al. 2001: 389–391, Figs 1a–h, 3a, b.

Type locality. Laguna de Términos, Campeche, Gulf of Mexico.

Type material. Gulf of Mexico, Mexico. Holotype UANL 4281 and paratype UANL 4282 (1), San Julián, Laguna de Términos, Campeche, 3 March 1984, Coll. V. Solís-Weiss et al.

156

Description. Holotype in regular conditions, body wall macerated, fragile, 13 mm long, 0.9 mm wide, 79 chaetigers; pharynx damaged, paragnaths not seen; paratype anterior end in regular conditions, 6 mm long, 1 mm wide, 21 chaetigers. Body pale, without pigmentation (Figs 25A– D). Prostomium longer than wide, anterior margin entire; antennae cirriform, half as long as prostomium; eyes subequal, rounded, black, in rectangular arrangement (Fig.25 C). Tentacular ring 1.5 times longer than first chaetiger; four pairs of tentacular cirri, longest one reaches chaetiger 5 (fragile in paratypes) (Figs 25B, C). Pharynx dissected previously; jaws brown, with 11 teeth, cutting edge completely dentate. Paragnaths not seen due to poor conditions of specimens. Parapodial cirri pattern: Dorsal cirri shorter and passing notopodial ligules throughout body, basally inserted in anterior-most chaetigers, displaced medially throughout body. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body. First two chaetigers type C, remaining biramous. In anterior chaetigers (Fig. 25E), dorsal cirri medial, 3/4 a long as notopodial ligules. Notopodial ligules subconical, slightly longer than notoacicular lobes; notoacicular lobes rounded to subconical, blunt tip, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, slightly longer than neuropodial ligules; postchaetal lobes digitiform, twice longer than neuroacicular lobes; neuropodial ligules rounded. Ventral cirri basal, 3/4 as long as neuropodial ligules. In posterior chaetigers (Fig. 25F), dorsal cirri medial, 3/4 as long as notopodial ligules. Notopodial ligules subconical; notoacicular lobes subconical to digitiform, slightly longer than notopodial ligules, 1.2 times longer than neuroacicular ones. Neuroacicular lobes subconical, 1.8 times longer than neuropodial ligules; neuropodial ligules digitiform. Ventral cirri basal, half as long as neuropodial ligules. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular ones. Neuropodial heterogomph falcigers falcate, serrate, half of inner edge dentate, distal tooth stout; supra-acicular falcigers stouter than sub-acicular ones (Figs 25G, H). Pygidium tripartite; anal cirri cirriform, as long as last two chaetigers (Fig. 25D).

157

Remarks. Type specimens have pharynx dissected, and the whereabouts of the specimen with pharynx everted (de León-González et al. 2001, Fig. 3) was not declared; due to the actual state of pharynx on specimens, paragnaths and their arrangement could not be confirmed, and additional dissections were avoided. This species lacks notopodial homogomph falcigers in posterior chaetigers, confirmed by examination under microscope of whole-mounted specimens. In addition, P. hutchingsae lacks the middorsal nuchal cushion found in other species such as P. australis and related species. de León-González et al. (2001) argued that this species and P. australis are the only species lacking notopodial falcigers, but Read (2007) found notopodial homogomph falcigers in juveniles of P. australis; therefore the revision of more specimens of P. hutchingsae might result in showing notopodial homogomph falcigers.

Distribution. Gulf of Mexico.

Platynereis lutea (Treadwell, 1928) inf. reinst. Figure 26A–J

Uncinereis lutea Treadwell 1928: 469–471, Fig. 179, Figs 40–49. Platynereis dumerilii Hartman 1956: 281 (partim, non Audouin & Milne-Edwards, 1834)

Type locality. Sargasso Sea (26°10’N 56°W), Northwestern Atlantic Ocean, on Sargassum.

Material examined. Northwestern Atlantic Ocean, Sargasso Sea. ECOSUR P0000 (3), R/V Pillsbury, Cruise 6406, St. 160 (29°28'N 70°05'W), NW Atlantic Ocean, on Sargassum sp., 15 m depth, 6 August 1964.

Description. Specimen complete, in two portions, 78 chaetigers, found inside a thick, membranous tube (Fig. 26B). Body pale, without pigmentation. Another complete specimen, inside a thinner tube, attached to a Sargassum leaf-like lamina (Fig. 26C).

158

Prostomium as long as wide; antennae cirriform, 1.5 times longer than prostomium; eyes rounded, subequal, reddish, in trapezoidal arrangement (Fig. 26A). Tentacular ring as long as first chaetiger; four pairs of tentacular cirri, longest reach chaetiger 22 (Fig. 26A). Pharynx dissected. Maxillary ring: I= 0, II= 0, III= 3 pectinate rows, IV= 0. Oral ring: V=0, VI= two pectinate rows; VII-VIII= 6 pectinate rows, in a discontinuous row. Parapodial cirri pattern: dorsal cirri longer than notopodial ligules throughout body, basally inserted in anterior-most chaetigers, displacing medially toward posterior chaetigers. Ventral cirri as long as neuropodial ligules in anterior-most chaetigers, becoming shorter than toward posterior end, basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 26F), dorsal cirri basal, 2.5 times longer than notopodial ligules. Notopodial ligules subconical, blunt tip, 2–3 times wider and 4–5 times longer than neuroacicular lobes. Neuroacicular lobes subconical; postchaetal lobes digitiform, twice longer than neuroacicular ones; neuropodial ligules digitiform, half as long as notopodial ones. Ventral cirri basal, as long as neuropodial ligules. In anterior chaetigers (Fig. 26G), dorsal cirri medial, 3–4 times longer than notopodial ligules. Notopodial ligules digitiform, as long as notoacicular lobes; notoacicular lobes rounded, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, as long as postchaetal lobes; postchaetal lobes digitiform; neuropodial ligules rounded, as long as neuroacicular lobes. Ventral cirri basal, half as long as neuropodial ligules. In middle and posterior chaetigers (Figs 26H–J), dorsal cirri medial, twice longer than neuropodial ligules, 1.5 times longer than in posterior chaetigers. Notopodial ligules subconical, blunt tip, as long as notoacicular lobes and 1.5 times longer than in posterior chaetigers; notoacicular lobes subconical, blunt tip, 3–4 times longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip; postchaetal lobes, as long as neuroacicular ones, inconspicuous in posterior chaetigers; neuropodial ligules digitiform, 1.2 times longer than neuroacicular lobes. Ventral cirri basal, 3/4 as long as neuropodial ligules. Notochaetae homogomph spinigers and falcigers. Neurochaetae homogomph spiniger and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular ones. Notopodial homogomph falcigers stout, pectinate, 3–6 teeth restricted to basal potion of the blades, with a tendon fused to inner edge; falcigers from middle chaetigers without a distal tooth

159 and stout blades (Fig. 25D), falcigers from posterior chaetigers with a mucron and narrow blades (Fig. 26E). Pygidium tripartite, anal cirri missing.

Remarks. Treadwell (1928) described Uncinereis lutea based upon specimens collected in the Sargasso Sea on Sargassum; there are other records from the Bahamas of P. dumerilii on Sargassum (Renaud 1956). The notopodial homogomph falciger drawn by Treadwell has a rounded distal end, it has no distal tooth or teeth on inner edge as he noted in the description, and it appears since chaetiger 16 in a non-type specimen (Treadwell 1928). The specimens examined have falcigers with almost smooth distal end as described by Treadwell in the first chaetigers with them, but they are replaced with falcigers with distal tooth toward posterior chaetigers. Probably Treadwell could not detect the minute distal tooth since the magnification he used (x250), even in farther magnifications (400x) it is difficult to be seen. Hartman (1945, 1956) synonymized several species with P. dumerilii, including three of Treadwell’s species (U. lutea, U. trimaculosa and Eunereis africana), arguing that they are “typical representatives”. The three species have some differences among them that allow their separation. P. lutea resembles P. trimaculosa in the shape of blades of notopodial homogomph falcigers, but they differ in some importan features. In P. lutea, the dorsal cirri are much longer throughout body than in P. trimaculosa. Furthermore, in P. lutea, the notopodial ligules are longer and wider than notoacicular lobes in middle and posterior chaetigers, whereas in P. trimaculosa they are subequal in length and width. Moreover, in P. lutea, the notopodial homogomph falcigers are narrower and longer in posterior chaetigers than in P. trimaculosa. This latter feature seems to be variable along body, requiring further analysis; however, the other features are sufficient to separate them.

Distribution. Sargasso Sea.

Platynereis mucronata de León-González, Solís-Weiss & Valadez-Rocha, 2001 Figure 27A–J

160

Platynereis mucronata de León González, Solís-Weiss & Valadez-Rocha 2001: 391–394, Figs 2a–h, 3c, d.

Type locality. La Pesca, Tamaulipas, Gulf of Mexico.

Material examined. Gulf of Mexico, Mexico. ECOSUR P0000 (5), Playa Manigua, Ciudad del Carmen, Campeche, 9 May 2000, on rocks, Coll. J.R. Bastida-Zavala, S.I. Salazar-Vallejo.

Description. Specimens complete, one incomplete; specimen used for description atokous female, complete, in two portions, 33 mm long, 2 mm wide, 87 chaetigers, filled with oocytes; body pale, without pigmentation (Fig. 27A). Prostomium as long as wide, margin entire; antennae cirriform, as long as prostomium, slightly passing palps; eyes rounded, subequal, black, in rectangular arrangement (Fig. 27A). Tentacular ring 1.5 times longer than first chaetiger (Fig. 27B); four pairs of tentacular cirri, incomplete (Fig. 27A), in a complete specimen longest reach chaetiger 8. Pharynx dissected; jaws brown with teeth. Maxillary ring: I= 0, II= 0, III= 3 small pectinate bars, IV= 4-5 pectinate bars, some discontinuous, in triangle. Oral ring: V= 0, VI= 2-2 pectinate bars, VII-VIII= 5 pectinate bars, in a discontinuous row. Parapodial cirri pattern: dorsal cirri longer than notopodial ligules in anterior chaetigers, becoming as long as notopodial ligues in posterior chaetigers, basally inserted in anterior-most chaetigers, displaced medially toward posterior chaetigers. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body. First two chaetigers uniramous, remaining ones biramous. In first two chaetigers (Fig. 27F), dorsal cirri twice longer than notopodial ligules. Notopodial ligules digitiform, twice longer than neuroacicular ones. Neuroacicular ones subconical, blunt tip, half as long as postchaetal lobes; postchaetal lobes digitiform; neuropodial ligules digitiform, as long as neuroacicular lobes. In anterior chaetigers (Fig. 27G), dorsal cirri basal, 1.2 times longer than notopodial ligules. Notopodial ligules subconical, blunt tip, as long as notoacicular lobes; notoacicular lobes digitiform to rounded, as long as neuroacicular ones. Neuroacicular lobes subconical, blunt tip, slightly longer than neuropodial ligules; postchaetal lobes as long as neuroacicular ones; neuropodial ligules digitiform. Ventral cirri basal, 3/4 as long as neuropodial ligules.

161

In middle and posterior chaetigers (Figs 27H–J), dorsal cirri basal, as long as notopodial ligules. Notopodial ligules subconical to digitate, as long as notoacicular ones; notoacicular lobes digitiform to subconical, 2–3 times longer than neuroacicula ones. Neuroacicular lobes subconical, blunt tip; postchaetal lobes as long as neuroacicular ones, in posterior chaetigers reduced, papillae-like; neuropodial ligules digitiform, slightly longer than neuroacicular lobes. Ventral cirri basal, half as long as neuropodial ligules. Notochaetae homogomph spinigers and falcigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular fascicles. Notopodial homogomph spinigers pectinate, fine teeth, decreasing toward distal end. Notopodial homogomph falcigers stout, pectinate, 3–6 teeth restricted to basal portion of the blades, with a tendon fused to inner margin, with a mucron, increasing slightly in length and number of teeth toward posterior chaetigers (Fig. 27G). Neuropodial heterogomph falcigers pectinate, fine teeth, with a tendon fused to inner margin, 1/3 of inner edge dentate; blades increasing their length toward posterior end, supra-acicular ones stouter than sub-acicular ones (Figs 27C, D). Pygidium tripartite; anal cirri cirriform, as long as last 11 segments (Fig. 27B).

Remarks. The specimens examined match well with the description available. Only the notopodial homogomph falcigers differs little; in specimens examined these chaetae are distally stouter than the one illustrated by de León-González et al. (2001). On the other hand, this species would be a junior synonym of a Treadwell's species, P. trimaculosa (Treadwell, 1940), discussed below. The most reliable differences between both species are the follow. In P. mucronata, notoacicular lobes in anterior chaetigers are as wide as neuroacicular ones, whereas in P. trimaculosa, they are much wider than neuroacicular ones; also, neuropodial heterogomph falciger illustrated by Treadwell (1940, Fig. 7) does not have a distal tooth with tendon as in P. mucronata. In other features are very similar, as the reach of tentacular cirri, and the proportions of dorsal cirri, notopodial ligules and notoacicular lobes. P. mucronata also resembles P. antillensis and P. lutea, and the differences indicated above among the last and P. trimaculosa, are valid for P. mucronata as well.

162

Platynereis sp. 1 sp. n. Figure 28A–N

Type material. Caribbean Sea, Mexico. Holotype USNM 0000 and paratypes USNM 0000 (8), R/V Gerda, Cruise 6717, St. 894 (21°11'N 86°19'W), Arrowsmith Bank, Quintana Roo, 190 m depth, 10 September 1967. Paratypes UMML P0000 (3), R/V Pillsbury, Cruise 670, St. 580 (21°05'N 86°23'W), Arrowsmith Bank, Quintana Roo, ? depth, 22 May 1967.

Additional material. Northwestern Atlantic Ocean, Bermuda. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6406, St. 133 (32°06’N 65°05’W), Challenger Bank, Bermuda, 59 m depth, 1 August 1964. Caribbean Sea, Mexico. ECOSUR P0000 (5), R/V Pillsbury, Cruise 6703, St. 581 (21°05'N 86°23'W), Arrowsmith Bank, Quintana Roo, 205 m depth, 22 May 1967. St. Barthélemy. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 981 (18°01'N 62°55'W), Saint-Barthélemy Channel, Lesser Antilles, 22 m depth, 22 July 1969.

Type locality. Arrowsmith Bank, Quintana Roo, Mexico, 190 m depth.

Description. Holotype complete, 45 m long, 3.2 mm wide, 79 chaetigers, tentacular and anal cirri incomplete. Paratypes some complete, in good condition. Body pale, without pigmentation (Fig. 28A). Prostomium cordiform, as long as wide; antennae as long as prostomium; eyes rounded, anterior pair slightly larger than posterior one, anterior pair very close to posterior ones (Fig. 28A). Tentacular ring 1.5 times longer than first chaetiger, middorsal nuchal cushion present, forming a small acute bridge; four pairs of tentacular cirri, incomplete, longest one reaches chaetiger 10 (Fig. 28A). Pharynx dissected; jaws amber, 15 teeth, cutting edge completely dentate (Fig. 28C). Maxillary ring: I= 0, II= 0, III= 0, IV= 4-5 pectinate rows. Oral ring: V= 0, VI= 0, VII-VIII= 3 groups of three rows each. Parapodial cirri pattern: Dorsal cirri shorter than notopodial ligules throughout body, basally inserted in anterior-most chaetiger, displaced medially toward posterior end, passing notopodial

163 ligules throughout body. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 28D), dorsal cirri basal, as long as notopodial ligules. Notopodial ligules digitiform, as long as nuropodial postchaetal lobes. Neuroacicular lobes rounded, 1/3 as long as postchaetal lobes; postchaetal lobes digitiform, twice longer than neuropodial ligules; neuropodial ligules digitiform. Ventral cirri basal, as long as neuropodial ligules. In anterior chaetigers (Fig. 28E), dorsal cirri medial, 1.2 times longer than notopodial ligules. Notopodial ligules subconical, blunt tip, as long as notoacicular lobes; notoacicular lobes subconical, blunt tip, papillae conspicuous. Neuroacicular lobes rounded to subconical; postchaetal lobes digitiform, twice longer than neuroacicular ones; neuropodial ligules digitiform, as long as notoacicular lobes. Ventral cirri basal, swollen base, 3/4 as long as neuropodial ligules. In middle and posterior chaetigers (Figs 28F, G), dorsal cirri medial, 3/4 as long as notopodial ligules. Notopodial ligules subconical, blunt tip, as long as notoacicular lobes; notoacicular lobes ovate, three times longer than neuroacicular ones, papillae conspicuous. Neuroacicular ligules subconical, blunt tip, as long as postchaetal lobes; postchaetal lobes digitiform; neuropodial ligules digitiform, slightly longer than notoacicular lobes. Ventral cirri basal, 3/4 as long as neuropodial ligules. In posterior-most chaetigers (Fig. 28H), dorsal cirri medial, 3/4 as long as notopodial ligules. Notopodial ligules subconical to digitiform, twice longer than notoacicular lobes; notoacicular lobes ovate, 4 times longer than neuroacicular lobes, papillae conspicuous. Neuroacicular lobes subconical, blunt tip, as long as postchaetal lobes; postchaetal lobes subconical; neuropodial ligules digitiform, twice longer than neuroacicular lobes. Ventral cirri basal, 3/4 as long as neuropodial ligules. Notochaetae homogomph spinigers and falcigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub- acicular fascicles. Notopodial homogomph spinigers pectinate, coarse teeth basally, decreasing in size toward distal end (Fig. 28K). Notopodial homogomph falcigers several times longer than wide, with a slight projection on dorsal end, tendon short, almost all inner edge dentate (Fig. 28N). Neuropodial homogomph spinigers as notopodial ones, blades shorter than notopodial ones; heterogomph

164 spinigers pectinate, fine teeth, decreasing in size toward distal end. Neuropodial heterogomph falcigers pectinate, fine teeth, half in inner edge dentate, distal tooth coarse, with a tendon reaching the half of inner edge; supra-acicular falcigers stouter than sub-acicular ones (Figs 28L, M). Pygidium crenulated; anal cirri incomplete, as long as last 10 chaetigers (Fig.28B).

Variation. Holotype and most paratypes from same station lack notopodial homogomph falcigers throughout body, just one specimen has notopodial falcigers, restricted to posterior-most chaetigers (Figs 28I, J). This is an unusual feature of this species, because in other species usually notopodial falcigers arise from 18-20 chaetigers. Holotype is the best-preserved, complete specimen and was preferred over incomplete specimens with notopodial falcigers; all other chaetal and parapodial features are shared with remaining specimens. Notopodial falciger appears since chaetiger 67 (of 98), in paratype, and in the smallest specimen (USNM P0000), the notopodial falcigers appear from chaetiger 22 (of 43) and change little their length. Conversely, in all specimens neuropodial falcigers tend to decrease in size toward posterior chaetigers.

Remarks. The shape of notopodial falcigers and their absence in most parapodia resemble P. australis and related species (Read 2007). The main differences among these species are in the heteronereis, along with pigmentation pattern in atokes. Also, Read (2007) found that the notopodial homogomph falcigers are found in juveniles only, and the appearing of the homogomph falcigers becomes more posterior according to the adition of new chaetigers in juveniles, as the case of specimen with 43 chaetigers. However, paratype with 98 chaetigers does not resemble to a juvenile stage. However, P. sp. 1 sp. n. differs from P. australis and related species in the following features. In P. australis paragnaths are present in areas III and VI, whereas P. sp. 1 sp. n. lacks paragnaths on these areas; furthermore, P. australis have several pectinate rows on areas IV, whereas P. sp. 1 sp. n. has a few rows. In addition, in P. sp. 1 sp. n., the notopodial homogomph falcigers have rounded or almost distal tooth, while in P. australis they have a distal tooth; also, in P. sp. 1 sp. n. the tendon is 1/5 as long as the inner margin, while in P. australis it is 1/3.

165

P. sp. 1 sp. n. resembles P. hutchingsae but they differ because P. sp. 1 sp. n. has a middorsal nuchal cushion, while P. hutchingsae lacks this feature. Further, in P. sp. 1 sp. n. the dorsal cirri are relatively longer throughout body than in P. hutchingsae. Moreover, in P. sp. 1 sp. n., the heterogomph falcigers have a well-developed distal tooth and a tendon, but these features are missing in P. hutchingsae.

Distribution. Bermuda, Mexican Caribbean, St. Barthélemy.

Species requiring further revision

Platynereis trimaculosa (Treadwell, 1940)

Uncinereis trimaculosa Treadwell 1940: 3, Figs 3-9.

Type locality. Galvestone, Texas, Gulf of Mexico.

Remarks. Treadwell (1940) described Uncinereis trimaculosa with specimens found in a floating log in Galveston, Texas, and Hartman (1956) synonymized it (and other species) with P. dumerilii. The notopodial homogomph falciger with a distal tooth and parapodia morphology resembles to P. mucronata. The latter and the fact that the type localities of both species are near to each other indicate that P. mucronata would be conspecific with P. trimaculosa, becoming a junior synonym; however, in order to clarify these doubts the revision of type material of both species is needed, meanwhile both names are provisionally retained.

Questionable records

Platynereis dumerilii (Audouin & Milne–Edwards, 1834)

Type locality. La Rochelle, France, North Sea.

Remarks. The name was introduced in several reports (Dean 2012; Salazar-Vallejo 1996). This species has a confused, problematic taxonomic history; problems must be solved to clarify the

166 taxonomic status of synonymized species and the possible proposal of new species for the Grand Caribbean and others regions.

Platynereis coccinea (delle Chiaje, 1822)

Type locality. Naples, Mediterranean Sea.

Remarks. The name was introduced by Amaral & Nonato (1975). This species has a confused, problematic taxonomic history; problems must be solved to clarify the taxonomic status of synonymized species and the possible proposal of new species for the Grand Caribbean and others regions.

Pseudonereis Kinberg, 1865

Type species. Pseudonereis gallapagensis Kinberg, 1865, by subsequent designation (Hartman 1948). Diagnosis (after Bakken 2007). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary ring with conical and p-bars paragnaths, arranged in comb-like series in areas I-IV, sometimes missing. Oral ring with conical, p-bars or smooth bars paragnaths in areas VII-VIII, areas V–VI with or without paragnaths. First two chaetigers type C. Notopodial ligules very projected, pennant-like toward posterior chaetigers. Prechaetal notopodial lobe present or absent; neuropodial postchaetal lobe present at least in anterior chaetigers. Notochaetae homogomph spinigers, rarely homogomph falcigers; neurochaetae homogomph spinigers and heterogomph spinigers and falcigers on supra-acicular fascicles, heterogomph falcigers and spinigers in sub-acicular fascicles.

Remarks. Although some species have been recorded for the Grand Caribbean region, all have been regarded as synonyms of two species: P. gallapagensis and P. variegata both described from Pacific South American coasts. Therefore, they are the only species recorded from Caribbean coasts (Dean 2012).

167

After the revision of paragnath morphology, some terms are changed here. Currently, Area VI is reported bearing conical shield-shaped paragnaths (Bakken 2007); however, the species described here have a flat, acute paragnath with a rectangular base, which fits the current definition of pointed paragnaths (P-bars). Therefore, P-bars in Area VI are included in the diagnosis.

Pseudonereis citrina sp. n. Figure 29A–P

Pseudonereis gallapagensis Liñero-Arana & Reyes-Vásquez 1979: 8-9, Pl. 7, Figs 1–10. Salazar- Vallejo & Jiménez-Cueto 1997: 369, Figs 35, 36, 51. (non Kinberg, 1865)

Type material. Caribbean Sea, Mexico. Holotype ECOSUR 0000, Mahahual, Quintana Roo, Mexico (18°43'27.06"N 87°42'1.53"W), 19 January 2001, in corals, Coll. S.I. Salazar-Vallejo. Paratypes ECOSUR 0000 (2), Akumal, Quintana Roo, Mexico (20°23'38.41"N 87°18'55.61"W), 23 February 1986, Coll. M Aguilar, A. Gómez. ECOSUR 0000 (2), Xamach, Quintana Roo, Mexico (19°55'36.85"N 87°26'9.17"W), 28 February 1986, Coll. L. Aguilar, J. Fernández. Additional material. Caribbean Sea, Mexico. ECOSUR P1108 (1), Mahahual, Quintana Roo, Mexico (18°43'27.06"N 87°42'1.53"W), 20 May 1997, Coll. S.I. Salazar-Vallejo, L.F. Carrera- Parra. ECOSUR P1222 (1), Playa Sol y Mar, Quintana Roo (18°46'36''N 87°40'14.4''W), 18 January 2000, Coll. S.I. Salazar-Vallejo, L.F. Carrera-Parra. ECOSUR P2278 (1), Vigía Chico, Quintana Roo, Mexico (19°46'22.84"N 87°35'11.38"W), 27 February 1986, Coll. S.I. Salazar- Vallejo. Panama. USNM 73683 (2), Galeta Reef Flat, Canal zone, 1973, Coll. H.O. Brattstrom.

Type locality. Mahahual, Quintana Roo, Mexico (18°43'27.06"N 87°42'1.53"W).

Etymology. The specific name stems from the Latin adjective citrinus, -a, -um, meaning lemon yellow, referring to the yellowish body and the very conspicuous, yellow glandular masses in notopodial ligules from posterior chaetigers.

Description. Holotype complete, 19 mm long, 1.5 mm wide, 62 chaetigers. Body pale, yellowish, slightly transparent in some areas, tapered, damaged by chaetiger 30.

168

Prostomium longer than wide; antennae subulate, half as long as prostomium, slightly passing the palps; eyes rounded, subequal, black, in trapezoidal arrangement (Fig. 29A). Tentacular ring 1.5 times longer than first chaetiger; tentacular cirri with cirrophore conspicuous, longest tentacular cirri reach chaetiger 5 (Fig. 29A). Pharynx dissected; jaws brown, 13 teeth, distal end very incurved, cutting edge completely dentate (Fig. 29C). Maxillary ring: I= 3 cones in vertical line, II= 4-5 pectinate bars in triangle, III= 4 pectinate bars in triangle, IV= 4-4 pectinate bars and additional cones and smooth bars, in sigmoid. Oral ring: V= 0, VI= 1-1 P-bar, VII-VIII= 21 cones in a row, alternating, forming a jagged line (Figs 29D, E). Parapodial cirri pattern: Dorsal cirri several times longer than notopodial ligules in anterior and middle body, largest in middle body, and shorter than in posterior-most chaetigers, basally inserted to dorsal ligules in most-anterior segments, displaced medially in anterior and medial chaetigers, subdistal in posterior chaetigers, and distally in posterior-most ones. Ventral cirri longer than neuropodial ligules in anterior-most chaetigers; subequal from anterior to posterior chaetigers, slightly shorter in posterior-most segments, basally inserted in anterior-most region, becoming distal in anterior segments, progressively closer in posterior region. First two chaetigers type C, remaining biramous; glandular masses on tips of ligules and lobes. In first two chaetigers (Fig. 29F), dorsal cirri basal, three times longer than dorsal lobes. Notopodial ligules subconical, blunt tip, as long as neuropodial ligules, as long as wide, slightly longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, twice wider than long; postchaetal lobes rounded, as long as neuroacicular ones; neuropodial ligules digitate. Ventral cirri basal, slightly longer than neuropodial ligules. In anterior and middle chaetigers (Figs 29G, H), dorsal cirri medial, 2–3 times longer than notopodial ligules. Notopodial ligules subconical, blunt tip, shorter than notoacicular lobes; notoacicular lobes subconical, blunt tip, shorter than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, twice wider than long, twice longer than neuropodial ligules; postchaetal lobes rounded, as long as neuroacicular ones; neuropodial ligules digitiform. Ventral cirri basal, subequal to neuropodial ligules. In posterior chaetigers (Fig. 29I), dorsal cirri subdistal, 1.5 times longer than notopodial ligules. Notopodial ligules subconical, blunt tip, shorter than notoacicular lobes; notoacicular lobes subconical, blunt tip, subequal to neuroacicular lobes. Neuroacicular lobes subconical, blunt tip,

169 twice wider than long, twice longer than neuropodial ligules; postchaetal lobes inconspicuous; neuropodial ligules digitiform. Ventral cirri subulate, shorter than neuropodial ligules. In posterior-most chetigers (Fig. 29J), dorsal cirri distal, 0.3 times length of notopodial ligules. Notopodial ligules pennant-like, several times longer than notoacicular lobes; notoacicular lobes subconical, blunt tip, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, 1.5 times wider than long, twice longer than neuropodial ligules; postchaetal lobes inconspicuous; neuropodial ligules digitiform. Ventral cirri subulate, shorter than neuropodial ligules. Notochaetae homogomph spinigers. Neurochaetae spinigers and heterogomph falcigers in supra- acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles. Notopodial homogomph spinigers pectinate, coarse teeth, decreasing in size toward end (Fig. 29O). Neuropodial homogomph spinigers as notopodial ones; heterogomph spinigers serrate, teeth basally coarse, decreasing in size toward distal end (Figs 29K, P). Neuropodial heterogomph falcigers falcate, pectinate, coarse teeth, half of inner edge edentate, distal tooth stout; supra-acicular falcigers stouter than sub-acicular ones, and falcigers become stouter toward posterior chaetigers (Figs 29L–N) Pygidium crenulated; anal cirri subulate, as long as last five chaetigers (Fig. 29B).

Variation. Specimens examined vary practically in size only. The paratypes have a minute cone in area V (Figs 29D, E).

Remarks. The yellowish colour of body in this species and the spotted notopodial ligules were observed in all material examined. The small size together with the color and enlarged dorsal cirri allow recognize easily to this species. Pseudonereis citrina sp. n. resembles P. gallapagensis, but they have several differences. Gravier (1909) provided a detailed description of P. gallapagensis based upon specimens collected near its type locality. The more evident difference is that in P. citrina sp. n. the dorsal cirri in anterior chaetigers are three times longer than notopodial ligules, and they extend 6–7 times beyond of notopodial ligules, whereas in P. gallapagensis the dorsal cirri in the same region are 2 times longer than notopodial ligules and they extend 3–4 times beyond notopodial ligules. Another difference is that in P. citrina sp. n. notopodial ligules from posterior chaetigers have several

170 yellow, glandular masses, whereas in P. gallapagensis they are only vascularized but not glandular.

Pseudonereis brunnea sp. n. Figure 30A–O

Type material. Gulf of Mexico, Mexico. Holotype ECOSUR 0000, La Mancha, Veracruz (19°36'0.30"N 96°22'19.61"W), 27 May 2004. Paratypes ECOSUR 0000 (6), La Mancha, Veracruz (19°35'10.16"N 96°22'38.88"W), 3 July 2003. ECOSUR P0000 (3), La Mancha, Veracruz, Mexico, 23 July 2003. Additional material. Gulf of Mexico, Mexico. ECOSUR P0000 (12), La Mancha, Veracruz (19°35'7.49"N 96°22'38.74"W), 23 July 2003. ECOSUR P0000 (10), La Mancha, Veracruz (19°35'13.64"N 96°22'39.24"W), 23 September 2004. ECOSUR P0000 (10), La Mancha, Veracruz (19°35'10.16"N 96°22'38.88"W), 27 April 2004. ECOSUR P0000 (1), La Mancha, Veracruz (19°35'7.49"N 96°22'38.74"W), 27 May 2004. ECOSUR P0000 (1), La Mancha, Veracruz (19°36'4.91"N 96°22'18.09"W), 28 September 2003. ECOSUR P0000 (6), Montepío, Veracruz (18°38'46.97"N 95°5'44.19" W), basaltic rocky shore, 1 April 2004.

Type locality. La Mancha Lagoon, Veracruz, Mexico, Gulf of Mexico.

Etymology. The specific name stems from the Latin adjective brunneus, -a, -um, meaning brown, referring to the brown pigmentation present in this species.

Description. Holotype complete, 24 mm long, 2.1 mm wide, 72 chaetigers. Body tapered, pigmented. Dark brown pigment on prostomium and tentacular ring, more intense in anterior end, discoloring posteriorly; two rectangles arising from base of antennae, becoming wider toward posterior end of prostomium, surrounding eyes, with a medial pale line between them; inner margin of palpophores and some yellowish spots on palpostyles (Fig. 30A). Prostomium longer than wide, pigmented; antennae cirriform, 3 times longer than wide, not passing the palps; eyes black, lenticulated, in trapezoidal arrangement, diameter as long as base of antennae (Fig. 30A), anterior pair semicircular, posterior pair oval (Fig. A). Tentacular ring as

171 long as first chaetiger; tentacular cirri with cirrophore conspicuous, longest tentacular cirri reach chaetiger 4 (Fig. 30A). Pharynx dissected, jaws brown with 11 teeth. Maxillary ring with pectinate bars (II, III and IV), smooth bars (IV) and conical paragnaths (I and IV); I= 2 in vertical line, II= 3/3 lines in triangle, III= 4 lines in triangle, IV= 5/5 pectinate lines and additional cones and bars, in sigmoid. Oral ring with conical (VII-VIII) and P-bar (VI and VII-VIII) paragnaths; V= 1, VI= 1-1, VII-VIII= 20 in a row, alternating, forming a jagged line (Figs 30I, J). Parapodial cirri pattern: Dorsal cirri longer than notopodial ligules in anterior region, becoming shorter toward posterior end, basally inserted to dorsal ligules in most anterior segments, displaced medially in medial segments, subdistal in posterior segments and distal in posterior ones, extending beyond notopodial ligules throughout body. Ventral cirri subequal to neuropodial ligules throughout body, basally inserted to neuropodial ligules throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 30C), dorsal cirri basal, 1.5 times longer than dorsal ligules. Dorsal ligules subconical, blunt tip, slightly longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, as long as neuropodial ligules; neuropodial ligules digitiform. Ventral cirri basal, subequal than neuropodial ligules. In anterior and middle chaetigers (Figs 30D, E), dorsal cirri medial, 1.5 times longer than notopodial ligules. Notopodial ligules subconical, blunt tip, slightly longer than notoacicular lobes; notoacicular lobes rounded to digitiform, subequal to neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, twice longer than neuropodial ligules; postchaetal lobes rounded, as long as neuroacicular ones; neuropodial ligules digitiform. Ventral cirri basal, as long as neuropodial ligules. In posterior chaetigers (Fig. 30F), dorsal cirri subdistal, 0.8 length of notopodial ligules. Notopodial ligules pennant-like, twice longer than notoacicular lobes; notoacicular lobes digitiform, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, blunt tip, slightly longer than neuropodial ligules; postchaetal lobes rounded, as long as neuroacicular ones; neuropodial ligules digitiform. Ventral cirri basal, as long as neuropodial ligules. In posterior-most chaetigers (Fig. 30G), dorsal cirri distal, 0.4 length of notopodial ligules. Notopodial ligules foliose, pennant-like, 3 to 4 times longer than notoacicular lobes; notoacicular lobes digitiform, twice longer than neuroacicular lobes. Neuroacicular lobes subconical, slightly

172 longer than neuropodial ligules; postchaetal lobes rounded, as long as neuroacicular ones; neuropodial ligules digitiform. Ventral cirri basal, slightly shorter than neuropodial ligules. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spiniger and falcigers in sub-acicular fascicles. Notopodial homogomph spiniger pectinate, fine teeth, decreasing in size toward distal end (Fig. 30N). Neuropodial homogomph spinigers as notopodial ones; neuropodial heterogomph spinigers pectinate, teeth basally coarse, decreasing in size toward distal end (Fig. 30M, O). Neuropodial heterogomph falcigers falcate, pectinate, fine teeth, half to two thirds of inner edge edentate, distal tooth stout; supra-acicular falcigers stouter than sub-acicular ones (Figs 30L–N). Pygidium crenulated; anal cirri cirriform, as long as last six segments (Fig. 30B).

Remarks. Specimens identified as P. gallapagensis by de León-González (1997) from Veracruz and Tamaulipas, Gulf of Mexico, and Dueñas-Ramírez & Quiros-Rodríguez (2012) from Córdoba, Colombian Caribbean, resemble to P. brunnea sp. n, but both records require confirmation with the revision of their specimens. The description of P. gallapagensis by Gravier (1909) shows similarities with the provided here, but some differences can be noted. In P. brunnea sp. n., the antennae are larger than in P. gallapagensis. Furthermore, in P. brunnea sp. n., the tentacular cirri reach chaetiger 4, while in P. gallapagensis they reach chaetiger 1. Moreover, in P. brunnea sp. n., notopodial ligules in posterior chaetigers are less developed than in P. gallapagensis.

Distribution. Gulf of Mexico.

Pseudonereis ferox (Hansen, 1882)

Nereis ferox Hansen 1882: 14, Pl. 4, Figs 34–39. Pseudonereis variegata Bakken 2007: 168–171, Figs 14A–C, 15A–H (non Kinberg, 1865)

Type locality. Rio de Janeiro, Brazil.

173

Remarks. Bakken (2007) redescribed this species, but synonymized it with P. variegata. However, there are several differences between them, and here P. ferox is considered as a distinct species. Further comments in remarks section about synonymy with P. variegata below. Fauvel (1914) identified under this name some specimens from São João dos Angolares, São Tomé, Gulf of Guinea. He recognized the similarity between P. ferox and P. gallapagensis, but based upon Gravier (1909), he retained the former as a valid species. However, specimens from the Gulf of Guinea have some differences as well, as noted by Fauvel; additionally, the large dorsal cirri and tentacular cirri make them closer to P. citrina sp. n. than to P. ferox, but the relative length of notopodial ligule and dorsal cirri-notopodial ligule ratio in posterior-most chaetigers, lack of paragnaths on Area V and shape of paragnaths on Area VI allow its separation of the Gulf of Guinea specimens from those from the Caribbean Sea.

Questionable records

Pseudonereis gallapagensis Kinberg, 1865

Pseudonereis gallapagensis Kinberg 1866: 174. Gravier 1909: 629–633, Pl, 16, Figs 15–20. Hartman 1948: 68–69. Bakken 2007: 157–159 (partim).

Type locality. Indifatigable Island (Santa Cruz Island), Galapagos.

Remarks. This species has been reported from several localities, requiring revision. Currently the species include P. formosa Kinberg, 1865 and Neanthes variegata Kinberg, 1865 from Hawaii after revision of type materials (Hartman 1948, Bakken 2007), but Bakken noted the poor conditions of specimens. P. gallapagensis and P. formosa deserves a new revision, including topotypic specimens in good conditions to ensure their conspecificity. P. gallapagensis is related to another problematic species, P. variegata, see further comments in remarks section of P. variegata below.

Pseudonereis variegata (Grube & Kröyer in Grube, 1858)

174

Nereis (Nereilepas) variegata Grube 1858: 164–165. Pseudonereis variegata Monro 1933: 45.

Type locality. Valparaiso, Chile, Pacific Ocean.

Remarks. Bakken (2007), following Ehlers (1901) and Augener (1934), redescribed this species with the syntype of Nereis ferox Hansen, 1882 from Rio de Janeiro, Brazil, because the type material of N. variegata Grube & Kröyer in Grube was not available. This decision was unfortunate because of the discrepancies of type localities, and because no topotype material was studied, and some publications that characterize the species from localities in South American Pacific coasts were overlooked. Salazar-Vallejo & Eibye-Jacobsen (2012) found the holotype in the Zoological Museum, University of Copenhagen (ZMUC-POL-1496).

Rozbaczylo & Bolados (1980) made a useful redescription of P. variegata with specimens collected near its type locality. Some relevant differences include that the area sVII-VIII of P. ferox Hansen have one jagged row with 18–21 paragnaths, whereas in P. variegata there are two jagged rows with 39–41 paragnaths, both species have alternating conical and P-bar types. Bakken (2007) concluded that this feature was not useful for separating P. variegata from P. gallapagensis as other authors have noted, especially because P. ferox and the P. gallapagensis have a similar pattern, but not P. variegata. Ranges are overlapped in other pharyngeal areas. In P. gallapagensis and P. ferox, dorsal cirri are terminally inserted in posterior chaetigers; however, in P. variegata they are slender and sub-terminally inserted in posterior chaetigers. Bakken (2007) noted this feature as useful to separate it from P. gallapagensis. Further, notopodial ligules from posterior chaetigers are quite broad and 4–5 times longer than notoacicular ligules in P. ferox; whereas in P. variegata they are narrower and 2–3 times longer than notoacicular lobes. Finally, heterogomph spinigers start from chaetiger 40 in P. ferox and continue throughout body in P. variegata. Even if heterogomph spinigers are overlooked or missed in the type material, parapodial and pharyngeal features provide sufficient features for separating them, therefore in this work P. ferox is considered as distinct species.

175

Subfamily Namanereidinae Hartman, 1959

Remarks. The subfamily Namanereidinae has species living in brackish, freshwater or subterranean environments and are characterized by having reduced morphologic features that difficult the separation of species (Glasby 1999); resulting in that some are regarded as widespread species or species complexes (Glasby et al. 2009). Due to their life habits, what are currently regarded as populations belonging to a single species could be largely isolated in some ways, and this generates genetic differentiation, although they could be morphologically indistinguishable. For example, Magesh et al. (2014), using molecular and morphological data, showed incipient speciation in some Namalycastis from India.

Namalycastis Hartman, 1959

Type species. Lycastis abiuma Müller & Grube in Grube, 1872, by original designation (Hartman 1959a).

Diagnosis (modified from Conde-Vela 2013). Prostomium with anterior margin entire or incised. Eyes generally present. Four pairs of tentacular cirri Maxillary ring bare. Oral ring bare, occasionally papillae on areas VI. All chaetigers type E. Dorsal cirri becoming larger and foliose toward posterior chaetigers. Neuroacicular lobe occasionally distally bilobed. Notochaetae generally present, homogomph spinigers; neurochaetae arranged in discrete fascicles, homogomph and heterogomph spinigers and heterogomph falcigers, ocassionally uppermost sub- acicular heterogomph falcigers with a graded series of lengths, or replaced by spinigers toward posterior chaetigers.

Remarks. The genus is easily recognizable by the distinct dorsal cirri in anterior and posterior chaetigers. Generally, dorsal cirri become foliose but in some species, this feature is attenuated. The genus has been revise by Glasby (1999) and few new species have been added, including one from the Grand Caribbean region (Conde-Vela 2013).

176

Namalycastis abiuma (Müller & Grube in Grube, 1872)

Lycastis abiuma Grube 1872: 47–49. Namalycastis abiuma Glasby 1999: 31, Figs 10a–h.

Type locality. Desterro (Santa Catarina Island), Brazil (27°35’S 48°31’W).

Remarks. Glasby (1999) redescribed the species based upon holotype, and proposed N. abiuma species group to include specimens closely related to it but from several localities, giving it a cosmopolitan status. Currently, a new species has been described from the species group (Glasby et al. 2007), other studies have shown incipient speciation in the group (Magesh et al. 2014).

Namalycastis arista Glasby, 1999

Namalycastis arista Glasby 1999: 35–37, Figs 11a–g.

Type locality. Paradise Beach, Guyana (15°34’N 57°57’W), in tidal zone.

Remarks. The species is recognizable by the presence of a mucron in distal end of blades of falcigers (Glasby 1999). Currently, the distribution of this species is restricted to the type locality.

Namalycastis borealis Glasby, 1999 Figure 31J–M

Namalycastis borealis Glasby 1999: 37–39, Figs 1c, 12a–g, 13 (partim). Conde-Vela 2013: 482– 485, Figs 3A–J.

Type locality. Beaufort, North Carolina (34°43’N 76°44’W), intertidal.

177

Type material. Northwestern Atlantic Ocean, United States. Holotype USNM 52926, Piver´s Island, Beaufort, North Carolina, 13 April 1973, Coll. C. Jenner, S.L. Gardiner. Paratypes USNM 57129 (2), East Goose Creek, Florida, 17 April 1966, Coll. E.V. Komarek.

Description. Type material in good condition, specimens very large. Holotype 70 mm long, 2 mm wide, 129 chaetigers. Pigmentation faded out in anterior and middle body, brown pigment toward posterior end. Prostomium wider than long, anteriorly incised, groove present; antennae subconical, half as long as prostomium, slightly passing palps; eyes rounded, anterior pair slightly larger than posterior one, obliquely aligned (Fig. 31J). Tentacular ring half as long as first chaetiger; four pairs of tentacular cirri, longest tentacular cirri reach chaetiger 2 (Fig. 31J). Pharynx smooth, without structures. Paratype with jaws brown, 9 teeth, with single terminal tooth (Fig. 31M). One parapodium from middle chaetiger has an abnormal dorsal cirrus (Fig. 31L). Posterior end with small chaetigers, possibly regenerating. Pygidium crenulated, anal cirri short.

Variation. The dorsal cirri in posterior chaetigers have different states of contraction, modifying their shape. Some dorsal cirri are as illustrated by Glasby (1999) while others are as those shown by Conde-Vela (2013). Other features were described by Glasby (1999).

Remarks. Glasby (1999) included in this species, material collected and studied by Rasmussen from Sapelo Island, Georgia. However, some features described by Rasmussen based upon living specimens do not match with material from North Carolina. As Glasby noted, living specimens from Georgia are translucent, such that even the jaws can be seen by transparency. Also, rate of length of dorsal cirri and neuroacicular lobe in posterior chaetigers ranges from 2 to 6.9. Finally, Rasmussen described the spiniger with a hair-like tip (i.e. slender and long). One specimen (UANL-3926) from Nautla, Mexico, Gulf of Mexico, does not have pigmentation, the rate of length of dorsal cirri and the neuroacicular lobe is near 7 and spinigers have this type of tip, and matches well with the description of Rasmussen, but not with the type material from North Carolina. Therefore, specimens from Georgia belong to a different species but they must be studied and compared with type material to clarify these doubts.

178

Distribution. Virginia to Florida, United States; Mexican Caribbean.

Namalycastis geayi (Gravier, 1901)

Lycastis geayi Gravier 1901: 399–402; 1902: 361–366, Figs 11–17. Namalycastis geayi Glasby 1999: 47–49, Figs 18a–g.

Type locality. Ouanary River, French Guiana.

Remarks. The species is remarkable because it only has spinigers and the dorsal cirri are not enlarged as usual for this genus (Glasby 1999). Currently restricted to type locality.

Distribution. Frech Guiana.

Namalycastis intermedia Glasby, 1999 Figure 31A–I

Namalycastis intermedia Glasby 1999: 55–57, Figs 1c, 19, 22a–i.

Type material. United States. Holotype USNM 178870; N Rockefeller Wildlife Refuge, Cameron Parish, Louisiana, April 1963, Coll. Whitehead. Paratypes USNM 31011 (15), data as in holotype.

Type locality. Cameron Parish, Louisiana.

Description. Type material in regular condition, most incomplete. Specimens pale, without pigmentation. Holotype incomplete, posterior end damaged, 14 mm long, 0.8 mm wide, 77 chaetigers (Fig. 31A). Prostomium slightly wider than long, anterior margin incised, groove present; antennae cirriform, half as long as prostomium; eyes with dispersed pigment, obliquely aligned (Fig. 31B).

179

Tentacular ring as long as first chaetiger; four pairs of tentacular cirri, longest one reach chaetiger 5 (Fig. 31B). Pharynx not everted. Jaws light brown, 8 teeth, with single terminal tooth (Fig. 31C). Parapodial cirri pattern: Dorsal cirri longer than neuroacicular lobe and basally inserted throughout body, becoming foliose toward posterior end. Ventral cirri shorter than neuroacicular lobe and basally inserted throughout body. Chaetiger type E throughout body. In anterior chaetigers (Fig. 31G), dorsal cirri cirriform, slightly longer than neuroacicular lobe. Neuroacicular lobe subconical, 2.3 times wider than base of dorsal cirri. Ventral cirri 1/5 as long as dorsal cirri. In middle chaetigers (Fig. 31H), dorsal cirri 2.5 times longer than neuroacicular lobe. Neuroacicular lobe subconical, blunt tip, half as wide as dorsal cirri bases. Ventral cirri 1/5 as long as dorsal cirri. In posterior chaetigers (Fig. 31I) dorsal cirri 1.5 times longer than neuroacicular lobes, foliose feature poorly developed. Neuroacicular lobe subconical, blunt tip, as wide as base of dorsal cirri. Ventral cirri 1/5 as long as dorsal cirri. Supra-acicular spiniger (Fig. 31F) pectinate, teeth basally coarse, decreasing in size distally. Supra-acicular falciger (Fig. 31E) pectinate, teeth basally coarse, decreasing in size distally. Sub- acicular falciger (Fig. 31D) pectinate, teeth basally coarse, decreasing size distally, blade larger than in supra-acicular ones.

Remarks. No discrepancies in description provided by Glasby (1999) were found. Currently restricted to Gulf of Mexico. The foliose aspect of dorsal cirri from posterior chaetigers is poorly developed, as in N. occulta Conde-Vela, 2013.

Distribution. Gulf of Mexico.

Namalycastis kartaboensis (Treadwell, 1926)

Namonereis (sic) kartaboensis Treadwell 1926: 101–103, Figs 33A–C. Namalycastis kartaboensis Glasby 1999: 57–59, Figs 23a–h.

180

Type locality. Kartabu Point (6°23'N 58°41'W), Guyana.

Remarks. The holotype is in very poor conditions (Glasby 1999); the species was redescribed with non-type specimens from Surinam and French Guiana by Glasby (1999). The species is easy recognize by having neuropodial heterogomph falcigers with minute teeth, almost inconspicuous and givig them an edentate appearance, and sub-acicular spinigers have very coarse basal teeth.

Distribution. Guyana.

Namalycastis macroplatis Glasby, 1999

Namalycastis macroplatis Glasby 1999: 61–63, Figs 24a–g. Santos & Lana 2001: 139-141, Figs 7-14.

Type locality. Marajos Island, Brazil (01°00’S 49°00’W), in beach.

Remarks. The species has been recorded from Brazilian coasts only. The presence of neuropodial heterogomph with long blades (several times longer than wide) is the main feature of this species, other diagnostic feature are the very long and narrow dorsal cirri in posterior chaetigers.

Distribution. Northern Brazil.

Namalycastis occulta Conde–Vela, 2013

Namalycastis occulta Conde-Vela 2013: 478–482, Figs 1A–K, 2A–C.

Type locality. Chetumal Bay, Mexico, intertidal.

181

Type material. Caribbean Sea, Mexico. Holotype ECOSUR 0151, Bahía Chetumal, Quintana Roo, México, 7 Jul. 2011, intertidal, under rocks, in sediment with leaves litter of red mangrove (Rhizophora mangle) (18°30'30.82" N, 88°16'43.61" W), col. V.M. Conde Vela, A.E. Te Gómez. Additional material. Caribbean Sea, Mexico. ECOSUR P0000 (4), Panto-Ha, Laguna Bacalar, Quintana Roo (18°34’15.62’’N 88°26’39.9’’W), 20 May 2015, Coll. L.F. Carrera-Parra, J. Jarquín-González. ECOSUR P0000(3), Raudales, Laguna Guerrero, Quintana Roo (18°42’24.12’’N 88°15’18.35’’W), 28 May 2015, on rocks with Batophora, 0.5 m depth, Coll. L.F. Carrera-Parra, J. Jarquín-González.

Remarks. The material examined extends the distribution of this species into freshwater environments, instead of brackish-water environments of Chetumal Bay, its type locality (Conde- Vela 2013). Also, they were found on seagrass and on rocks with the green algae Batophora sp. After the revision of the holotype and one non-type specimen, both with the pharynx everted, two massive structures were found on area VI, as in some specimens of N. hummelincki. The presence of these structures was highlighted in the diagnosis for the genus.

Distribution. Mexican Caribbean.

Questionable records

Namalycastis senegalensis (de Saint-Joseph, 1901)

Lycastis senegalensis de Saint-Joseph 1901: 217–224, Pl. 1, Figs 1–7. Namalycastis senegalensis Glasby 1999: 65–70.

Type locality. Marsassoum, Songrougou Stream, Senegal.

Remarks. Gravier (1901) described L. ouanaryensis and noted differences with L. senegalensis. Later, Glasby (1999) revised the type material of L. ouanaryensis from French Guiana, recognizing two species mixed in it: some were identified as a new species, N. macroplatis, while the remaining ones were transferred to N. senegalensis. The author considered the differences

182 highlighted by Gravier as intraspecific variations and synonymyzed them. However, this synonymy is questionable because of the enormous distance between type localities.

Namanereis Chamberlin, 1919

Type species. Lycastis quadraticeps Blanchard in Gay, by original designation (Chamberlin, 1919).

Diagnosis (after Glasby 1999). Prostomium with anterior margin entire or incised. Eyes generally present. Four or three pairs of tentacular cirri. Maxillary ring bare. Oral ring bare, occasionally papillae on areas VI. All chaetigers type E. Dorsal cirri with similar development throughout body. Neuroacicular lobe occasionally bilobed distally. Notochaetae generally absent, homogomph spinigers if present; neurochaetae arranged in discrete fascicles, homogomph and heterogomph spinigers and heterogomph falcigers, ocassionally uppermost sub-acicular heterogomph falcigers with a graded series of lengths or replaced by spinigers toward posterior chaetigers.

Remarks. After the revision of the subfamily by Glasby (1999), some new species have been described. The species from the Caribbean Sea were reviewed by Williams (2004). Namanereis species are easily recognized to Namalycastis in having dorsal cirri with similar size in all chaetigers.

Namanereis beumeri (Augener, 1922) inf. reinst.

Lycastopsis beumeri Augener 1922: 42.1936: 346. Wesenberg-Lund 1958: 14–17, Figs 9-10 (partim).

Type locality. Habana, Cuba.

183

Remarks. Augener (1922) erected the genus Lycastopsis based upon a new species, N. beumeri, by the presence of three pairs of tentacular cirri instead of four. Wesenberg-Lund (1958) considered as valid the species, but synonymyzed L. pontica with it, but Glasby (1999) regarded the latter as valid. Also, Wesenberg-Lund recognized the cushion-shaped papillae on areas VI in this species. Few differences can be noted after the current descriptions. Wesenberg-Lund (1958) described the jaws having one or two teeth in cutting edge, matching with bifid-type jaws, without teeth in inner edge, while lectotype of N. littoralis have five teeth (Glasby 1999). In addition, dorsal cirri of N. beumeri were reported as triangular, while in N. littoralis they are digitiform. Glasby (1999) examined the type material and referred (as others) it to N. littoralis (Müller & Grube in Grube, 1972) from Santa Catarina Island, Brazil; however, the author did not include more details about his decision, and unfortunately, the original description by Augener (1922) is very scarce. The species is considered as valid here, requiring a detailed revision of the N. littoralis species group.

Distribution. Caribbean Sea.

Namanereis hummelincki (Augener, 1933)

Lycastopsis hummelincki Augener 1933: 352–355, Figs 1a–d. Wesenberg-Lund 1958: 12-14, Figs 5–8. Namanereis hummelincki Hartman 1959: 163. Glasby 1999: 86-89, Figs 37a–g.

Type locality. Fontein, Bonaire (12°15’N 68°27’W).

Material examined. Caribbean Sea, Netherlands Antilles. Topotypes USNM 29715 (3); Fontein Bron, Bonaire, 30 March 1937. USNM 29716 (1), Brondi Pos Bronswinkle, Bonaire, 28 August 1955. USNM 29713 (3), Bron Di Rooi Sanchez, Curacao, 11 November 1937. USNM 29714 (12), Bron Cajoeda, Curacao, 26 September 1948.

184

Description. Complete and incomplete specimens, in diverse condition. Most specimens pale, without pigmentation, some light brown (USNM 29714). Topotypes one juvenile, another one broken in two fragments, and one portion without anterior end. Specimen broken complete 10 mm long, 0.5 wide, 64 chaetigers. Prostomium 1.5 to 2.0 wider than long, margin entire or very slightly incised, groove present; antennae cirriform, 4/5 as long as prostomium; eyes absent. Tentacular ring as long as first chaetiger; three pairs of tentacular cirri, longest one reaches chaetiger 4. Pharynx everted in some specimens (USNM 29713, 29716). Jaws dark brown, without teeth on inner edge, distal tooth bifid. Oral ring with two massive structures, papillae-like in area VI, remaining areas bare. Parapodial cirri pattern: Dorsal cirri subequal or slightly shorter than neuroacicular lobes in anterior and middle body, becoming slightly longer toward posterior end, basally inserted throughout body. Ventral cirri shorter tan neuropodial ligules and basally inserted throughout body. Chaetigers type E throughout body. Dorsal cirri as long as neuroacicular lobes in anterior chaetigers, slightly shorter in middle ones, and slightly longer in posterior ones. Neuroacicular lobe subconical, decreasing their width posteriorly. Ventral cirri bases diminishing their length toward posterior chaetigers. Chaetae lost in most chaetigers. Pygidium tripartite; anal cirri cirriform, as long as last two segments.

Variation. In some specimens, the dorsal cirri from posterior chaetigers are two times longer than neuroacicular lobes.

Remarks. Wesenberg-Lund (1958) noted two massive structures on areas VI, also found in specimens examined, not being juvenile specimens, and in Namalycastis occulta. Also, Glasby (1999) found papillae in some juveniles of this species from Curaçao and in juveniles of N. amboinensis (Pflugfelder, 1933). As Glasby (1999) noted, because most specimens did not expose the pharynx, their dissection is difficult due to their very small size. This is an important feature because the definition of this subfamily includes species without pharyngeal structures. A similar feature is present in

185

Ceratonereis, having papillae on area VI, but they are currently disregarded in phylogenetic studies. The pharyngeal armature is relevant for phylogenetic analysis, and this feature must be considered in future studies.

Namanereis minuta Glasby, 1999

Namanereis minuta Glasby 1999: 96–98, Figs 41a–f.

Type locality. Nallee Grand’ Anse, Haiti, Hispaniola (18°35’55’’N 74°10’46’’W).

Remarks. The species belongs to a group of species characterized by having bifid jaws with edentate cutting edges, lacking eyes and having heterogomph falcigers with relatively long blades that are replaced by heterogomph spinigers toward posterior chaetigers (Conde-Vela 2013). More studies are required to elucidate their phylogenetic affinities among them and with other namanereidins. Currently, N. minuta is only known from la Hispaniola (Glasby 1999).

Namanereis serratis Glasby, 1999

Namanereis serratis Glasby 1999: 106–108, Figs 47a–f.

Type locality. Étang Saumâtre (Lake Azuei), Haiti, Hispaniola.

Remarks. The species is very close to three other species, but it can be recognized by chaetal features (Glasby 1999), also the blades of heterogomph falcigers are similar to those found in N. occulta. Currently only known from la Hispaniola (Glasby 1999).

Namanereis stocki Glasby, 1999

Namanereis stocki Glasby 1999: 108–110, Figs 48a–g.

Type locality. St. Ann´s Great River, Jamaica (18°25’44’’N 77°06’32’’W).

186

Remarks. The species is closely related to N. hummelincki; however, they differ in the number of tentacular cirri, a feature not related with the size of specimens (Glasby 1999). Currently known from Jamaica and Cuba (Glasby 1999).

Namanereis sublittoralis Glasby, 1999

Namanereis sublittoralis Glasby 1999: 110–112, Figs 49a–f.

Type locality. St. Eustatius, Lesser Antilles (17°33’N 63°00’W).

Remarks. The species resembles N. littoralis, differing only in the relative length of falcigers, the high number of teeth and in the articulation of spinigers; currently known from type locality (Glasby 1999).

Namanereis christopheri n. sp. Figure 32A–O

Namanereis cavernicola Glasby 1999: 83–86, Figs 1c, 8c, 34, 35a–g (partim)

Type locality. St. Vincent, Lesser Antilles.

Etymology. The specific name is after Christopher J. Glasby, as a means for recognizing his numerous contributions in polychaete taxonomy, especially in namanereidins, and because he identified this species as new in first instance, and separated it from other type material, but changed his mind later.

Type material. Holotype LACM-AHF 0000 and paratypes LACM-AHF 0000 (1) and LACM- AHF 0000 (10), Golden Grove, near Chateaubelair Bay, Saint Vincent, Saint Vincent and the Grenadines (13°15’N 61°11’W), 31 July 1972, 435 m above sea level, spring pool in Colocasia swamp, Coll. J.J. Rankin.

187

Description. Holotype complete, 32 mm long, 1 mm wide, 95 chaetigers; body with several incisions along middle chaetigers, otherwise in good condition. Paratypes complete, in good condition. Body pale, without pigmentation (Figs 32A, B). Prostomium wider than long, anterior margin entire, groove present; antennae cirriform, as long as prostomium; eyes absent (Fig. 32A). Tentacular ring as long as first chaetiger; three pairs of tentacular cirri, longest one reach chaetiger 3 (Fig. 32A). Pharynx dissected; jaws edentulate, two distal teeth, with bifid appearance (Fig. 32E). Pharynx smooth, with a cushion-shaped papillae on area VI (Fig. 32D). Parapodial cirri pattern: Dorsal cirri longer than neuroacicular lobes and basally inserted throughout body. Ventral cirri shorter tan neuropodial ligules and basally inserted throughout body. Chaetigers type E throughout body. In anterior chaetigers (Fig. 32F), dorsal cirri 3 times longer than neuroacicular lobes; neuroacicular lobes digitiform to rounded, as long as ventral cirri; ventral cirri half as long as dorsal ones. In middle chaetigers (Figs 32G, H), dorsal cirri twice longer than neuroacicular lobes; neuroacicular lobes digitiform to rounded, twice longer than ventral cirri; ventral cirri one-third to one-half as long as dorsal ones. In posterior chaetigers (Fig. 32I), dorsal cirri twice longer than neuroacicular lobes; neuroacicular lobes digitiform to rounded, 3 times longer than ventral cirri; ventral cirri half as long as dorsal ones. Notochaetae absent. Neurochaetae in type D arrangement, i.e. supra-acicular chaetae heterogomph falcigers and sesquigomph spinigers in pre- and post-acicular fascicles respectively; sub-acicular chaetae heterogomph falcigers with short and long blades in pre-acicular fascicles. Supra-acicular sesquigomph spiniger pectinate, fine teeth, decreasing in size toward distal end (Figs 32J, N); heterogomph falcigers pectinate, coarse teeth, decreasing in length toward distal end (Fig. 32L). Sub-acicular falcigers pectinate, fine teeth, falcate tip, decreasing in size toward distal end; upper heterogomph falcigers long bladed, blades 2–3? times longer than lower ones, lower falcigers stouter than upper ones (Figs 32K–M, O).

Pygidium tripartite; anal cirri cirriform, short, as long as the last chaetiger (Fig. 32B).

188

Variation. Specimens vary little in all features, only size of body varies among specimens.

Remarks. The material of this species was previously examined by Glasby (1999) and identified as N. cavernicola. The three vials from the Los Angeles Museum include two labels. One label has the name “Lycastopsis F”, identified in 1988, and includes the designation of holotypes and paratypes. The other label has other name, “Lycastopsis pectinalis n. sp.”, identified in 1990. Therefore, he thought it was a new species, but later he changed of mind and included this material in N. cavernicola. The result was that in description of. N. cavernicola there are mixed features and high ranges of variation. The species has some differences with N. cavernicola that ensure their separation. First, N. christopheri sp. n. has an anterior margin of prostomium entire and antennae are shorter than prostomium, while in N. cavernicola the anterior margin is incised and antennae are longer than prostomium. In addition, in N. christopheri the jaws are much broader than in N. cavernicola. The incised anterior margin of prostomium and other features were compared with L. alticola but not with N. cavernicola, arguing that the only difference among Caribbean and Pacific materials were the longer dorsal cirri in the former ones. The recurved tip in N. cavernicola is regarded here as specific feature. Glasby (1999) noted, after revision of several parapodia of N. cavernicola, diverse grades of curvature in notoaciculae and even in neuroaciculae, absent in N. christopheri sp. n. This and other features as cleft prostomium and articulated antennae and tentacular cirri was used to establish Lycastilla, but after Glasby, they are present in other species and regarding them as specific features is incorrect. All specimens were whole-mounted and examined under microscope, no spinigers were found in sub-acicular fascicles. The species does not replace upper long bladed falcigers by spinigers toward posterior chaetigers, as in other species with type D chaetal arrangement; however, it was regarded as variant of this type.

Questionable records

Namanereis amboinensis (Pflugfelder, 1933)

Lycastopsis amboinensis Pflugfelder 1933: 69–71, Figs 9–11.

189

Namanereis amboinensis Hartman 1959a: 163. Glasby 1999: 76–79, Figs 31a–g (partim).

Type locality. Ambon Island, Moluccas, Indonesia.

Remarks. Glasby (1999) included specimens from several localities in this species, regarding it as circumtropical and circumsubtropical. The synonymy with L. tecolutlensis Rioja, 1946 and the inclusion of material from the Caribbean Sea extended the distribution to the Grand Caribbean. Here this decision is regarded as questionable due to the enormous distance among localities (Glasby et al. 2009), requiring a detailed revision to restrictof the distribution of regional species.

Namanereis cavernicola (Solís-Weiss & Espinasa, 1972)

Lycastilla cavernicola Solís-Weiss & Espinasa 1991: 632–635, Figs 1a–e, 2a–f. Namanereis cavernicola Glasby 1999: 83–86, Figs 35a–g (partim).

Type locality. Isote Cavern, Guerrero, Mexico, in cavern at 1650 m above sea level.

Remarks. Glasby (1999) redescribed the species with paratypes and non-type materials from Caribbean Sea. However, parts of the materials from Caribbean Sea were regarded as a distinct species, N. christopheri sp. n. (see above).

Namanereis pontica (Bobretzky, 1872)

Lycastis pontica Bobretzky 1872: 1–3, Pl. 14, Figs 1–4. Namanereis pontica Glasby 1999: 98–100, Figs 43a–e.

Type locality. Sevastopol Bay, Black Sea.

Remarks. The name was introduced by Perkins & Savage (1975) following Hartman (1959a), who considered probable the inclusion of several names under N. quadraticeps, including L. pontica. On the contrary, Pettibone (1963) included several names under N. pontica, including L.

190 tecolutlensis and L. beumeri as synonyms, and her material came from Massachusetts. Later, Hartmann-Schröder (1980) reported N. pontica with specimens from the Caribbean Sea. Glasby (1999) referred these materials and others to N. littoralis species group; however, the resolution is inconclusive, requiring further revision of materials.

Namanereis quadraticeps (Blanchard in Gay, 1849)

Lycastis quadraticeps Blanchard 1849: 24–25, Fig. 7. Namanereis quadraticeps Hartman 1959a: 162. Hutchings & Glasby 1985: 106–108, Figs 3a–e. Glasby 1999: 100-103, Figs 44a–g.

Type locality. North of San Gregorio Bay, Chile, Strait of Magellanes (52°34’S 70°10’W)

Remarks. Glasby (1999) found indistinguishable specimens from several localities (Chile, Australia, New Zealand, South Africa, Namibia), and regarded them as forming a species group, instead of following his alternative scheme used for N. littoralis and N. abiuma. As the resolution is inconclusive, and because more than one species are involved (Glasby et al. 2009), the species deserves further revision.

Subfamily Gymnonereidinae Banse, 1977

Ceratocephale Malmgren, 1867

Type species. Ceratocephale loveni Malmgren, 1867, by original designation.

Diagnosis (after Hutchings & Reid 1990). Prostomium with anterior margin incised. Eyes present. Four pairs of tentacular cirri. Maxillary ring bare. Oral ring with papillae only in areas V–VIII. First two chaetigers type B. Notopodial ligules absent. Notopodial postchaetal lobes present at least on anterior chaetigers; neuropodial postchaetal lobe present, at least in anterior

191 and middle chaetigers. Notochaetae and neurochaetae homogomph and sesquigomph spinigers and occasionally sesquigomph falcigers, arranged in dense bundles around neuroaciculae.

Remarks. Chamberlin (1919) proposed Chaunorhynchus for replacing Ceratocephale or its alternate common form Ceratocephala, the latter preoccupied for a trilobite genus, and Monro (1933: 46) followed the change. Hartman (1959b: 237) considered Ceratocephala as the correct spelling and Ceratocephale as the variant, overlooking the commentary of their homonymy by Chamberlin. However, the original name is Ceratocephale (Malmgren 1867: 60) and Ceratocephala is an incorrect subsequent spelling, therefore an invalid, not substitute name (ICZN 1999, Art. 33.3). If the argument of Chamberlin about the ‘common’ use of Ceratocephala was true, it could be preserved (ICZN 1999, Art. 33.3.1), but the most common name is the original spelling, therefore Chaunorhynchus becomes in a synonym of Ceratocephale. A remarkable feature of this genus is the presence of densely arranged homogomph spinigers, gold-colored (C. aureola and C. setosa by Hutchings & Reid, 1990 were named after these features), in circular arrangement (Fig. 33C, F).

Ceratocephale oculata Banse, 1977 Figure 33A–I

Ceratocephale oculata Banse 1977: 620–621, Figs 3a–f.

Type locality. Beaufort, North Carolina, 35 m depth.

Type material. Northwestern Atlantic Ocean, United States. Holotype UNSM 51079, off Beaufort, North Carolina (34°24'N 76°05'W), 35 m depth, 6 April 1965, Coll. J.H. Day. Paratypes USNM 52096 (4), same data as holotype.

Additional material. Gulf of Mexico, United States. USNM 1183121 (1), Off Port Isabel, Texas, 27 September to 7 October 1977, Coll. MSI Port Aransas.

192

Description. Holotype incomplete posteriorly, in regular conditions, 1.3 mm long, 0.4 mm wide, 47 chaetigers (Fig 33A). Paratypes (USNM 52096) in several fragments, incomplete posteriorly, in regular condition; best preserved one incomplete, 3 mm long, 0.4 mm wide, 19 chaetigers (Fig. 33B). Prostomium as long as wide, anterior margin deeply incised, amlost reaching anterior pair of eyes; antennae cirriform, as long as prostomium; eyes subequal, reddish, slightly faded, in trapezoidal arrangement (Fig. 33D). Tentacular ring as long as first chaetiger; four pairs of tentacular cirri, longest reach chaetiger 2 (Fig. 33D). Parapodial cirri pattern: dorsal cirri as long as neuroacicular lobes in anterior chaetigers, becoming longer than neuroacicular lobes toward posterior end. Ventral cirri shorter than neuroacicular lobes and basally inserted throughout body; cirri duplicated on each parapodium, of different length. First two chaetigers type B, remaining biramous. In first two chaetigers (Fig. 33F), dorsal cirri basal, slightly longer than neuroacicular lobes. Neuroacicular lobes digitiform, twice longer than postchaetal lobes and neuropodial ligules; pre- or postchaetal lobes digitiform, distally inserted; neuropodial ligules digitiform. Ventral cirri duplicated, of different length, largest ones as long as neuropodial ligules. In anterior chaetigers (Fig. 33G), dorsal cirri medial, half as long as notoacicular lobes. Notoacicular lobes subconical to digitiform, twice longer than neuroacicular ones. Neuroacicular lobes digitiform; postchaetal lobe inconspicuous; neuropodial ligule medially inserted, digitiform, 3/4 as long as neuroacicular lobes. Ventral cirri duplicated, of different length, largest ones slightly shorter than neuropodial ligules. In middle chaetigers (Fig. 33H), dorsal cirri medial; cirrophores very swollen and vascularized, cirrostyles several times longer than neuroacicular lobes. Notoacicular lobes absent, aciculae embedded in cirrophore. Neuroacicular lobes subconical, as long as cirrophores; neuropodial ligules medially inserted, subconical, half as long as neuroacicular lobes. Ventral cirri duplicated, largest one twice longer than neuropodial ligules. Notochaetae and neurochaetae homogomph spinigers throughout body. In chaetigers 2–9, notopodial and neuropodial spinigers arranged in dense bundles, surrounding neuroaciculae, decreasing rapidly in number in subsequent segments and toward posterior end (Figs 33C, E, G, I).

193

Remarks. The specimens match the description by Banse (1977); in order to improve the description and to show the current state of type material, some features are detailed and illustrated. Records for this species are mainly for the Northern Atlantic coasts of the United States. The report by Taylor (1984) seems to refer to the same species; however, the report by Gobin (1990) did not include a description and consequently the distribution into the Caribbean Sea is regarded as doubtful.

Distribution. North Carolina, United States, Gulf of Mexico.

Questionable records

Ceratocephale loveni Malmgren, 1867

Ceratocephale loveni Malmgren 1867: 61–62, Pl. 5, Figs 33, 33A–D. Banse 1977: 615-617 (partim)

Type locality. Bohüslan (Lindö and Koster Islands), Skagerrak coast of Sweden.

Remarks. Banse (1977) added some notes about the specimens based upon type and non-type material of C. loveni and two other species, unfortunately this decision increased greatly the distribution and the definition of the species, currently with a cosmopolitan status. C. loveni requires a revision for its possible restriction to temperate Northeastern Atlantic coasts.

Gymnonereis Horst, 1919

Type species. Gymnorhynchus sibogae Horst, 1918, by original designation (Horst 1918).

Diagnosis (after Darbyshire 2014). Prostomium with anterior margin incised. Eyes present. Four pairs of tentacular cirri. Maxillary ring bare. Oral ring with papillae only in areas V–VIII. First two chaetigers type B. Notopodial ligules absent, projection of the cirrophores present in

194 anterior chaetigers. Notopodial postchaetal lobes present at least on anterior chaetigers; neuropodial postchaetal lobe present at least in anterior and middle chaetigers. Notochaetae and neurochaetae homogomph and sesquigomph spinigers and occasionally homogomph and sesquigomph falcigers, arranged in dense bundles around notaciculae and neuroaciculae.

Remarks. As discussed in other section, here the so-called accessory dorsal cirri were regarded as a projection of the cirrophore. This structure does not present a true joint as observed in the dorsal cirri, and even this structure could be regarded as a notopodial ligule. In Ceratonereis species, notopodial ligules decrease in size toward posterior chaetigers, and the cirrophore eventually carry the reduced notopodial ligules; therefore, the homology about the projection of cirrophores with dorsal cirri or the notopodia ligules require further studies. This genus, as Ceratocephale, has a great amount of spinigers in anterior chaetigers, being a distinctive of these genera. The only species recorded from Grand Caribbean region is G. crosslandi, a Pacific Ocean species.

Questionable records

Gymnonereis crosslandi (Monro, 1933)

Chaunorhynchus crosslandi Monro 1933: 46–49, Text Figs 20a–f. Gymnonereis crosslandi Banse 1977: 623.

Type locality. Gorgona Island, Pacific Panama, at 30 m depth.

Remarks. Monro (1933) placed this species in Chaunorhynchus Chamberlin, 1919, and considered it allied to Ceratocephale and Gymnonereis species. The report by Gillet (1986) from Guadeloupe cannot be corroborated after the short description, and it probably belongs to another species. The description by Taylor (1984) is similar and does not allow to separate it from G. crosslandi, deserving further revision for their possible separation.

195

Kainonereis Chamberlin, 1919

Type species. Kainonereis alata Chamberlin, 1919, by original designation.

Diagnosis (after Chamberlin 1919). Prostomium with anterior margin entire. Antennae bifid in epitokes. Eyes present. Four pairs of tentacular cirri. Pharynx bare. First two chaetigers type C. Dorssal cirri modified, calyx-like, from chaetigers 5–7 notopodial ligules with similar development throughout body. Notopodial prechaetal or postchaetal lobes absent; neuropodial postchaetal lobes present at least on anterior chaetigers. Notochaetae homogomph spinigersand falcigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles.

Remarks. Chamberlin established Kainonereis for a single species with highly modified, bifid antennae (tentacle) and dorsal cirri elytriform in chaetigers 5-7. The type species was regarded as species inquirenda (de León-González & Trovant 2013, Read 2014); however, the genus is valid because there are type and additional materials in good conditions and its features are unique and distinctive. Only two species of this genus have been described, both from the Pacific Ocean and some comments are necessary to clarify its taxonomic status for one of them, and the formation of a new combination for the other. Also, a new species from the Caribbean Sea is diagnosed, being the first one recorded in the Atlantic Ocean. The genus is easily identified by having dorsal cirri elytriform in chaetigers 5-7 only. For both Pacific Ocean species, only epitokes are known, but both atokous and epitokous of K. chamberlini sp. n. were found. There are some doubts concerning how dorsal cirri are modified into elytra-like features, or if they are independent structures. In K. elytrocirra n. comb., the dorsal cirri are inserted to elytral stem, whereas in K. alata and K. chamberlini sp. n., dorsal cirri are distally inserted to elytrae. A recent revision of type material of K. elytrocirra confirms the absence of the basal cirrostyles.

Kainonereis chamberlini sp. n. Figure 16 E-H

196

Type material. Caribbean Sea, Colombia. Holotype USNM P0000, R/V Pillsbury, Cruise 6806, St. –768 (12°33'N 71°11'W), NE Peninsula de la Guajira, Colombia, 65 m, 28 July 1968. Bahamas. Paratype USNM P0000, R/V Pillsbury, Cruise 6406, St. 197 (27°59'N 79°20'W), NW Little Bahama Bank, 576 m, 11 August 1964.

Additional material examined. Central Pacific Ocean, Kiribati. Type material of Kainonereis alata Chamberlin, 1919. Holotype USNM 19388 and paratypes USNM 19386 (5), R/V Albatross, Sta. H3964, Off Apaiang, Gilbert Islands, Kiribati, 3 January 1900, Coll. United States Fish Comission.

Type locality. Northeastern of Península de la Guajira, Colombia, 65 m depth.

Etymology. The name is after Ralph Vary Chamberlin, a very prolific taxonomist in several groups, as a recognizing of his work in polychaete taxonomy, and by proposing the genus Kainonereis.

Description. Holotype complete, 8 mm long, 0.6 mm wide, 56 chaetiger. Paratype complete, prostomium and posterior end damaged, anal cirri missing, 9 mm long, 5 mm wide, 58 chaetigers. Body yellowish, tapering posteriorly, stained with green-methyl to improve contrast (Figs 16E, F). Prostomium longer than wide; antennae subulate, passing the palps; eyes subequal, in trapezoidal arrangement (Fig. E). Tentacular ring as long as first chaetiger; tentacular cirri jointed, longest tentacular cirri reach chaetiger 6 (Fig. 16E). Parapodial cirri pattern. Dorsal cirri calyx-like from 5-7 chaetigers (Figs 16G, H). Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles; heterogomph spinigers and falcigers in sub-acicular fascicles. Pygidium not modified; anal cirri subulate, as long as last 3 segments.

Remarks. Some differences can be advanced between K. chamberlini sp. n and K. alata. The bifid antennae and the notopodial homogomph falcigers, present in epitokous of K. alata, are

197 absent in epitokes of K. chamberlini sp. n. Moreover, the calyx-like dorsal cirri in K. chamberlini is relatively larger than in K. alata. The description is short because the genus will be revised elsewhere.

Laeonereis Hartman, 1945

Type species. Nereis culveri Webster, 1879 by original designation (Hartman 1945).

Diagnosis (after Pettibone 1971). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary ring with verticillated papillae. Oral ring with triangular papillae in areas VI, missing from areas V and VII-VIII. First two chaetigers type C. Notopodial ligules with distinct development throughout body. Notopodial prechaetal lobes and neuropodial postchaetal lobes present at least on anterior chaetigers. Notochaetae homogomph spinigers; neurochaetae homogomph spinigers in both supra- and sub-acicular fascicles, few homogomph falcigers in sub-acicular fascicles in middle and posterior chaetigers.

Remarks. The genus was partially revised by Pettibone (1971); unfortunately, she regarded most species as synonyms of L. culveri. Other authors have rejected this opinion (Orenzanz & Gianuca 1974; Santos & Lana 2001), decision followed here.

Laeonereis nota (Treadwell, 1941)

Leptonereis nota Treadwell 1941: 1–3, Figs 7–10. Laeonereis culveri Pettibone 1971:14–19, Figs 5c, d (partim, non Webster, 1879) Laeonereis nota De Jesús-Flores et al. 2015?

Type locality. Offats Bayou, Galveston, Texas, Gulf of Mexico.

Remarks. Pettibone (1971) regarded most species of Laeonereis described up to date as synonyms of L. culveri (Webster, 1879), regarding them as just variants, althought the different

198 parapodial topologies. However, de Jesús-Flores et al. (2015) recognized differences in relation to L. culveri. In L. nota, the notopodial ligules are subconical with blunt tips in anterior chaetigers, while in L. culveri they are basally very wide, almost rounded structures. In addition, in L. nota, the notopodial ligules are subequal in anterior chaetiger and becoming shorter than toward posterior chaetigers; whereas in L. culveri, notopodial ligules are longer than notopodial ligules in anterior chaetigers and becoming as subequal in posterior ones. In L. nota, the notoacicular and neuroacicular lobes are subequal throughout body and they are more projected than remaining parapodial structures, whereas in L. culveri notoacicular lobes are longer than neuroacicular ones in middle and posterior chaetigers, and notopodial ligules and notoacicular lobes are the projected structures.

Laeonereis culveri (Webster, 1879) Figure 34A–N

Nereis culveri Webster 1879a: 111–113, Pl. 3, Figs 23-30, Pl. 4, Figs 31–32. Laeonereis culveri Hartman 1945: 21. Pettibone 1971: 14–19, Figs 5a, b, 6a–l (partim)

Type locality. Beesley’s Point, Great Egg Harbor, New Jersey.

Type material. Northwestern Atlantic Ocean, United States. Syntypes of Nereis culveri USNM 541 (10), Great Egg Harbor, New Jersey, Coll. H.E. Webster. USNM 28178 (11), Beaufort, North Carolina, Coll. C.L. Culver.

Additional material. Gulf of Mexico, United States. USNM 22093 (5), Ocean Springs, David Bay, Mississippi, 17 August 1949, Coll. GCRL. USNM 22243 (3), Sand Beach, Grand Isle, Louisiana, 12 August 1942, Coll. E.H. Behre.

Description. Syntypes complete, in perfect condition; one specimen selected for dissection, complete, 48 mm long, 1.8 mm wide, 105 chaetigers. Body pale, without pigmentation; greenish glandular masses very conspicuous in both dorsal and ventral sides of body, forming a distinct pattern as follows. In anterior chaetigers, 3–4 well-defined glands on notopodial ligules, one

199 gland on ventral base of parapodium (Figs 34D, E). In middle chaetigers, three well-defined glands on notopodial ligule and a short line on the beginning of parapodium, one gland very reduced on ventral base of parapodium, almost inconspicuous (Figs 34F, G). In posterior chaetigers, 2–3 slightly dispersed glands and three long lines forming a row and crossing each chaetiger, one gland reduced on ventral base of parapodium, glands progressively reduced toward posterior end (Figs 34H, I). Prostomium as long as wide, anterior margin slightly incised, medial groove present along middle prostomium (Figs 34A, B); eyes sometimes faded (Fig. A), or with pigment dispersed (Fig. 34B); antennae conical, half as long as prostomium, not passing palps. Tentacular ring up to 1.5 times longer than first chaetiger dorsally, 3–4 times longer ventrally; four pairs of tentacular cirri, cirrophores conspicuous, largest reach chaetiger 4 (Figs 34A, B). Parapodial cirri pattern: Dorsal cirri much shorter than notopodial ligules throughout body, basally inserted throughout body, slightly displaced toward posterior end. Ventral cirri shorter than neuropodial ligules throughout body, posterolaterally displaced with respect to parapodia and often overlapped in mounted parapodia, basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 34J), dorsal cirri basal, digitiform, half as long as notopodial ligules. Notopodial ligules subconical, blunt tip, 1.5 times longer than neuroacicular ones. Neuroacicular lobes ovate; postchaetal lobes digitiform, 2– 3 times longer than neuroacicular ones; twice longer than neuropodial ligules. Ventral cirri digitiform, half as long as neuropodial ligules. In anterior chaetigers (Fig. 34K), dorsal cirri basal, reduced, digitiform. Notopodial ligules subconical, swollen, lobulated aspect, several times wider than remaining structures, twice longer than prechaetal lobes; notoacicular lobes digitiform, slightly longer than prechaetal ones; prechaetal lobes digitiform. Neuroacicular lobes rounded; postchaetal lobes digitiform, twice longer than neuroacicular ones. Ventral cirri short, 1/3 as long as neuropodial ligules, subequal to dorsal ones. In middle chaetigers (Fig. 34L), dorsal cirri basal, reduced, digitiform. Notopodial ligules digitiform, 1.2 times longer than notoacicular lobes; notoacicular lobes subconical, blunt tip, twice longer than neuroacicular ones; prechaetal lobes absent. Neuroacicular lobes subconical, twice wider than notopodial ones, 1-2 times longer than neuropodial ligules; postchaetal lobes

200 absent; neuropodial ligules digitiform. Ventral cirri button-shaped, posteriorly displaced, covered by surrounding tissue, 1/3 as long as dorsal cirri. In posterior chaetigers (Figs 34M, N), dorsal cirri basal, reduced, digitiform. Notopodial ligules subconical, subequal to notoacicular lobes; prechaetal lobes absent; notoacicular lobes subconical, 2.0-2.5 times longer than neuroacicular ones. Neuroacicular lobes subconical, blunt tip, twice longer than neuropodial ligules; postchaetal lobes absent; neuropodial ligules digitiform. Ventral cirri button-shaped, 1/3 as long as dorsal cirri. Notochaetae homogomph spinigers. Neurochaeta homogomph spiniger in both supra- and sub- acicular fascicles, few homogomph falcigers in sub-acicular fascicles in middle and posterior chaetigers. Pygidium annulated, as long as two segments; anal cirri cirriform, as long as pygidium and last segment together (Fig. 34C).

Variation. There are specimens with different grades of contraction and their widths vary widely, especially in anterior segments. However, parapodial proportions are consistent in all cases, making it difficult a misidentification. In some specimens, the greenish glandular pattern is slightly faded.

Remarks. Syntypes are very large specimens and have a greenish glandular pattern in both dorsal and ventral sides of body; this has been regarded as a specific feature for this species. The glandular pattern and parapodial morphology is almost identical in the additional material examined, only the body size is different; syntypes are larger than additional ones. This species has been largely confused with other Laeonereis species after Pettibone (1971), and recorded from several localities. Material illustrated by Pettibone for Sapelo Island, Georgia (USNM 33294) has some differences with L. culveri. In specimens from Sapelo Island, glandular pattern is represented by several isolated spots in anterior and middle chaetigers, and the very developed notopodial ligule, passing greatly the remaining parapodial structures; possibly specimens belong to a distinct species. Recently, de Jesús-Flores et al. (2015) noted useful morphological differences to separate L. culveri to L. nota and including a key to all known species.

201

Nicon Kinberg, 1865

Type species. Nicon pictus Kinberg, 1865, by subsequent designation (Hartman, 1948).

Diagnosis (after de León-González & Trovant 2013). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Pharynx bare. First two chaetigers type C. Notopodial ligules with variable development throughout body. Notopodial prechaetal lobes present at least on anterior and middle chaetigers; neuropodial postchaetal lobe present throughout body. Notochaetae homogomph spinigers; neurochaetae homogomph spinigers and falcigers in supra- acicular fascicles, homogomph and heterogomph spinigers and heterogomph and sesquigomph falcigers in sub-acicular fascicles.

Remarks. The genus has a high morphologic variation and deserves a further revision for the possible restriction of Nicon and the proposal of new genera. The only species recorded from the Grand Caribbean is N. moniloceras, a species originally described from the Pacific Ocean. After the revision of type material of N. moniloceras, a new species is recognized here from the Grand Caribbean.

Nicon vossae sp. n. Figure 35A–M

Type material. Caribbean Sea, Mexico. Holotype USNM 0000 and paratypes USNM 0000 (2), R/V Gerda, Cruise, St. 951 (21°06’N 86°28’W), W Arrowsmith Bank, Mexico, 249 m depth, 28 January 1968. Venezuela. Paratypes UMML P0000 (3) R/V Pillsbury, Cruise 6806, St. 736 (11°00'N 65°55'W), W Isla La Tortuga, Venezuela, 112 m, 22 July 1968.

Additional material. Northwestern Atlantic Ocean, Bahamas. ECOSUR P0000 (3), R/V Gerda, Cruise 6528, St. 697 (26°29'N 78°42'W), Dundee Bay, Grand Bahama Island, 310 m depth, 22/Jul/1965. Unites States. ECOSUR P0000 (1), R/V Gerda, Cruise 6505, St. 452 (25°03'N 80°10'W), E Key Largo, Florida, 185 m depth, 22 January 1965. ECOSUR P0000 (6), R/V Gerda, Cruise 6802, St. 978 (24°32'N 81°07'W), Pourtalès Terrace, Florida, 183 m depth, 3

202

February 1968. ECOSUR P0000 (8), R/V Pillsbury, Cruise 6406, St. 197 (27°59'N 79°20'W), E Florida, 576 m depth, 11 August 1964. Caribbean Sea, Mexico. ECOSUR P0000 (1), R/V Gerda, Cruise 6717, St. 898 (21°04'N 86°19'W), Arrowsmith Bank, Mexico, 352 m depth, 10 September 1967. Guatemala. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6802, St. 613 (15°58'N 88°20'W), E Punta de Manabique, Bay of Honduras, 39 m depth, 19 March 1967. Grenade. ECOSUR P0000 (1), R/V Pillsbury, Cruise 6907, St. 857 (12°23'N 61°22'W), SE Carriacou, Lesser Antilles, 178 m depth, 3 July 1969. Martinica. ECOSUR P0000 (2), R/V Pillsbury, Cruise 6907, St. 907 (14°27'N 60°58'W), S Martinica, 164 m depth, 9 July 1969. Antigua and Barbuda. ECOSUR P0000 (3), R/V Pillsbury, Cruise 6802, St. 975 (17°29'N 61°55'W), SW Barbuda, Lesser Antilles, 29 m depth, 21 July 1969. St. Vincent and the Grenadines. ECOSUR P0000 (3), R/V Pillsbury, Cruise 6806, St. 876 (13°14'N 61°05'W), E Saint Vincent, 245 m depth, 6 July 1969.

Etymology. This species is named after Nancy Voss, in recognition for her important efforts maintaining the invertebrate collections resulting from the University of Miami Deep Sea Expeditions, and especially the collection of polychaetes because it was the main source of specimens for this study.

Description. Holotype complete, 25 mm long, 2 mm wide, 65 chaetigers; paratypes complete and incomplete, in good condition, one in tube. Body pale, without pigmentation, slightly translucent toward posterior end. Lateral whitish glandular masses, with the same pattern throughout body, more conspicuous toward posterior end. Glands arranged as follows. One gland on dorsal base of parapodium; three glandular masses on notopodial ligules in anterior chaetigers, merging into a large one in middle and posterior chaetigers; a large glandular mass on notoacicular lobes and neuropodial ligules, and a small one on neuroacicular lobes, and two glands on bases of ventral cirri; a discontinuous series crossing each chaetiger (Figs 35B, C). Prostomium longer than wide, anterior margin entire; antennae as long as prostomium, not passing palps; eyes rounded, anterior pair larger than posterior one, in trapezoidal arrangement (Fig. 35A). Tentacular ring 1.5 times longer than first chaetiger; four pairs of tentacular cirri, longest one reaches chaetiger 8 (Fig. 35A).

203

Parapodial cirri pattern: Dorsal cirri longer than dorsal ligule in anterior-most chaetigers, becoming shorter in anterior chaetigers and as long as notopodial ligules in middle and posterior ones, basally inserted in anterior region, displaced medially in middle and posterior chaetigers. Ventral cirri shorter than neuropodial ligules and basally inserted throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 35D), dorsal cirri basal, twice times longer than notopodial ligules. Notopodial ligules digitiform. Neuroacicular lobes subconical, blunt tip, 1/3 as long as postchaetal lobes; postchaetal lobes digitiform, as long as dorsal ligules; neuropodial ligules subconical to digitiform, twice wider than dorsal ligules, as long as postchaetal lobes. Ventral cirri basal, 4/5 as long as neuropodial ligules. In anterior chaetigers (Fig. 35E), dorsal cirri medial, 4/5 as long as notopodial ligules. Notopodial ligules subconical, blunt tip, as long as notoacicular lobes; notoacicular lobes subconical, 2.5 times longer than prechaetal ones; prechaetal lobes rounded. Neuroacicular lobes subconical, blunt tip, half as long as postchaetal lobes; postchaetal lobes digitiform, as long as neuropodial ligules; neuropodial ligules subconical, blunt tip. Ventral cirri basal, half as long as neuropodial ligules. In middle and posterior chaetigers (Figs 35F, G), dorsal cirri medial, 3/4 or as long as notopodial ligules. Notopodial ligules subconical, blunt tip 1.3 times longer than notoacicular lobes; notoacicular lobes subconical, blunt tip, 4 times longer than prechaetal ones; prechaetal lobes rounded. Neuroacicular lobes subconical, blunt tip, half as long as postchaetal lobes; postchaetal lobes digitiform, subequal to neuropodial ligules; neuropodial ligules subconical to digitiform. Ventral cirri basal, 4/5 as long as neuropodial ligules. Notochaetae homogomph spinigers. Neurochaetae homogomph spiniger and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular ones. Notopodial homogomph spinigers pectinate, fine teeth, decreasing in size toward distal end. Neuropodial homogomph spinigers similar to notopodial ones, but blades half as long; heterogomph spinigers pectinate, fine teeth, decreasing in size toward distal end, blades length similar to supra-acicular ones. Neuropodial heterogomph falcigers pectinate, coarse teeth, half to one third of inner margin edentulate, distal tooth hooked, with a long tendon fused to inner margin, sometimes with a mucron (Figs 35H-J); supra-acicular falcigers stouter than sub-acicular ones (Figs 35H, I). Pygidium crenulated; anal cirri subulate, as long as last 10 chaetigers (Fig. 35B)

204

Variation. The length of longest tentacular cirri varied from chaetiger 8 to 15.

Remarks. Nicon vossae sp. n. resembles N. moniloceras in some respects, as in the similar reach of tentacular cirri, but there are evident features that allow separating them. In N. vossae sp. n. the dorsal cirri are shorter than neuropodial ligule in anterior and middle chaetigers, becoming slightly longer in posterior chaetigers; whereas in N. moniloceras they are shorter than notopodial ligules throughout body. In addition, in N. vossae sp. n., the notopodial and notoacicular lobe are subequal in anterior chaetiger, and the notoacicular lobes become markedly thinner and digitiform toward posterior end; while in N. moniloceras, they are subequal and subconical throughout body. Reports of N. moniloceras need to be evaluated to corroborate their belonging to N. vossae sp. n. or if are distinct species.

Note about the tube morphology. Other Nicon species have been found in tubes (Hutchings & Reid 1990). The inner surface of the tube is quite regular and smoother than outer surface. The external surface has attached several foreign bodies such as shell fragments, other biogenic structures and sediment particles. Because there are no algae in the samples examined N. vossae sp. n. constructs their tube within or at the sediment surface. Also, the structure of the tube have a large amount of chaetae derived of the ; their embedment is not merely superficial, but they are embed on the tube matrix together with other hard structures as spicules, shells and remaining of algae.

Questionable records

Nicon moniloceras (Hartman, 1940)

Leptonereis glauca moniloceras Hartman 1940: 217, Pl. 34, Figs 42–46. Nicon moniloceras Hartman 1958: 265; 1959a: 245, 274. ? Platynereis sp. Pettibone 1971: 48.

Type locality. Off Catalina Island, California, 73 m.

205

Type material. Northeastern Pacific Ocean, United States. Holotype of Leptonereis glauca moniloceras LACM-AHF POLY 815 and paratypes LACM-AHF POLY 816 (3), R/V Velero III, Sta. 900-38 (33°24'35"N 118°21'15"W), off Long Point, Santa Catalina Island, California, 12.1 m depth, on rocks, brachiopods and sponges, 18 September 1938, Coll. Allan Hancock Pacific Expedition.

Remarks. Currently this species has a wide distribution. Pettibone (1971), based on the literature, transferred it to Platynereis arguing that parapodial and chaetal morphology is closer to Platynereis than to Nicon, regarding it as questionable. However, a re-examination of type materials by Hutchings & Reid (1990) confirmed the absence of paragnaths and their belonging to Nicon. de León-González & Solís-Weiss (2000) also revised the type series and noted some differences in tentacular cirri between the holotype and paratypes. After the revision of type material, the species is a valid Nicon species as stated by the cited authors. Furthermore, the reports of this species for Atlantic localities might belong to N. vossae sp. n, described above.

Olganereis Hartmann-Schröder, 1977

Type species. Ceratocephala edmondsi Hartman, 1954 by, original designation (Hartmann- Schröder 1977c).

Diagnosis (after Hartmann-Schröder 1977b). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary and oral rings with papillae only. First two chaetigers type C. Notopodial ligules with similar development throughout body. Notopodial postchaetal lobes present, at least on anterior chaetigers; neuropodial postchaetal lobes present throughout body. Notochaetae homogomph spinigers; neurochaetae homogomph spinigers in supra-acicular fascicles, heterogomph falcigers and spinigers in sub-acicular fascicles.

Remarks. The genus resembles Laeonereis and Tylorhynchus by having papillae in both pharyngeal rings, but the shape of the papillae in the former and parapodial features in the latter

206 distinguish them from Olganereis. The genus is only known from Australia, this is the first species from other region.

Olganereis salazari sp. n. Figure 36A–O

Type material. Holotype ECOSUR 0000 and paratypes ECOSUR 0000 (3), R/V Pillsbury, Cruise 6806, St. 761 (11°52'N 70°22'W), W Península de Paraguaná, Venezuela, 35 m, 27 Jul 1968.

Additional material. ECOSUR P0000 (1), Mochima, Cumaná, Venezuela, in Millepora, 1 March 2002, Coll. M. Liñero. Other material examined. Australia. Paratypes of Ceratocephala edmondsi Hartman, 1954. LACM-AHF 845 (7), American River, Kangaroo Island (35°46’45”S 127°46’28”E), Coll, S.J. Edmonds.

Type locality. West of Península de la Guajira, Gulf of Venezuela.

Etymology. The specific name is after Dr. Sergio I. Salazar-Vallejo, mentor, friend and colleague, as a recognition of his numerous contributions to polychaete taxonomy.

Description. Holotype complete, body semi-transparent, yellowish, tapered posteriorly, 16 mm long, 0.8 mm wide, 75 chaetigers. Massive, yellowish glands in dorsal cirri bases, lobes and ligules throughout body, more conspicuous in posterior chaetigers (Figs 36B, C). Prostomium wider than long, antennae subulate, slightly passing palps; anterior pair of eyes larger than posterior ones, rounded, in trapezoidal arrangement (Fig. 36A). Tentacular ring twice longer than first chaetiger; tentacular cirri with conspicuous cirrophores, jointed, longest tentacular cirri reaching chaetiger 3 (Fig. 36A). Pharynx everted; jaws amber with 15 teeth, distal ones rounded, basal ones acute, inner edge completely dentate (Fig. 36K). Maxillary ring: I= 0, II= one large papilla, III= 0, IV= one papilla. Oral ring: V= 0, VI= one small papilla, VII-VIII= five small papillae in one row (Figs 36D, E).

207

Parapodial cirri pattern: Dorsal cirri extending beyond upper notopodial ligules in anterior and posterior chaetigers, smaller, not passing them in middle region chaetigers; basally inserted in anterior region, displaced medially in medial and posterior regions. Ventral cirri shorter than both neuropodial ligules and dorsal cirri throughout body; basally inserted in anterior region, migrating ventrally throughout body. First two chaetigers type C, remaining biramous. In first two chaetigers (Fig. 36F), dorsal cirri basal, subequal to notopodial ligules. Notopodial ligules digitiform, three times longer than neuroacicular lobes, twice longer than postchaetal lobes; postchaetal lobes digitiform. Neuroacicular lobes subconical, blunt tip; postchaetal lobes digitiform, twice longer than neuroacicular ones; neuropodial ligules digitiform, subequal to postchaetal lobes. Ventral cirri basal, half as long as neuropodial ligules. In anterior chaetigers (Fig. 36G), dorsal cirrus medial, subequal to notopodial ligules. Notopodial ligules subconical, blunt tip, as long as wide, three times longer than both notoacicular and postchaetal lobes; notoacicular lobes digitiform, papillae very conspicuous; postchaetal lobes digitiform. Neuroacicular lobes rounded, subequal to postchaetal ones; postchaetal lobes rounded; neuropodial ligules digitiform, subequal to neuroacicular lobes. Ventral cirri basal, half as long as neuropodial ligule. In middle chaetigers (Figs 36H, I), dorsal cirri medial, slightly shorter than notopodial ligules. Notopodial ligules subconical, blunt tip, slightly longer than notoacicular lobes, 2–3 times longer than postchaetal lobes; notoacicular lobes subconical, papillae conspicuous; postchaetal lobes digitiform. Neuroacicular lobes subconical, blunt tip, shorter than or subequal to neuropodial ligules; postchaetal lobes rounded or slightly bilobed, subequal to neuroacicular lobes; neuropodial ligules digitiform. Ventral cirrus basal, half as long as neuropodial ligule. In posterior chaetigers (Fig. 36J), dorsal cirri medial, subequal to notopodial ligule. Notopodial ligules and notoacicular lobes subconical, blunt tip, subequal, 3 times longer than neuroacicular lobes, papillae conspicuous; postchaetal lobes very reduced. Neuroacicular lobes subconical, blunt tip, subequal to postchaetal ones; postchaetal lobes rounded; neuropodial ligules digitiform, twice longer than neuroacicular lobes. Ventral cirri basal, half as long as neuroacicular lobes. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles.

208

Notopodial homogomph spinigers pectinate, fine teeth, decreasing in size toward distal end (Fig. 36N). Neuropodial homogomph spinigers as notopodial ones; heterogomph spinigers pectinate, fine teeth, decreasing in size toward distal end (Fig. 36O). Neuropodial heterogomph falcigers with short blade, 6–5 teeth, distal tooth incurved, fused to blade; supra-acicular falcigers with distal tooth larger than and shaft stouter than sub-acicular ones (Figs 36L, M). Pygidium without modification; anal cirri subulate, as long as last 5-6 segments (Fig. 36B).

Remarks. The new species has similar morphology to N. edmondsi, and match well with the current definition of the genus (Hartmann-Schröder 1977c), therefore the new species is placed into Olganereis. O. salazari sp. n. differs, however, from O. edmondsi (Hartman, 1954) in some important features. In O. salazari sp. n., the pharynx has small, rounded papillae arranged into a series, whereas in O. edmondsi the papillae are long and filiform and set into distinct bands. The shape of the papillae in both species was regarded as a specific feature. In addition, in O. salazari sp. n., the dorsal cirri and notoacicular lobes are subequal to each other, whereas in O. edmondsi dorsal cirri are several times than all remaining parapodial structures. As in the new species. O. edmondsi have glandular patches at dorsal and ventral bases of chaetigers, increasing their size posteriorly, especially the dorsal ones.

Profundilycastis Hartmann-Schröder, 1977 Reinstated

Type species. Namalycastis profundus Hartman, 1965, by original designation (Hartmann- Schröder 1977c).

Diagnosis (after Hartmann-Schröder 1977b). Prostomium with an anterior constriction at the tentacular cirri region, anterior margin entire. Eyes absent. Four pairs of tentacular cirri. Pharynx bare. First two chaetigers type C. Notopodial ligules absent throughout body. Notoacicular lobes and neuropodial ligules long and narrow. Neuropodial postchaetal lobes inconspicuous. Notochaetae homogomph spinigers and falcigers, neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph falcigers and spinigers in sub- acicular fascicles.

209

Remarks. This genus was synonymized with Rullierinereis by Böggemann (2009). However, neither parapodia illustrated resembles those found in the type species, Rullierinereis zebra (Rullier, 1963) (Pettibone 1971), nor in other species; other differences are found in the shape of prostomium and the tentacular ring. In Profundilycastis, the prostomium has a constriction at the tentacular cirri region, givig it a funnel-like shape, and ligules are digitate, large and narrow; whereras in Rullierinereis the prostomium and tentacular ring have a similar width, and the ligules and lobes are subconical to digitiform. In addition, the notopodial homogomph falcigers in Profundilycastis have long blades with several teeth; while in Rullierinereis, they have short blades and few teeth, if present. The genus resembles the deep-sea genus Typhlonereis by the absence of chaetae in first two chaetigers, the bare pharynx and by having enlarged, conical parapodial processes, but they differ in the presence of notopodial homogomph falcigers in Profundilycastis. Here the genus is regarded as distinct after the revision of the holotype of P. profundus.

Profundilycastis profundus (Hartman, 1965) Figure 16 I, J

Namalycastis profundus Hartman 1965:84–85, Pl. 11, Figs a, b. Profundilycastis profundus Hartmann-Schröder 1977b: 149–150; Hartmann-Schröder 1977c: 45– 48.

Type locality. Off Bermuda (32°21’18’’N 64°33’00’W), 1000 m.

Type material. Holotype of Namalycastis profundus LACM-AHF 841, R/V Stranger, Sta. Bermuda 8 (32° 21'18" N, 64° 33' W), Bermuda Slope, 1000 m depth, 5 September 1961, Coll. L. Sanders.

Remarks. The species was redescribed by Hartmann-Schröder and she proposed a new genus. The only specimen is in poor condition; therefore, additional dissections were not performed, but it was whole-mounted and examined under microscope.

210

This species is very similar to N. caecoides. However, the absence of notopodial ligules throughout the body, the pharyngeal ormanentation, and the distinct shape of blades of falcigers would be sufficient for separating them. The absence of notopodial ligules was corroborated in the holotype, but their presence cannot be completely discarded since in N. caecoides they detach easily. Paragnaths or papillae were regarded as absent by Hartmann-Schröder. No structures were seen by transparency but dissection was not performed due to the poor condition of the specimen. Analysis of additional material of the genus, as those examined by Böggeman (2009), would be useful for determining if Profundilycastis species have pharyngeal armature, or in fact, the pharynx is bare. Deep-sea nereidids deserve further revisions to evaluate their generic affinities. Böggemann (2009) identified as P. profundus some specimens from the Angola Basin (3945– 3995 m depth), in the Southeastern Atlantic. He regarded Profundilycastis as a synonym of Rullierinereis, based upon the presence of notopodial homogomph falcigers and similar presence of lobes and ligules in parapodia. However, as stated in the remarks above, the shape of both anterior end and parapodia differ from the current definition of Rullierinereis, and consequently Profundilycastis is regarded as distinct. On the other hand, the specimens from Bermuda differ from Angola specimens by having notoacicular lobes 2–3 times longer than neuroacicular ones, presumably they belong to a different species.

Rullierinereis Pettibone, 1971

Type species. Leptonereis zebra Rullier, 1963, by original designation (Pettibone 1971).

Diagnosis (after Pettibone 1971). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Pharynx bare; papillae or paragnaths present on oral ring of R. bahamensis (Hartman-Schröder, 1958) only. First two chaetigers type C. Notopodial ligules with similar development throughout body, sometimes reduced or absent. Prechaetal or postchaetal notopodial lobes absent; neuropodial postchaetal lobes present throughout body. Notochaetae homogomph spinigers and falcigers; neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph falcigers and spinigers in sub-acicular fascicles.

211

Remarks. The genus is closely related with the genus Nicon, but it can be recognized by the absence of a notopodial prechaetal lobe and the presence of notopodial homogomph falcigers, replacing homogomph spinigers in posterior chaetigers. Valuable features for the recognition of species are the presence or absence of notopodial ligules and the shape and dentition of homogomph falcigers (Núñez & Brito 2006).

Rullierinereis bahamensis (Hartmann-Schröder, 1958)

Tylorrynchus bahamensis Hartmann-Schröder 1958: 238–239, Pl. 39, Figs 14–18. Rullierinereis bahamensis Pettibone 1971: 5, 31-33.

Type locality. Bimini, Bahamas, 1 m depth.

Remarks. The sample of this species was taken from a hole made to 2 m from the water and at 30 cm depth, in the entrance of the Bimini Lagoon (Hartmann-Schröder 1958). The species appears to be related with genera Websterinereis, Kinberginereis and Gymnonereis by the presence of papillae in oral ring only, but parapodial and chaetal features separate them. The record by Liñero-Arana & Reyes-Vásquez (1979) from the Gulf of Cariaco does not belong to this species because their specimens have pharyngeal armature; the authors reported papillae but it could be paragnaths due to the small size of specimens. Therefore, the specimens would belong to the Nereis group with homogomph falcigers with strongly bifid blades such as N. panamensis and N. lanai.

Questionable records

Rullierinereis mexicana (Treadwell, 1942)

Type locality. Topolobampo Bay, Sinaloa, Pacific Ocean.

212

Remarks. Type locality corrected by Salazar-Vallejo & Jiménez-Cueto (1997). Pettibone (1971) extended the distribution to Galapagos Island and Florida, but the Atlantic specimens might belong to other species.

Stenoninereis Wesenberg–Lund, 1958

Type species. Stenoninereis martini Wesenberg-Lund, 1958, by original designation.

Diagnosis (after de León-González & Solís-Weiss 1997). Prostomium with anterior margin incised. Antennae present. Eyes present. Four pairs of tentacular cirri. Pharynx bare. First two chaetigers type A. Notopodial surfaces ciliated at least in anterior chaetigers. Dorsal cirri with cirrophores very distinct. Notopodial ligules present in anterior chaetigers, absent in posterior- most ones, neuropodial ligules absent throughout body. Notochaetae homogomph spinigers; neurochaetae homogomph spinigers and heterogomph falcigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles. Pygidium wing-like.

Remarks. Currently, Stenoninereis is the only endemic nereidid genus from the Grand Caribbean region, here all previously known species, and a new one are described. de León González & Solís-Weiss (1997) used the deeply incised anterior margin of prostomium as a specific feature of S. tecolutlensis, and noted that S. martini has a shallow incision as well. Stenoninereis has some other distinctive features, such as the sub-acicular spinigers with very coarse teeth and the wing- like pygidium (i.e. lobes flattened, plate-like), a term proposed by Wesenberg-Lund (1958).

Stenoninereis lackeyi (Hartman, 1958) reinst. Figure 37A–M

Nicon lackeyi Hartman 1958: 263–265, Figs 1–5. Stenoninereis martini Pettibone 1971: 39-41 (partim, non Wesenberg-Lund, 1958)

Type locality. Warm Mineral Springs, Florida, Gulf of Mexico.

213

Type material. Gulf of Mexico, United States. Paratypes of Nicon lackeyi LACM-AHF 806 (8), Warm Mineral Springs, Sarasota County (27°02’43”N 82°17’35”W), Coll. J. Lackey.

Additional material. Gulf of Mexico, United States. USNM 45699 (60), Tampa Bay, Florida, 1963, Coll. J.L. Taylor.

Description. Paratypes (LACM-AHF 806) complete, some parapodia previously dissected; one complete paratype selected for description, 4 mm long, 0.5 mm wide, 31 chaetigers (Fig. 37A). Specimens from Florida (USNM 45699), most complete, some filled with oocytes, generally two per segment (Fig. 37B); one used for variation, 4.8 mm long, 0.4 mm wide, 33 chaetigers (Fig. 37B). All specimens with body pale, without pigmentation. Prostomium wider than long, anterior margin shallowly incised; antennae cirriform, half as long as prostomium; eyes black, anterior pair reniform, slightly larger than rounded posterior pair, anterior and posterior pairs slightly overlapped (Fig. 37C). Tentacular ring half as long as first chaetiger; four pairs of tentacular cirri, longest reach chaetiger 7, anterodorsal pair lanceolate (Figs 37A, C). Pharynx previously removed in paratype. In non-type material, jaws light brown, 10 teeth restricted to half cutting edge (Fig. 37E). Pharynx bare. Parapodial cirri pattern: cirrostyles shorter than, to subequal to, neuroacicular lobes and basally inserted to notoacicular lobes throughout body; dorsal cirri subequal to or slightly longer than notoacicular lobes in anterior chaetigers, becoming longer toward posterior chaetigers. Ventral cirri shorter and basally inserted throughout body. First two chaetigers type A, remaining biramous. In first two chaetigers (Fig. 37I), notopodia small lobes. Neuroacicular lobes subconical, blunt tip, twice longer than ventral cirri. In remaining chaetigers (Figs J–M), dorsal cirri shorter than, or as long as, the width of their respective chaetigers; cirrophores twice longer than cirrostyles in anterior chaetiger, becoming 1.5 times longer toward posterior chaetigers. Notoacicular lobes subconical, blunt tip, twice longer than wide in anterior chaetigers, up to 4 times longer than wide toward posterior chaetigers; notopodial ligules digitiform, half as long as, or subequal to, notoacicular lobes in anterior chaetigers, becoming shorter toward middle chaetigers and disappearing from chaetigers 23–25. Neuroacicular lobes subconical, blunt tip, 2.0–2.5 times longer than wide, subequal to

214 notoacicular lobes throughout body. Ventral cirri cirriform, half as long as neuroacicular lobes throughout body. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles. Pygidium wing-like; anal cirri subulate, as long as last four chaetigers (Figs 37B, D).

Variation. The reach of tentacular cirri in specimens from Tampa Bay ranged from 6 to 8 chaetigers. The ratio between the length of dorsal cirrus and notoacicular lobe ranged from 1.5 to 2.0. The most variable feature, as in all species of Stenoninereis, was the relative length of notopodial ligules; in S. lackeyi, ratio between the length of notopodial ligule and notoacicular lobes ranged from 0.5 to 1.0 in anterior chaetigers. The lanceolate aspect of anterodorsal pair of tentacular cirri sometimes is not fully developed.

Remarks. The holotype and part of the paratype series deposited in USNM type material was not found during my research visit, but redescribed with paratypes deposited in LACM. The main distinctive feature for this species is the very incised aspect of the region between parapodial rami. This feature is given by the slender noto- and neuroacicular lobes (several times longer than wide). Pettibone (1971) regarded S. lackeyi as a junior synonym of S. martini; however, there are some features for recognizing S. lackeyi as a valid species. In S. lackeyi cirrostyles are 1.5 to 2 times longer than their respective cirrophores, while in S. martini they are subequal or up to 1.5 longer than their respective cirrophores. In addition, the anterodorsal pair of tentacular cirri is usually lanceolate in S. lackeyi, whereas in S. martini this attribute was not observed in syntypes examined, nor reported by other authors; the lanceolate tentacular cirri are also present in S. elisae sp. n. On the other hand, Wesenberg-Lund (1958) pointed out that there is only one oocyte per segment, while in S. lackeyi there are two. Consequently, S. lackeyi can be easily recognized from the other Gulf of Mexico species: S. tecolutlensis. In S.lackeyi the prostomium is shallowly incised, and the eyes are minute, whereas in S. tecolutlensis the prostomium is deeply incised and the eyes occupy a larger prostomial surface. In addition, in S. lackeyi the neuroacicular lobes are subconical, whereas in S. tecolutlensis they are acuminate. On the other hand, the main difference between S. lackeyi and S. elisae sp. n. is the different development of the dorsal cirri in anterior and middle chaetigers.

215

Stenoninereis martini Wesenberg-Lund, 1958 Figure 38D-F

Stenoninereis martini Wesenberg-Lund 1958: 9–12, Figs 2, 3, 4a–c. Pettibone 1971: 39–41 (partim). Stenoninereis martinsi (sic) Hartmann-Schröder 1977a: 60–62, Figs 25–28.

Type locality. Fontein, Bonaire.

Type material. Caribbean Sea, Netherlands Antilles. Syntypes USNM 29726 (2), Devil’s Hole Swamp, St. Martin, 26 July 1955, Coll. P.W. Hummelinck.

Description. Syntypes in poor conditions, almost dried, dissections previously performed (Figs D, E). One syntype incomplete posteriorly, some anterior parapodia dissected previously, 2.4 mm long, 0.4 mm wide, 16 chaetiger (Fig. E). Another one in two portions, anterior end very damaged, 3.5 mm long, 0.3 mm wide, 21 chaetigers, some posterior chaetigers previously dissected (Fig. D). Prostomium wider than long, anterior margin shallowly incised; antennae cirriform, half as long as prostomium; eyes black, subequal, anterior pair reniform, posterior eyes rounded, anterior and posterior pairs overlapped (Fig. 38F). Tentacular ring half as long as first chaetiger; four pair of tentacular cirri, longest reach chaetiger 8. Parapodial cirri pattern: dorsal cirri passing all parapodial structures and basally inserted throughout body. Ventral cirri half as long as neuroacicular lobes and basally inserted throughout body. First two chaetigers type A, remaining biramous. In first two chaetigers, notopodia as small lobes. Neuroacicular lobes twice longer than ventral cirri. In remaining chaetigers, dorsal cirri twice longer than notoacicular lobes throughout body, as long as, or slightly longer than, the width of their respective chaetigers throughout body; cirrostyles as long as cirrophores in anterior and posterior chaetigers, slightly shorter in middle chaetigers. Notoacicular lobes subconical to digitiform; notopodial ligules half as long as notoacicular lobes in anterior chaetigers, becoming shorter toward middle chaetigers and

216 disappearing in posterior-most chaetigers. Neuroacicular lobes slightly shorter than notoacicular lobes in anterior chaetigers, as long as notoacicular lobes in middle and posterior ones. Ventral cirri cirriform. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers in supra-acicular fascicles, heterogomph spiniger and falcigers in sub-acicular fascicles. Pygidium wing-like; anal cirri missing (Fig. 38D).

Remarks. The syntypes revised are in bad conditions, almost dried. Parapodia dissections were no performed, but specimens were whole-mounted and examined under microscope. Original parapodia drawings match the specimens examined. Pettibone (1971) redescribed the syntypes from the USNM; however, some illustrated features are doubtful. For example, the parapodia illustrated have lobes and ligules with distinct development from than observed in syntypes and illustrated by Wesenberg-Lund (1958). Pettibone emended the diagnosis including subbiramous parapodia in first two chaetigers, i.e. with notoaciculae and dorsal cirri, instead of uniramous ones (here type A); in fact, the first parapodium (Pettibone 1971, Fig. 23f) was figured with two aciculae, and the second with a dorsal cirrus (Figs 23a, g). However, notopodial rami are very reduced and neither notoaciculae nor dorsal cirri are actually present in the first two chaetigers. De León-González & Solís-Weiss (1997), perhaps following Pettibone, included subbiramous parapodia in their diagnosis, but in the description of S. tecolutlensis they declared the first two chaetigers as uniramous. Finally, one of the pygidium illustrated by Pettibone (1971, Fig. 23e) corresponds at the current state of the syntypes, but the whereabouts of the second posterior end is unknown. The relative size of cirrostyles and cirrophore vary among adjacent chaetigers. However, the ratio between the total length of dorsal cirri and the width of their respective segment varies little; it was used to distinguish S. martini from S. elisae sp. n.

Distribution. Caribbean Sea.

Stenoninereis tecolutlensis de León-González & Solís-Weiss, 1997 Figure 38A–C

217

Stenoninereis tecolutlensis de León-González & Solís-Weiss 1997: 199–202, Figs 2A-F.

Type locality. Estero de Larios, Tecolutla, Mexico, Gulf of Mexico.

Type material. Gulf of Mexico, Mexico. Holotype USNM 174870, Estero de Larios, Tecolutla, Veracruz, 19 November 1994, Coll. J.A. de León-González.

Description. Holotype complete, 14 mm long, 0.7 mm wide, 49 chaetigers. Body pale, brown pigmentation present in prostomium, around eyes and palpophores; in first segments slightly faded, bands occupying almost complete wide of each segment, discoloring toward posterior end (Fig. 38A). Prostomium wider than long, anterior margin deeply incised, incision reaches posterior pair of eyes; eyes black, rounded, anterior pair larger tan posterior one, slightly overlapped (Fig. 38B). Tentacular ring as long as first chaetiger; four pairs of tentacular cirri (3 tentacular cirri missing in holotype), longest reach chaetiger 10 (Fig. 38A). Notopodial ligules inconspicuous since chaetiger 37. Pygydium wing-like; anal cirri missing (Fig. 38C).

Remarks. The description match well with the specimen, no redescription or further details seem to be necessary because S. tecolutlensis is easily distinghished by the deep incision in the prostomium, while in the remaining species of the genus it is shallow.

Distribution. Gulf of Mexico.

Stenoninereis elisae sp. n. Figure 39A–M

Type material. Caribbean Sea, Puerto Rico. Syntypes USNM 55366 (3), Espiritu Santo River, Rio Grande, April 1977, Coll. W.R. Bhajan.

218

Additional material. Caribbean Sea, Puerto Rico. USNM 55360 (3), 50 m above Castanon confluent, Espiritu Santo River, Rio Grande, Coll. W.R. Bhajan.

Etymology. The name is after Elise Wesenberg-Lund, by their extense work in polychaete taxonomy, especially freshwater ones, and by proposing Stenoninereis. The name is a noun in apposition, and was formed by elision of the last vowel for euphony.

Description. Three syntypes complete, in good condition. Specimen dissected complete, 62 mm long, 0.4 m wide, 27 chaetigers. Body pale, without pigmentation (Figs 39A, B). Prostomium wider than long, anterior end shallowly incised; antennae cirriform, half as long as prostomium; eyes black, anterior eyes smaller than posterior ones and fused with, forming a piriform spot (Fig. 39C). Tentacular ring as long as first chaetiger; four pairs of tentacular cirri, longest reach chaetiger 6, anterodorsal pair slightly wider than remaining ones or lanceolate (Figs 39A–C). Pharynx dissected; jaws brown, 8 teeth. Pharynx surface bare. Parapodial cirri pattern: dorsal cirri longer than neuroacicular lobes and basally inserted to notoacicular lobes throughout body. Ventral cirri shorter and basally inserted throughout body. First two chaetigers type A, remaining biramous. In first two chaetigers, notopodia as small lobes. Neuroacicular lobes twice longer than ventral cirri. In remaining chaetigers (Figs 39E–H), dorsal cirri 2.0–2.5 times longer than the width of their respective chaetigers in anterior and middle body, becoming as long as in posterior-most chaetigers; cirrostyles much shorter than cirrophores in anterior and middle chaetigers, 1/3 as long as cirrophores in posterior ones. Notoacicular lobes subconical, blunt tip, 1.5 to 2 times longer than wide; notopodial ligules digitiform, twice longer than notoacicular lobes in anterior chaetigers, becoming shorter toward middle chaetigers and disappearing from chaetigers 23–25. Neuroacicular lobes subconical, blunt tip, 1.0–1.5 times longer than wide, slightly shorter than notoacicular lobes in anterior and middle chaetigers, becoming slightly longer than notoacicular lobes in posterior ones. Ventral cirri cirriform, half as long as neuroacicular lobes throughout body. Notochaetae homogomph spinigers. Neurochaetae homogomph spinigers in supra-acicular fascicles, heterogomph spinigers and falcigers in sub-acicular fascicles.

219

Notopodial homogomph spinigers (Fig. 39M) pectinate, fine teeth, decreasing in size toward distal end. Neuropodial homogomph spinigers as notopodial ones; heterogomph spinigers serrate, 2/3 of inner margin dentate, distal tooth subulate (Fig. 39I). Neuropodial heterogomph falcigers with long blade, blade increasing their length toward posterior chaetigers; supra-acicular blades shorter than sub-acicular ones, blades increasing their length from dorsal to ventral position in the same fascicle (Figs 39J–L). Pygidium wing-like; anal cirri as long as last five chaetigers (Fig. 39D).

Variation. The specimens vary little in body length and shape. The length of notopodial ligule is the most variable feature in anterior chaetiger, ranged from 1.5 o 2.5.

Remarks. Due to the variability observed in species belonging to Stenoninereis, designation of syntypes was preferred. This species is easily recognized by the enlarged dorsal cirri in anterior and middle chaetigers, being two or three times longer than their respective segments, and by the very short cirrostyles in those segments. Stenoninereis elisae sp. n. resembles S. martini, but they differ in some parapodial features. In S. elisae sp. n., the notopodial ligules are longer than notoacicular lobes in anterior chaetiger and become as long as them in middle chaetigers, whereas in S. martini they are shorter than notoacicular lobes throughout body. Furthermore, in S. elisae sp. n., the dorsal cirri in anterior and middle chaetigers are 2–3 times the width of segment (without parapodia), whereas in S. martini their length is up to 1.5 times the width of their respective chaetigers. Moreover, in S. elisae sp. n., the cirrostyle is much shorter than cirrophore throughout body, while in S. martini they are subequal throughout body. Furthermore, in S. elisae sp. n., the upper falcigers in sub- acicular fascicles have conspicuous teeth toward distal end, whereas the falcigers in S. martini lack these teeth. On the other hand, S. elisae sp. n. differs from S. tecolutlensis in these same features. In S. elisae sp. n., the dorsal cirri extend greatly beyond notoacicular lobes throughout body, while in S. tecolutlensis they are subequal throughout body. This difference can be noticed by comparing the new species and S. lackeyi. In addition, the medial groove of prostomium in S. tecolutlensis reaches the posterior pair of eyes, whereas in S. elisae sp. n. the groove is very small and does not reach the anterior eyes.

220

Distribution. Puerto Rico.

Websterinereis Pettibone, 1971

Type species. Nereis tridentata Webster, 1880, by original designation (Pettibone 1971).

Diagnosis (after Pettibone 1971). Prostomium with anterior margin entire. Eyes present. Four pairs of tentacular cirri. Maxillary ring bare. Oral ring with papillae only; areas VI and VII-VIII present, V absent. First two chaetigers type C. Notopodial ligules with similar development throughout body. Notopodial prechaetal lobes present at least in anterior chaetigers; neuropodial postchaetal lobes present throughout body. Notochaetae homogomph spinigers; neurochaeta homogomph spiniger in supra-acicular fascicles, heterogomph spinigers and falciger in sub- acicular fascicles.

Remarks. The species are distributed in Atlantic, Pacific and Indian Oceans. New species have not been described since the establishment of the genus by Pettibone. Websterinereis species are characterized by the presence of papillae on oral ring only in combination with their parapodial topology.

Questionable records

Websterinereis tridentata (Webster, 1879)

Nereis tridentata Webster 1879a: 113–115, Pl. 4, Figs 33–40; 1886: 142, Pl. 7, Figs 33–40. Websterinereis tridentata Pettibone 1971: 21–23, Figs 8a–g, 9a–k.

Type locality. New Jersey, Northwestern Atlantic Ocean.

221

Remarks. Reports of this species from the Grand Caribbean region include Cuba and Colombia (Dean 2012), but any of those reports included descriptions to ensure its presence in the region; however the presence cannot be completely discarded until a further revision of materials from Northwestern Atlantic Ocean and the Caribbean Sea.

Acknowledgements

This work was made possible thanks to the support of some colleagues. Also, I thank the curators and collections managers by facilitating materials: Karen Osborn and Geoff Keel (USNM), Norma Emilia González (ECOSUR), and Leslie Harris (AHF). Sergio I. Salazar-Vallejo, Luis F. Carrera-Parra and Jesús Á. de León-González read carefully the manuscript and made valuable suggestions, and for allowing the revision of materials in their respective collections. A special acknowledgement to S.I. Salazar-Vallejo by the persistent support in several ways for carrying out this research. I am grateful to Leslie Harris and David Ocker by their hospitality in Los Angeles. I thank my wife Astrid. E. Te Gómez by her full support during this research. The author and a whole generation of Latin American polychaetologist are in debt with Nancy Voss by allowing us to examine the polychaete material from the University of Miami Deep Sea Expeditions.

References

Ackermann, C., Dorresteijn, A.W.C. & Fischer, A. (2005) Clonal Domains in Postlarval Platynereis dumerilii (Annelida: Polychaeta). Journal of Morphology 266, 258–280.

Aguado, M.T. & San Martín, G. (2009) Phylogeny of Syllidae (Polychaeta) based on morphological data. Zoologica Scripta 38, 379–402.

Amaral, A.C.Z. & Nonato, E.F. (1975) Algunos anélidos poliquetos encontrados en paneles de substrato artificial en el Golfo de Cariaco, Cumana, Venezuela. Boletín del Instituto Oceanográfico de Venezuela 14, 233–242.

222

Amoureux, L. (1985) Annélides benthiques récoltés à l’entrée de la lagune de la Manche-à-Eau, Guadeloupe (Antilles). Bulletin du Muséum National d’Histoire Naturelle, Paris, 4 sér., 7, section A 1, 93–107.

Arias, A. & Carrera-Parra, L.F. (2014) First record of the genus Kuwaita (Annelida: Lumbrineridae) in Europe with the description of a new species and new ultramorphological data for the genus. Zootaxa 3887, 68–78.

Arteaga-Flórez, C. & Londoño-Mesa, M.H. (2015) Neréididos (Nereididae, Polychaeta, Annelida) asociados a raíces de mangle rojo, Rhizophora mangle, en islas San Andrés y Providencia, Caribe Colombiano. Boletín de Investigaciones Marinas y Costeras 44, 163– 184.

Augener, H. (1906) Reports on the results of dredging, under the supervision of Alexander Agassiz, in the Gulf of Mexico and the Caribbean Sea, and on the east coast of the United States, 1877 to 1880, by the U.S.S. Coast Survey Steamer “Blake”, Lieut. Commander C.D. Sigsbe. Bulletin of the Museum of Comparative Zoology 43, 91–196.

Augener, H. (1918) Polychaeta. In: W. Michaelsen (Ed), Beiträge zur Kenntnis der Meeresfauna Westafrikas. Band II. L. Friederischen & Co., Hamburg, pp. 67–625.

Augener, H. (1922) Ueber litorale Polychaeten von Westindien. Sitzungsberichte der Gesellschaft Naturforschender Freunde zu Berlin, Jahrgang 1922, 38–53.

Augener, H. (1922) Revision der Australischen Polychaeten-Typen von Kinberg. Arkiv för Zoologi 14, 1–42.

Augener, H. (1925) Über westindische und einige andere Polychaeten-Typen von Grube (Öersted), Krøyer, Mörch und Schmarda. Publikationer fra Universitetets Zoologiske Museum Kobenhavn 39, 1–47.

Augener, H. (1933a) Polychaeten aus den Zoologischen Museen von Leiden und Amsterdam. Zoologische Mededeelingen 15, 177–260.

Augener, H. (1933b) Zoologische Ergebnisse einer Reise nach Bonaire, Curaçao und Aruba im Jahre 1930. 3. Süβwasser Polychaeten von Bonaire. Zoologische Jahrbücher, Abteilung für Systematik, Ökologie und Geographie der Tiere 64, 351–356.

223

Augener, H. (1934) Polychaeten aus den zoologischen Museen von Leiden und Amsterdam. IV (schluss). Zoologische Mededeelingen 17, 67–160.

Augener, H. (1936) Polychaeten aus den marinen Salinen von Bonaire und Curaçao. Zoologische Jahrbücher, Abteilung für Systematik, Ökologie und Geographie der Tiere 67, 337–352.

Bakken, T. (2003) Redescription and resurrection of Typhlonereis gracilis Hansen, (Polychaeta, Nereididae). Sarsia 88, 346–352.

Bakken, T. (2007) Revision of Pseudonereis (Polychaeta, Nereididae). Zoological Journal of the Linnean Society 150, 145–176.

Bakken, T., Glasby, C.J. & Wilson, R.S. (2009) A review of paragnath morphology in Nereididae (Polychaeta). Zoosymposia 2, 305–316.

Bakken, T., Kongsrud, J.A., Oug, E., Cochrane, S.K.J., Moen, T.L. & Solbakken, B.E.B. (2010) Polychaetes from Jan Mayen (Annelida, Polychaeta). Polar Research 29, 1–21.

Bakken, T. & Wilson, R.S. (2005) Phylogeny of nereidids (Polychaeta, Nereididae) with paragnaths. Zoologica Scripta 34, 507–547.

Banse, K. (1977) Gymnonereidinae new subfamily : the Nereididae ( Polychaeta ) with bifid parapodial neurocirri. Journal of Natural History 11, 609–628.

Bass, N.R. & Brafield, A.E. (1972) The life-cycle of the polychaete Nereis virens. Journal of the Marine Biological Association of the United Kingdom 52, 701–726.

Blanchard, E. (1849) Anelides. In: C. Gay (Ed), Historia Física y Política de Chile. Zoología, Tomo III. Casa del Autor, París, pp. 9–52.

Böggemann, M. (2009) Polychaetes (Annelida) of the abyssal SE Atlantic. Organisms, Diversity & Evolution 9, 251–428.

Carrera-Parra, L.F. & Salazar-Vallejo, S.I. (2011) Redescriptions of Eunice filamentosa and E. denticulata and description of E. tovarae n. sp. (Polychaeta: ), highlighted with morphological and molecular data. Zootaxa 2880, 51–64.

Chamberlin, R.V. (1919) The Annelida Polychaeta. Memoirs of the Museum of Comparative Zoology at Harvard College 48, 1–514.

224

Chambers, S. & Garwood, P. (1992) Polychaetes from Scottish Waters. A Guide to Identification. 3. Family Nereidae. National M. Edinburgh.

Conde-Vela, V.M. (2013) Namalycastis occulta n. sp. and a new record of N. borealis (Polychaeta: Nereididae: Namanereidinae) from the Northwestern Caribbean Sea. Zootaxa 3721, 475–487.

Conde-Vela, V.M. & Salazar-Vallejo, S.I. (2015) Redescriptions of Nereis oligohalina (Rioja, 1946) and N. garwoodi González-Escalante & Salazar-Vallejo, 2003 and description of N. confusa sp. n. (Annelida, Nereididae). Zookeys 518, 15–49.

Cossu, P., Maltagliati, F., Lai, T., Casu, M., Curini-Galleti, M. & Castelli, A. (2012) Genetic structure of Hediste diversicolor (Polychaeta, Nereididae) from the northwestern Mediterranean as revealed by DNA inter-simple sequence repeat (ISSR) markers. Marine Ecology Progress Series 452, 171–178.

Coutihno, M.C.L., Paiva, P.C. & Santos, C.S.G. (2015) Morphometric analysis of two sympatric species of Perinereis (Annelida: Nereididae) from the Brazilian coast. Journal of the Marine Biological Association of the United Kingdom, 1–7.

Darbyshire, T. (2014) Intertidal and nearshore Nereididae (Annelida) of the Falkland Islands, southwestern Atlantic, including a new species of Gymnonereis. Zookeys 427, 75–108.

Day, J.H. (1973) New Polychaeta from Beaufort, with a key to all species recorded from North Carolina. NOAA Technical Report NMFS CIRC-375, 153. de León-González, J.Á. (2009) Nereididae. In: J. Á. de León-González, J. R. Bastida-Zavala, L. F. Carrera-Parra, M. E. García-Garza, A. Peña-Rivera, S. I. Salazar-Vallejo, and V. Solís- Weiss (Eds), Poliquetos (Annelida: Polychaeta) de México y América Tropical. Universidad Autónoma de Nuevo León, Monterrey, pp. 325–354. de León-González, J.Á. & Goethel, C.A. (2013) A new species of Perinereis (Polychaeta, Nereididae) from Florida , USA , with a key to all Perinereis from the American continent. Zookeys 312, 1–11. de León-González, J.Á. & Solís-Weiss, V. (1997) A new species of Stenoninereis (Polychaeta: Nereididae) from the Gulf of Mexico. Proceedings of the Biological Society of Washington

225

110, 198–202. de León-González, J.Á. & Solís-Weiss, V. (1998) The genus Perinereis (Polychaeta: Nereididae) from Mexican littoral waters, including the description of three new species and the redescriptions of P. anderssoni and P. elenacasoae. Proceedings of the Biological Society of Washington 111, 674–693. de León-González, J.Á. & Solís-Weiss, V. (2000) A review of the polychaete Family Nereididae from Western Mexico. Bulletin of Marine Science 67, 549–569. de León-González, J.Á., Solís-Weiss, V. & Ochoa-Rivera, V. (1999) Nereidids (Polychaeta) from the Caribbean Sea and adjacent coral islands of the southern Gulf of Mexico. Proceedings of the Biological Society of Washington 112, 667–681. de León-González, J.Á., Solís-Weiss, V. & Valadez-Rocha, V. (2001) Two new species of Platynereis (Polychaeta: Nereididae) from Eastern Mexican shores. Proceedings of the Biological Society of Washington 114, 389–395. de León-González, J.Á. & Trovant, B. (2013) A new species of Nicon Kinberg, 1866 (Polychaeta, Nereididae) from Ecuador, Eastern Pacific, with a key to all known species of the genus. Zookeys 269, 67–76.

Dean, H.K. (2001) Some Nereididae (Annelida: Polychaeta) from the Pacific Coast of Costa Rica. Revista de Biología Tropical 40, 37–67.

Dean, H.K. (2012) A literature review of the Polychaeta of the Caribbean Sea. Zootaxa 3596, 1– 86. delle Chiaje, S. (1822) Memorie sulla Storia e Notomia degli Animali senza Vertebre del Regno di Napoli. Publisher not indicated, Figures, 109 plates.

Dueñas-Ramírez, P.R. & Quiros-Rodríguez, J.A. (2012) Ocurrence of Pseudonereis gallapagensis Kingberg, 1865 (Annelida: Polychaeta) in the Colombian Caribbean Coast. Revista Colombiana de Ciencia 4, 454–457.

Ehlers, E. (1868) Die Borstenwürmer (Annelida Chaetopoda) Nach Systematischen und Anatomischen Untersuchungen. Erster Band. Wilhelm Engelmann, Leipzig.

Ehlers, E. (1887) Reports on the results of dredging, under the direction of L. F. Pourtalès, during 226

the years 1868-1870, and the Alexander Agassiz, in the Gulf of Mexico (1877-78) and in the Caribbean Sea (1878-79), in the U.S. Coast Survey Steamer “Blake”. XXI. Report on the annelids. Memoirs of the Museum of Comparative Zoology at Harvard College 15, 1–335.

Ehlers, E. (1901) Die Polychaeten des magellanischen und chilenischen Strandes. Ein faunistischer Versuch. Festschrift zur Feier des Hundertfünfzigjährigen Bestehens des Königlichen Gesellschaft der Wissenschaften zu Göttingen. Abhandlungen der Mathematisch-Physikalischen Klasse, 1–232.

Einfeldt, A.L., Doucet, J.R. & Addison, J.A. (2014) Phylogeography and cryptic introduction of the ragworm Hediste diversicolor (Annelida, Nereididae) in the Northwest Atlantic. Invertebrate Biology 133, 232–241.

Eisig, H. (1870) Nereis hircinicola. (Nova Species.). Zeitschrift Für Wissenschaftliche Zoologie 20, 103–105.

Ezhova, E.E. (2011) Spawning and Early Ontogenesis of the Littoral Polychaete Namanereis littoralis (Grube , 1876) (Nereididae , Namanereidinae). Russian Journal of Developmental Biology 42, 160–167.

Fauchald, K. (1972) Benthic polychaetous annelids from deep waters off Western Mexico and adjacent areas in the Eastern Pacific Ocean. Allan Hancock Monographs in Marine Biology 7, 1–575.

Fauchald, K. (1977a) Polychaetes from Intertidal Areas in Panama, with a Review of Previous Shallow-Water Records. Smithsonian Contributions to Zoology 221, 1–81.

Fauchald, K. (1977b) Some nereid polychaetes from deep waters. In: D. J. Reish and K. Fauchald (Eds), Essays on the polychaetous annelids in memory of Dr. Olga Hartman. Allan Hancock Foundation, Los Angeles, pp. 157–171.

Fauchald, K. (1977c) 28 Natural History Museum of Los Angeles County, Science Series The polychaete worms: Definitions and keys to the orders, families and genera.

Fauvel, P. (1914) Annélides polychètes de San-Thomé (Golfe de Guinée) recueillies par M. Ch. Gravier. Archives de Zoologie Expérimentale et Générale 54, 105–155.

Fauvel, P. (1919) Annélides polychètes de la Guyane francaise. Bulletin du Muséum National

227

d’Histoire Naturelle 25, 472–479.

Fauvel, P. (1921) Annélides polychètes de Madagascar, du Muséum R. d’Histoire Naturelle recueillies par M. le Dr. W. Kaudern en 1912. Arkiv för Zoologi 13, 1–32.

Fauvel, P. (1923a) Annélides polychètes des Illes Gambier et de la Guyane. Memorie della Pontificia Accademia Romana dei Nuovi Lincei, Series II 6, 1–59.

Fauvel, P. (1923b) Polychètes Errantes. Faune de France 5, 1–488.

Fauvel, P. (1932) Annelida Polychaeta of the Indian Museum, Calcutta. Memoirs of the Indian Museum 12, 1–261.

Fauvel, P. (1953) Annélides Polychètes de la Croisière du “Président Théodore Tissier” aux Antilles (1951). Bulletin de l’Institut Océanographique 1033, 1–23.

Feuerborn, H.J. (1931) Eine Rhizocephale und zwei Polychaeten aus dem Süßwasser von Java und Sumatra. Verhandlungen der Internationalen Vereinigung für theoretische und angewandte Limnologie, Stuttgart 5, 618–660.

Fischer, A.H.L., Henrich, T. & Arendt, D. (2010) The normal development of Platynereis dumerilii. Frontiers in Zoology 7, 1–39.

Fitzhugh, K. & Wolf, P.S. (1990) Gross Morphology of the Brain of Pilargid Polychaetes: Taxonomic and Systematic Implications. American Museum Novitates 2992, 1–16.

Gardiner, S.L. (1975) Errant polychaete annelids from North Carolina. Journal of the Elisha Mitchell Scientific Society 91, 77–220.

Gillet, P. (1986) Contribution à l’étude des Annélides Polychètes des lagunes de la Manche-à- Eau et de Belle-Plaine (Guadeloupe) Description d'un nouveau Capitellidae : Scyphoproctus guadalupensis n. sp. Bulletin du Muséum National d’Histoire Naturelle, Paris, 4 sér., 8, section A 4, 803–817.

Glasby, C.J. (1991) Phylogenetic Relationships in the Nereididae (Annelida: Polychaeta), Chiefly in the Subfamily Gymnonereidinae, and the Monophyly of the Namanereidinae. Bulletin of Marine Science 48, 559–573.

Glasby, C.J. (1999) The Namanereidinae (Polychaeta: Nereididae). Part 1, Taxonomy and

228

Phylogeny. Records of the Australian Museum, Supplement 25, 1–129.

Glasby, C.J., Fauchald, K. & Hutchings, P. (2000) Glossary. In: P. L. Beesley, G. J. B. Ross, and C. J. Glasby (Eds), Polychaetes and Allies: The southern synthesis. Fauna of Australia. Vol. 4A Polychaeta, Myzostomida, Pogonophora, Echiura, Sipuncula. CSIRO Publishing, Melbourne, pp. 401–412.

Glasby, C.J. & Hsieh, H.-L. (2006) New Species and New Records of the Perinereis nuntia Species Group (Nareididae: Polychaeta) from Taiwan and Other Indo-West Pacific Shores. Zoological Studies 45, 553–577.

Glasby, C.J., Miura, T., Nishi, E. & Junardi (2007) A new species of Namalycastis (Polychaeta: Nereididae: Namanereidinae) from the shores of South-east Asia. The Beagle, Records of the Museums and Art Galleries of the Northern Territory 23, 21–27.

Glasby, C.J., Timm, T. & Muir, A.I. (2009) Catalogue of non-marine Polychaeta (Annelida) of the World. Zootaxa 2070, 1–52.

Glasby, C.J., Wei, N.-W. V. & Gibb, K.S. (2013) Cryptic species of Nereididae (Annelida : Polychaeta) on Australian coral reefs. Invertebrate Systematics 27, 245–264.

Glasby, C.J., Wilson, R.S. & Bakken, T. (2011) Redescription of the Indo-Pacific polychaete Neanthes pachychaeta (Fauvel, 1918) n. comb. (Annelida, Phyllodocida, Nereididae) and its synonyms. Zoosystema 33, 361–375.

Gobin, J.F. (1990) A checklist of marine polychaetous annelids (Polychaeta) for the Gulf of Paria, Trinidad, West Indies. Caribbean Marine Studies 1, 37–47.

González-Escalante, L.E. & Salazar-Vallejo, S.I. (2003) A new estuarine species, Nereis garwoodi (Polychaeta: Nereididae), from Bahía Chetumal, Mexican Caribbean coast. Revista de Biología Tropical 51, 155–164.

Gravier, M.C. (1901) Sur deux nouvelles espèces du genre Lycastis Savigny, Aud. et Edw. rev., de la Guyane française. Bulletin du Muséum National d’Histoire Naturelle 7, 397–402.

Gravier, M.C. (1902) Sur le genre Lycastis Savigny (Audouin et M.-Edwards rev.). Bulletin de la Société d’Histoire naturelle d'Autun 14, 373–379.

Gravier, M.C. (1909) Annélides polychètes recueillis a Payta (Pérou) par M. le Dr Rivet, membre 229

de la Mission géodésique de l’Équateur. Archives de Zoologie Expérimentale et Générale, 4e Série 10, 617–659.

Grube, A.E. (1858) Annulata örstediana. Enumeratio Annulatorum, quae in itinere per Indiam occidentalem et Americam centralem annis 1845-1848 suscepto legit cl. A.S. Örsted, adjectis speciebus nonnullis a cl. H. Kröyero in itinere ad Americam meridionalem collectis, (Fortsættelse [continued]) [2. Familia Euniceae - F. Syllidea]. Videnskabelige Meddelelser fra Dansk Naturhistorisk Forening for 1857, 158–186.

Grube, A.E. (1859) Annulata örstediana. Enumeratio Annulatorum, quae in itinere per Indiam occidentalem et Americam centralem annis 1845-1848 suscepto legit cl. A.S. Örsted, adjectis speciebus nonnullis a cl. H. Kröyero in itinere ad Americam meridionalem collectis, (Slutning [conclusion]) [3. Familia Ariciaea - F. Sipunculacea and Index]. Videnskabelige Meddelelser fra Dansk Naturhistorisk Forening for 1858, 105–120.

Grube, A.E. (1872) Über die Gattung Lycastis und ein paar neue Arten derselben. Jahres-Bericht der Schlesischen Gesellschaft für Vaterländische Cultur 49, 47–48.

Grube, A.E. (1873) Die familie der Lycorideen und die aufstellung von gruppen in der gattung Nereis. Jahres-Bericht der Schlesischen Gesellschaft für Vaterländische Cultur 51, 56–73.

Grube, A.E. (1878) Annulata Semperiana. Beitrage zur Kenntnis der Annelidenfauna der Philippinen nach den von Herrn Prof. Semper mitgebrachten sammlungen. Memoires de L’Académie Impériale des Sciences de St. Pétersbourg, 7 sér. 25, 1–300.

Haddad, N. El, Azzati, R.C. & García-Carrascosa, A.M. (2013) Annotated checklist of Capitellidae (Annelida, Polychaeta) from the Iberian Peninsula, Chafarinas, Balearic and Canary Islands. Graellsia 69, 97–116.

Hansen, A. (1882) Recherches sur les Annélides Recueillies par M. le professeur Édouard Van Beneden pendant son Voyage au Brésil et à La Plata. Memoires Couronnes et Memoires des Savants Etrangers publies par L’Academie Royale des Sciences, des Lettres et des Beaux- Arts de Belgique 44, 1–29.

Harper, D.E. (1979) Nereis (Neanthes) micromma n. sp. (Polychaeta: Nereididae) from the Northern Gulf of Mexico with a note on the structure of nereidid palps. Contributions in Marine Science 22, 91–103. 230

Hartman, O. (1938a) Annotated list of the types of polychaetous annelids in the Museum of Comparative Zoology. Bulletin of the Museum of Comparative Zoology at Harvard College 85, 1–31.

Hartman, O. (1938b) Nomenclatorial changes involving types of polychaetous annelids of the Family Nereidae in the United States National Museum. Journal of the Washington Academy of Sciences 28, 13–15.

Hartman, O. (1940) Polychaetous annelids. Part II. Chrysopetalidae to Goniadidae. Allan Hancock Pacific Expeditions 7, 173–287.

Hartman, O. (1944) Polychaetous Annelids. Allan Hancock Atlantic Expedition 3, 1–33.

Hartman, O. (1945) The marine annelids of North Carolina. Duke University Marine Station Bulletin 2, 1–51.

Hartman, O. (1948) The marine annelids erected by Kinberg with notes on some other types in the Swedish State Museum. Arkiv för Zoologi 42A, 1–137.

Hartman, O. (1954) Polychaetous annelids of the Gulf of Mexico. In: P. S. Galtsoff (Ed), Gulf of Mexico. Its Origin, Waters, and Marine Life. Fishery Bulletin of the Fish and Wildlife Service. Volume 55., pp. 413–417.

Hartman, O. (1956) Polychaetous annelids erected by Treadwell, 1891 to 1948, together with a brief cronology. Bulletin of the American Museum of Natural History 109, 239–310.

Hartman, O. (1958) A new nereid worm from Warm Mineral Springs, Fla., with a review of the genus Nicon Kinberg. Journal of the Washington Academy of Sciences 48, 263–266.

Hartman, O. (1959a) Capitellidae and Nereidae (Marine Annelids) from the Gulf side of Florida, with a review of freshwater Nereidae. Bulletin of Marine Science 9, 153–168.

Hartman, O. (1959b) Catalogue of the polychaetous annelids of the world. Part I. Allan Hancock Foundation Publications. Occasional Paper. 23, 1–353.

Hartman, O. (1964) Polychaeta Errantia of Antarctica. Antarctic Research Series 3, 1–287.

Hartman, O. (1965a) Catalogue of the polychaetous annelids of the world. Supplement 1960- 1965 and index. Allan Hancock Foundation Publications. Occasional Paper. 23, 1–197.

231

Hartman, O. (1965b) Deep-water benthic polychaetous annelids off New England to Bermuda and other North Atlantic areas. Allan Hancock Foundation Publications. Occasional Paper. 28, 1–378.

Hartmann-Schröder, G. (1958) Einige Polychaten aus dem Küstengrundwasser der Bimini-Inseln (Bahamas). Kieler Meeresforschungen 14, 233–240.

Hartmann-Schröder, G. (1977a) Die Polychaeten der Kubanisch-Rumänischen Biospeologischen Expedition nach Kuba 1973. Résultats des expéditions biospéologiques cubano-roumaines à Cuba 2, 51–63.

Hartmann-Schröder, G. (1977b) Profundilycastis profundus (Hartman, 1965) (Nereidae, Polychaeta): Wiederbeschreibung des Holotypus. Mitteilungen aus dem Hamburgischen Zoologischen Museum und Institut 74, 45–48.

Hartmann-Schröder, G. (1977c) The genera Ceratocephale Malmgren, Olganereis n. gen., and Profundilycastis n. gen. (Nereidae, Polychaeta) with a key to the nereid genera without chitinous paragnaths. In: D. J. Reish and K. Fauchald (Eds), Essays on the polychaetous annelids in memory of Dr. Olga Hartman. Allan Hancock Foundation, University of South California, Los Angeles, California, pp. 141–155.

Hartmann-Schröder, G. (1980) Die polychaeten der Amsterdam-Expeditionen nach Westindien. Bijdragen tot de Dierkunde 50, 387–401.

Hartmann-Schröder, G. (1985) Revision der gattung Ceratonereis Kinberg (Nereididae, Polychaeta) (Mit besonderer Berücksichtigung der Arten mit eigenschnittenem Prostomium). Mitteilungen aus dem Hamburgischen Zoologischen Museum und Institut 82, 37–59.

Hernández-Alcántara, P. & Solís-Weiss, V. (1991) New records of errantiate polychaetous annelids from the Gulf of California. Bulletin of Marine Science 48, 251–260.

Hoagland, R.A. (1919) Polychaetous Annelids from Porto Rico, The Florida Keys, and Bermuda. Bulletin of the American Museum of Natural History 41, 571–591.

Hong, J.-S., Choi, B.-M., Kubo, A. & Sato, M. (2012) Redescription of the giant mud worm Paraleonnates uschakovi Khlebovich and Wu, 1962 (Polychaeta: Nereididae) with special

232

reference to the synonymy of Periserrula leucophryna Paik, 1977 and the difference from Paraleonnates bolus (Hutchings and Reid, 1991). Zootaxa 3490, 49–62.

Horst, R. (1918) On a species of Lycastis and three aberrant forms of Nereidae from the dutch East Indies. Zoologische Mededeelingen 4, 246–250.

Hutchings, P. & Glasby, C.J. (1985) Additional nereidids (Polychaeta) from Eastern Australia, together with a redescription of Namanereis quadraticeps (Gay) and the synonymising of Ceratonereis pseudoerythraeensis Hutchings & Turvey with C. aequisetis. Records of the Australian Museum 37, 101–110.

Hutchings, P. & Reid, A. (1990) The Nereididae (Polychaeta) from Australia - Gymnonereidinae sensu Fitzhugh, 1987: Australonereis, Ceratocephale, Dendronereides, Gymnonereis, Nicon, Olganereis and Websterinereis. Records of the Australian Museum 42, 69–100.

Hutchings, P., Reid, A. & Wilson, R.S. (1991) Perinereis (Polychaeta, Nereididae) from Australia, with redescriptions of six additional species. Records of the Australian Museum 43, 241–274.

Hylleberg, J. & Nateewathana, A. (1988) Polychaetes of Thailand. Nereididae (Part 2): Ceratocephale and Gymnonereis, with description of two new species and notes of the subfamily Gymnonereidinae. Phuket Marine Biological Center Research Bulletin 49, 1–20.

Hylleberg, J., Nateewathana, A. & Bussarawit, S. (1986) Polychaetes from Thailand, Nereidae (Part 1): Perinereis and Pseudonereis with notes on species of commercial value. Phuket Marine Biological Center Research Bulletin 43, 1–22.

ICZN (1999) International Code of Zoological Nomenclature, 4th ed. International Trust for Zoological Nomenclature (The Natural History Museum), London.

Imajima, M. (2003) Polychaetous Annelids from Sagami Bay and Sagami Sea collected by the Emperor Showa of Japan and deposited at the Showa Memorial Institute, National Science Museum, Tokyo (II) : Orders included within the Phyllodocida, Amphinomida, Spintherida and Eunicid. National Science Museum Monographs 23, 1–221.

233

Khlebovich, V.V. (1996) Fauna of Russia and Neighbouring Countries. Polychaetous Annelids, Volume III. Polychaetes of the Family Nereididae of the Russian Seas and the Adjacent Waters. NAUKA publishing house, St. Petersburg, 221 pp. Kinberg, J.G.H. (1865) Annulata Nova (Nereidum Dispositio Nova. Leonnatidea, Nereidea, Aretidea, Pisenoidea, Niconidea). Öfversigt af Kongelige Vetenskaps-Akademiens Förhandlingar 22, 167–179.

Klesch, W.L. (1970) The reproductive biology and larval development of Laeonereis culveri Webster (Polychaeta: Nereidae). Contributions in Marine Science 15, 71–85.

Kulakova, M., Bakalenko, N., Novikova, E., Cook, C.E., Eliseeva, E., Steinmetz, P.R.H., Kostyuchenko, R.P., Dondua, A., Arendt, D., Akam, M. & Adreeva, T. (2007) Hox gene expression in larval development of the polychaetes Nereis virens and Platynereis dumerilii (Annelida, Lophotrochozoa). Development Genes and Evolution 217, 39–54.

Langerhans, P. (1880) Die Wurmfauna Madeiras. II. Zeitschrift Für Wissenschaftliche Zoologie 33, 271–316.

Leuckart, R.S. (1847) Verzeichnis der zur Fauna Helgoland’s gehørenden wirbellosen Seethiere. In: H. Frey and R. S. Leuckart (Eds), Beiträge zur Kenntnis Wirbelloser Thiere mit Besonderer Berücksichtigung der Fauna des Norddeutschen Meeres. Friedrich Vieweg und sohn, Braunschweig, pp. 136–168.

Linnaeus, C. (1758) Systema Naturae Per Regna Tria Naturae, Secundum Classes, Ordines, Genera, Species, Cum Characteribus, Differentiis, Synonymis, Locis. 10th ed. Holmiae: Laurentii Salvii.

Liñero-Arana, I. (1983) Dos Nuevas Especies de Nereidae (Polychaeta: Errantia) de la Costa Oriental de Venezuela. Boletín del Instituto Oceanográfico de Venezuela 22, 3–6.

Liñero-Arana, I. & Díaz-Díaz, Ó. (2007) Nuevas adiciones de Nereididae (Annelida: Polychaeta) para las costas de Venezuela. Boletín del Instituto Oceanográfico de Venezuela 46, 149– 159.

Liñero-Arana, I. & Reyes-Vásquez, G. (1979) Nereidae (Polychaeta, Errantia) del Golfo de Cariaco, Venezuela. Boletín del Instituto Oceanográfico de la Universidad de Oriente 18, 3–12. 234

Magesh, M., Kvist, S. & Glasby, C.J. (2014) Incipient speciation within the Namalycastis abiuma (Annelida: Nereididae) species group from southern India revealed by combined morphological and molecular data. Memoirs of Museum Victoria 71, 169–176.

Malmgren, A.J. (1867) Annulata Polychaeta: Spetsbergiae, Groenlandiae, Islandiae et Scandinaviae. Hactenus Cognita. Öfversigt af Kongelige Vetenskaps-Akademiens Förhandlingar 4, 127–235.

Mazurkiewicz, M. (1975) Larval development and habits of Laeonereis culveri (Webster) (Polychaeta: Nereididae). Biological Bulletin 149, 186–204.

McIntosh, W.C. (1885) Report on the Annelida Polychæta dredged by H.M.S. Challenger during the Years 1873-1876. Report of the Scientific Results of the Voyage of H.M.S. Challenger During the Years 1873-76. Zoology Part XXXIV. 12, 1–554.

McIntosh, W.C. (1910) A Monograph of the British Annelids. Vol. II. Part II. Polychaeta. Syllidae to Ariciidae. The Ray Society of London 2(2), 233–524.

Mikkelsen, P.S. & Virnstein, R.W. (1982) An illustrated glossary of polychaete terms. Harbor Branch Foundation. Technical Report. 46, 1–92.

Miloslavich, P., Díaz, J.M., Klein, E., Alvarado, J.J., Díaz, C., Gobin, J.F., Escobar-Briones, E., Cruz-Motta, J.J., Weil, E., Cortés, J., Bastidas, A.C., Robertson, R., Zapata, F., Martín, A., Castillo, J., Kazandjian, A. & Ortíz, M. (2010) Marine biodiversity in the Caribbean : regional estimates and distribution patterns. Plos One 5, 1–25.

Monro, C.C.A. (1933a) The polychaeta Errantia collected by Dr. Crossland at Colón, in the Panama Region, and the Galapagos Islands during the Expedition of the S.Y. ‘St. George.’ Proceedings of the Zoological Society of London 1933, 1–96.

Monro, C.C.A. (1933b) XXVII. On a collection of polychaeta from Dry Tortugas, Florida. Annals and Magazine of Natural History, 10 ser. 12, 244–269.

Monro, C.C.A. (1939) Polychaeta of the “Rosaura” Expedition. Novitates Zoologicae 41, 345– 354.

Moore, J.P. (1903) Descriptions of two new species of Polychaeta from Wood’s Hole, Massachussetts. Proceedings of the Academy of Natural Sciences of Philadelphia 55, 720–

235

726.

Müller, O.F. (1776) Zoologiae Danicae Prodromus, seu Animalium daniae et Norvegiae Indigenarum Characters, Nomina, et Synonyma Imprimis Popularium. Havniae.

Núñez, J. & Brito, M. del C. (2006) A new species of Rullierinereis and new records of Nereididae from sublittoral sandy bottoms off Lanzarote (Canary Islands). Scientia Marina 70, 145–150.

Orenzanz, J.M. & Gianuca, N.M. (1974) Contribuiçao ao conhecimento dos anelídos poliquetas do Rio Grande do sul, Brasil. I. lista sistematica preliminar e descriçao de très novas espécies. Comunicaçoes do Museo de Ciências da PUCRGS Porto Alegre 4, 1–37.

Paxton, H. (1983) Revision of the genus Micronereis (Polychaeta: Nereididae: Notophycinae). Records of the Australian Museum 35, 1–18.

Paxton, H. & Åkesson, B. (2007) Redescription of Ophryotrocha puerilis and O. labronica (Annelida, Dorvilleidae). Marine Biology Research 3, 3–19.

Perkins, T.H. (1980) Review of species previously referred to Ceratonereis mirabilis, and descriptions of new species of Ceratonereis, Nephtys, and Goniada (Polychaeta). Proceedings of the Biological Society of Washington 93, 1–49.

Perkins, T.H. & Savage, T. (1975) A bibliography and checklist of polychaetous anelids of Florida, the Gulf of Mexico, and the Caribbean Region. Florida Marine Research Publications 14, 1–62.

Pesch, C.G., Pesch, C.E. & Mueller, C. (1988) Chromosome complements from two populations of the marine worm Neanthes arenaceodentata (Annelida: Polychaeta). Ophelia 28, 163– 167.

Pettibone, M.H. (1956) Some polychaete worms of the families Hesionidae, Syllidae, and Nereidae from the east coast of North America, West Indies, and Gulf of Mexico. Journal of the Washington Academy of Sciences 46, 281–294.

Pettibone, M.H. (1963) Marine polychaete worms of the New England Region. I. Families Aphroditidae through Trochochaetidae. Bulletin of the United States National Museum 227, 1–356.

236

Pettibone, M.H. (1971) Revision of some species referred to Leptonereis, Nicon, and Laeonereis (Polychaeta: Nereididae). Smithsonian Contributions to Zoology 104, 1–53.

Pettibone, M.H. (1976) Revision of the genus Macellicephala McIntosh and the subfamily Macellicephalinae Hartmann-Schröder (Polychaeta: Polynoidae). Smithsonian Contributions to Zoology 229, 1–71.

Pettibone, M.H. (1989) Revision of the aphroditoid polychaetes of the Family Acoetidae Kinberg (= Polyodontidae Augener) and reestablishment of Acoetes Audouin and Milne-Edwards, 1832, and Euarche Ehlers, 1887. Smithsonian Contributions to Zoology 464, 1–138.

Pflugfelder, O. von (1933) Landpolychäten aus Niederländisch-Indien. (Ergebnisse der Sunda- Expedition der Notgemeinschaft der Deutschen Wissenschaft 1929/30.). Zoologischer Anzeiger 105, 65–75.

Pleijel, F. (1998) Phylogeny and classification of Hesionidae (Polychaeta). Zoologica Scripta 27, 89–163.

Pleijel, F. & Gustavsson, L. (2010) Chambered chaetae in nereidiform polychaetes (Annelida). Zoomorphology 129, 93–98.

Prevedelli, D. & Cassai, C. (2001) Reproduction and larval development of Perinereis rullieri Pilato in the Mediterranean Sea (Polychaeta: Nereididae). Ophelia 54, 133–142.

Qiu, J.-W. & Qian, P.-Y. (2000) Revision of the genus Leonnates Kinberg, 1866 (Polychaeta: Nereididae), with descriptions and comments on other species described in Leonnates. Proceedings of the Biological Society of Washington 113, 1111–1146. de Quatrefages, A. (1865) Histoire Naturelle des Annelés Marins et d’Eau Douce: Annélides et Géphyriens. Tome Premier. Collection des Suites a Buffon formant avec les Oeuvres de cet auteur un Cours Complet d’Histoire Naturelle. Librairie Encyclopédique de Roret, Paris.

Ramírez-Hernández, A., Hernández-Alcántara, P. & Solís-Weiss, V. (2015) Nereis alacranensis, a new species of polychaete (Annelida, Nereididae) from Alacranes Reef, southern Gulf of Mexico, with a key to Nereis from the Grand Caribbean. Zootaxa 4012, 156–166.

Read, G.B. (2007) Taxonomy of sympatric New Zealand species of Platynereis, with description of three new species additional to P. australis (Schmarda). Zootaxa 1558, 1–28.

237

Reish, D.J. (1957) The Life History of the Polychaetous Neanthes caudata (delle Chiaje), Including a Summary of Development in the Family Nereidae. Pacific Science 11, 216–228.

Reish, D.J., Anderson, F.E., Horn, K.M. & Hardege, J. (2014) Molecular phylogenetics of the Neanthes acuminata (Annelida: Nereididae) species complex. Memoirs of Museum Victoria 71, 271–278.

Renaud, J.C. (1956) A report of some polychaetous annelids from the Miami-Bimini area. American Museum Novitates 1812, 1–40.

Reuscher, M.G. & Shirley, T.C. (2014) Diversity, distribution, and zoogeography of benthic polychaetes in the Gulf of Mexico. Marine Biodiversity 44, 519–532.

Rouse, G.W. & Pleijel, F. (2001) Polychaete anatomy. In: G. W. Rouse and F. Pleijel (Eds), Polychaetes. Oxford University Press, New York, pp. 18–35.

Rozbaczylo, N. & Bolados, J. (1980) Nereidos de Iquique, Chile (Polychaeta: Nereidae). Boletín del Museo Nacional de Historia Natural, Chile 37, 205–224.

Rullier, F. & Amoureux, L. (1979) Campagne de la Calypso au large des côtes atlantiques de ’Amérique du sud (1961-1962). I. 33. Annélides Polychètes. Résultats Scientifiques des Campagnes de la Calypso, 11, Annales de l’Institut Océanographique, Suppl. 55, 10–206.

Ruta, C., Nygren, A., Rousset, V., Sundberg, P., Tillier, A., Wiklund, H. & Pleijel, F. (2007) Phylogeny of Hesionidae (Aciculata, Polychaeta), assessed from morphology, 18S rDNA, 28S rDNA, 16S rDNA and COI. Zoologica Scripta 36, 99–107. de Saint-Joseph, M. (1901) Sur quelques invertébrés marins des cotes du Sénégal (Annélides polychètes, nématoide endoparasite d’annélide polychète et crustacé decapode parasite). Annales des Sciences Naturelles (Zoologie), Série 8 12, 217–248.

Salazar-Vallejo, S.I. (1989) Enrique Rioja y su contribución al estudio de los poliquetos (Anelida: Polychaeta) en México. Brenesia 30, 39–65.

Salazar-Vallejo, S.I. (1996) Lista de especies y bibliografía de poliquetos (Polychaeta) del Gran Caribe. Anales del Instituto de Biología de la Universidad Nacional Autónoma de México, Serie Zoología 67, 11–50.

Salazar-Vallejo, S.I. (2000) Biogeografía marina del Gran Caribe. Interciencia 25, 7–12. 238

Salazar-Vallejo, S.I. & Eibye-Jacobsen, D. (2012) Annulata örstediana: publication dates, composition and annotated taxonomic list, with some comments on Hemipodus (Polychaeta: Glyceridae). Revista de Biología Tropical 60, 1391–1402.

Salazar-Vallejo, S.I. & Jiménez-Cueto, M.S. (1997) Neréididos (Polychaeta) del Caribe mexicano con una clave para las especies del Gran Caribe. Revista de Biología Tropical 44, 361–377.

Salazar-Vallejo, S.I. & Londoño-Mesa, M.H. (2004) Lista de especies y bibliografía de poliquetos (Polychaeta) del Pacífico Oriental Tropical. Anales del Instituto de Biología, Universidad nacional Autónoma de México, Serie Zoología 75, 9–97.

Santos, C.S.G. (2007) Nereididae from Rocas Atoll (North-East, Brazil). Arquivos do Museu Nacional do Rio de Janeiro 65, 369–380.

Santos, C.S.G. & Lana, P. da C. (2001) Nereididae (Annelida, Polychaeta) da Costa Nordeste do Brasil. II. Gêneros Namalycastis, Ceratocephale, Laeonereis e Rullierinereis. Iheringia, Série Zoologia 91, 137–149.

Santos, C.S.G. & Lana, P. da C. (2003) Nereididae (Polychaeta) Da Costa Nordeste Do Brasil. III. Gêneros Ceratonereis e Nereis. Iheringia, Série Zoologia 93, 5–22.

Santos, C.S.G., Pleijel, F., Lana, P. da C. & Rouse, G.W. (2005) Phylogenetic relationships within Nereididae (Annelida: Polychaeta). Invertebrate Systematics 19, 557–576.

Sato, M. & Nakashima, A. (2003) A review of Asian Hediste species complex (Nereididae, Polychaeta) with descriptions of two new species and a redescription of Hediste japonica (Izuka, 1908). Zoological Journal of the Linnean Society 137, 403–445.

Schmarda, L.K. (1861) Neue wirbellose thiere beobachtet und gesammelt auf einer reise um die erde 1853 bis 1857. Erster Band. Turbellarian, Rotatorien un Anneliden. Zweite Hälfte. Wilhelm Engelmann, Leipzig.

Solís-Weiss, V. & Espinasa, L. (1991) Lycastilla cavernicola, a new freshwater nereidid from an inland mexican cave (Polychaeta: Nereididae: Namanereidinae). Proceedings of the Biological Society of Washington 104, 631–639.

Spalding, M.D., Fox, H.E., Allen, G.R., Davidson, N., Ferdaña, Z.A., Finlayson, M., Halpern,

239

B.S., Jorge, M.A., Lombana, A., Lourie, S.A., Martin, K.D., McManus, E., Molnar, J., Recchia, C.A. & Robertson, J. (2007) Marine Ecoregions of the World: A Bioregionalization of Coastal and Shelf Areas. BioScience 57, 573–583.

Steinmetz, P.R.H., Kostyuchenko, R.P., Fischer, A. & Arendt, D. (2011) The segmental pattern of otx, gbx, and Hox genes in the annelid Platynereis dumerilii. Evolution & Development 13, 72–79.

Suárez, A.M. & Fraga, R. (1978) Poliquetos bentósicos cubanos I: Lista de poliquetos errantes. Investigaciones Marinas (Havana) 33, 1–60.

Taylor, J.L. (1984) Family Nereidae Johnston, 1845. In: J. M. Uebelacker and P. G. Johnson (Eds), Taxonomic guide to the polychaetes of the Northern Gulf of Mexico. Final report to the Minerals Management Service, contract 14-12-001-29091. Barry A. Vittor & Associates, Inc., Mobile, Alabama, pp. 31 1–42.

Treadwell, A.L. (1901) The polychaetous annelids of Porto Rico. Bulletin of the United States National Fish Commission 20, 183–210.

Treadwell, A.L. (1924) Polychaetous annelids collected by the Barbados-Antigua Expedition from the University of Iowa in 1918. University of Iowa Studies in Natural History 10, 3– 23.

Treadwell, A.L. (1926) A new polychaetous annelid from Kartabo, British Guiana. Zoologia 7, 101–104.

Treadwell, A.L. (1928) Polychaetous annelids from the Arcturus oceanographic expedition. Zoologica 8, 449–485.

Treadwell, A.L. (1929) New species of polychaetous annelids in the Collections of the American Museum of Natural History, from Porto Rico, Florida, Lower California, and British Somaliland. American Museum Novitates 392, 1–13.

Treadwell, A.L. (1931) New species of polychaetous annelids from California, Mexico, Porto Rico, and Jamaica. American Museum Novitates 482, 1–7.

Treadwell, A.L. (1939) Polychaetous annelids of Porto Rico and vicinity. In: Scientific Survey of Porto Rico And the virgin Islands. Volumen 16, pt. 2. New York Academy of Sciences, New

240

York, pp. 151–118.

Treadwell, A.L. (1940) A new genus and two new species of polychaetous annelids from Texas and one new species from the Philippine Islands. American Museum Novitates 1809, 1–4.

Treadwell, A.L. (1941) New species of polychaetous annelids from the vicinity of Galveston, Texas. American Museum Novitates 1139, 1–3.

Villalobos-Guerrero, T.F. & Carrera-Parra, L.F. (2015) Redescription of Alitta succinea (Leuckart, 1847) and reinstatement of A. acutifolia (Ehlers, 1901) n. comb. based upon morphological and molecular data (Polychaeta: Nereididae). Zootaxa 3919, 157–178.

Villalobos-Guerrero, T.F. & Tovar-Hernández, M.A. (2013) Una especie nueva de Pseudonereis (Polychaeta: Nereididae) de Mazatlán, golfo de California, incluyendo una clave para las especies del mundo. Revista Mexicana de Biodiversidad 84, 774–781.

Virgilio, M., Fauvelot, C., Costantini, F., Abbiati, M. & Backeljau, T. (2009) Phylogeography of the common ragworm Hediste diversicolor (Polychaeta: Nereididae) reveals cryptic diversity and multiple colonization events across its distribution. Molecular Ecology 18, 1980–1994.

Warren, W.S. (1942) A Survey of the annelid worms of the Grand Isle Region. Bios 13, 39–46.

Webster, H.E. (1879a) Annelida Chaetopoda of New Jersey. Annual Reports on the New York State Museum of Natural History 32, 101–128.

Webster, H.E. (1879b) Annelida Chaetopoda of the Virginian Coast. Transactions of the Albany Institute 9, 202–269.

Webster, H.E. (1884) Annelida from Bermuda, collected by G. Brown Goode. Bulletin of the United States National Museum 25, 305–327.

Weinberg, J.R., Starczak, V.R., Mueller, C., Pesch, G.C. & Lindsay, S.M. (1990) Divergence between populations of a monogamous polychaete with male parental care: premating isolation and chromosome variation. Marine Biology 107, 205–213.

Wesenberg-Lund, E. (1958) Lesser Antillean Polychaetes, chiefly from brackish water with a Survey and a Bibliography of Fresh and Brackish – Water Polychaetes. Studies on the Fauna of Curaçao and other Caribbean Islands 8, 1–41. 241

Willey, A. (1905) Report on the Polychaeta Collected by Professor Herdman, at Ceylon, in 1902. Report to the Government of Ceylon on the Pearl Oyster Fisheries of the Gulf of Manaar, With Supplementary Reports Upon the Marine Biology of Ceylon, by Other Naturalists. Part IV. Supplementary Reports 4, 243–324.

Williams, D.D. (2004) Review of the polychaete genus Namanereis (Nereididae) in the Caribbean Region, with a Record of N. hummelincki from Deep Freshwater Wells on Barbados. Caribbean Journal of Science 40, 401–407.

Wilson, D.P. (1932) The development of Nereis pelagica Linnæus. Journal of the Marine Biological Association of the United Kingdom 18, 203–217.

Wilson, R.S. & Glasby, C.J. (1993) A revision of the Perinereis nuntia species group (Polychaeta: Nereididae). Records of the Australian Museum 45, 253–277.

Winchell, C.J., Valencia, J.E. & Jacobs, D.K. (2010) Confocal analysis of nervous system architecture in direct-developing juveniles of Neanthes arenaceodentata (Annelida, Nereididae). Frontiers in Zoology 7, 1–19.

Yáñez-Rivera, B. & Carrera-Parra, L.F. (2012) Reestablishment of Notopygos megalops McIntosh, description of N. caribea sp. n. from the Greater Caribbean and barcoding of “amphiamerican” Notopygos species (Annelida, Amphinomidae). Zookeys 223, 69–84.

242

Figures

Figure 1. Drawings from parapodia and parapodial structures showing terminology used in the present study. A–E. Parapodial types used in text, sorted from A to E, respectively. F. Neuropodial lobe, distal view, showing the general position of postchaetal lobes. G. Neuropodial ramus (Vc omitted) from a chaetiger of Paraleonnates species in posterior (left) and anterior (right) views, showing the arising of both superior and inferior lobes from the same structure (AcLo). H–K. Schemes of neuropodial lobes, seen from above, showing different developments of neuropodial postchaetal lobes. H. PosLo shorter than AcLo. I. As long as it. J and K. longer than it. L. Scheme of a parapodium, showing the points and segments used. M. Scheme of a dorsal ligule, showing the procedure to obtain the reach of dorsal cirrus. N. Scheme of a dorsal ligule, showing the procedure to obtain the insertion of dorsal cirrus. Abbreviations. Dc, Dorsal cirrus; NoLi, Notopodial ligule; NeLi, Neuropodial ligule; NeLo, Neuropodial lobe; PosLo, Postchaetal lobe; NePa, Neuropodial papillae; AcLo, Neuropodial acicular lobe; InfLo, Inferior lobe; SupLo, Superior lobe; Vc, Ventral cirrus; Ax, Bx, Cx, ax, bx, cx, mAxBx, maxbx, reference points to form segments; DoEx, dorsal extension; VeEx, Ventral extension; Ba, Base; Sub, Subdistal insertion; Med, Medial insertion; tEx, Tip Extension; LDc, Dorsal cirri length. (G, redrawn from Hong et al. 2012).

243 Figure 2. Ceratonereis gracilis (Webster, 1884). Syntypes of N. gracilis A, B, D–F, L (USNM 4787). Non-types C, G–K, M–P (ECOSUR P1148), Q (ECOSUR Pilbs0965). A. Anterior end, dorsal view. B. Whole specimens, dorsal view. C. Anterior end, dorsal view. D. Chaetiger 2. E. Chaetiger 8. F. Chaetiger 16 (Dc incomplete). G. Chaetiger 2. H. Chaetiger 3. I. Chaetiger 10. J. Chaetiger 25. K. Chaetiger 44. L. Sub. Heterogomph falciger, chaetiger 2. M. Sub. Heterogomph falciger, chaetiger 2. N. Not. Sesquigomph falciger, chaetiger 44. O. Sup. Heterogomph falciger, chaetiger 44. P. Sub. Heterogomph falciger, chaetiger 44. Q. Right jaw. Scale bars. A, C, 0.5 mm; B, 1 mm; D–F, 0.25 mm; G–K, 0.3 mm; L–P, 0.02 mm; Q, 0.2 mm.

244 Figure 3. Ceratonereis longicirrata Perkins, 1980. Holotype A–C (USNM 58740); non- type D, E, H–O (ECOSUR P1164); non-type F, G (ECOSUR P0000 Las Coloradas). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Whole specimen, dorsal view. D. Anterior end, dorsal view. E. Left jaw, dorsal view. F. Pharynx, ventral view. G. Same, dorsal view. H. Not. Sesquigomph falciger, chaetiger 19. I. Sub. Heterogomph falciger, chaetiger 17. J. Sub. Heterogomph falciger, chaetiger 19. K. Chaetiger 2. L. Chaetiger 3 (Dorsal cirrus incomplete). M. Chaetiger 10. N. Chaetiger 17. O. Chaetiger 19. Scale bars. A, B 0.5 mm; C, D, 1 mm; E, 0.2 mm; F, G, 0.5 mm; H–J, 0.01 mm; K–O, 0.1 mm.

245 Figure 4. Ceratonereis setosa sp. n. Holotype, atokous female A–H, J–N (ECOSUR P0000); paratype C, I, P (ECOSUR P0000 Catoche). A and C. Anterior ends, dorsal view. B. posterior end, dorsal view. D. Not. Sesquigomph falciger, chaetiger 81. E. Sup. Heterogomph falciger, chaetiger 81. F. Sub. Heterogomph falciger, chaetiger 81. G. Sub. Heterogomph spiniger, chaetiger 81. H. Not. Sesquigomph spiniger, chaetiger 81. I. Sub. Heterogomph falciger, chaetiger 1. J. Left jaw, dorsal view. K. Chaetiger 1. L. Chaetiger 12. M. Chaetiger 30. N. Chaetiger 81. O. Chaetiger 121. P. Chaetiger 65. Scale bars. A, 2 mm; B, O 0.5 mm; C–H, 0.1 mm; I, 0.25 mm; J–N, P, 0.3 mm.

246 Figure 5. Composetia versipedata (Ehlers, 1887). Non-types A–O (ECOSUR Pilsb 989). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Pharynx, dorsal view. D. Same, ventral view. E. Left jaw, dorsal view. F. Chaetiger 1. G. Chaetiger 10. H. Chaetiger 28. I. Chaetiger 40. J. Chaetiger 53. K. Sup. Heterogomph falciger, chaetiger 10. L. Sub. Heterogomph falciger, from same. M. Sup. Heterogomph falciger, chaetiger 53. N. Sub. Heterogomph falciger, from same. O. Sup. Homogomph spiniger, from same. Scale bars A–D, 1 mm; E, O, 0.5 mm; F-J, 0.1 mm; K–N, 0.01 mm. 247 Figure 6. Composetia irritabilis (Webster, 1879). Syntypes of Nereis irritabilis female A, F, K–N (USNM 532); male D, E, G–J (USNM 533); atoke B, C (USNM 531). A. Anterior end, dorsal view. B. Pharynx, ventral view. C. Right jaw, dorsal view. D. Anterior end, dorsal view. E and F. Posterior ends, dorsal view. G. Chaetiger 2. H. Chaetiger 7. I. Chaetiger 3. J. Chaetiger 62. K. Chaetiger 2. L. Chaetiger 7. M. Chaetiger 60. N. Chaetiger 173. Scale bars. A, D, 1 mm; B, E, F, 0.5 mm; C, G–J, 0.4 mm; K–N, 0.3 mm.

248 Figure 7. Micronereis piccola Banse, 1977. Holotype male A, B, E, F (USNM 54514); paratype female D (USNM 60505); non-types C (LACM-AHF 7380), G–I (LACM-AHF 7387). A. Whole specimen, dorsal view. B. Anterior end, dorsal view. C. Middle chaetigers with oocytes. D. Female right jaw, dorsal view (current position in specimen). E. Male right jaw, dorsal view (current position in specimen). F. Homogomph spinigers, chaetiger 7 (arrow indicate damages in chaetae). G. Chaetiger 1. H. Chaetiger 3. I. Chaetiger 15. Scale bars. A, 0.5 mm; B, 0.3 mm; C, G–I, 0.1 mm; F, 0.01 mm; D, E, 0.05 mm.

249 Figure 8. Neanthes galetae Fauchald, 1977. Paratype A–K (LACM-AHF 1132). A. Whole specimen, dorsal view. B. Anterior end, dorsal view. C. Posterior end, dorsal view. D. Sup. Heterogomph falciger, chaetiger 27. E. Sub. Heterogomph falciger, chaetiger 27. F. Not. Homogomph spiniger, chaetiger 27. G. Sup. Homogomph spiniger, chaetiger 27. H. Chaetiger 1. I. Chaetiger 8. J. Chaetiger 27. K. Chaetiger 52. Scale bars. A, 1 mm; B, C, 0.5 mm; D–F, 0.01 mm; G, 0.02 mm; H–K, 0.1 mm.

250 Figure 9. Neanthes micromma (Harper, 1979). Holotype A, B (USNM 55575); paratype C–K (USNM 55576). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Sup. Homogomph spiniger, chaetiger 62. D. Not. Homogomph spiniger, chaetiger 62. E. Sub. Heterogomph spiniger, chaetiger 62. F. Sub. Heterogomph falciger, chaetiger 2. G. Sub. Heterogomph falciger, chaetiger 62. H. Chaetiger 2. I. Chaetiger 12. J. Chaetiger 62. K. Chaetiger 120. Scale bars. A, B, 0.5 mm; C–G, 0.01 mm; H–K, 0.1 mm.

251 Figure 10. Neanthes 1 sp. n. Holotype A, B, D–K (ECOSUR P0000); paratype C (ECOSUR P0000). A. Anterior end, dorsal view. B. Oral ring, frontal view. C. Posterior end, dorsal view. D. Sub. Heterogomph falciger, chaetiger 35. E. Sub. Heterogomph falciger, chaetiger 35. F. Sup. Heterogomph falciger, chaetiger 35. G. Not. Homogomph spiniger, chaetiger 35. H. Chaetiger 2. I. Chaetiger 11. J. Chaetiger 25. K. Chaetiger 35. Scale bars. A, C, 0.5mm; B, 0.25 mm; D–G, 0.01mm; H-K, 0.2 mm.

252 Figure 11. Neanthes arenaceodentata (Moore, 1903). Topotypes A–D, F–L (USNM 28093) and E (USNM 28100). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Whole specimen, dorsal view. D. Segments from middle body, ventral view. E. Whole, extended oral rings. F. Sup. Heterogomph falciger, chaetiger 45. G. Sub. Heterogomph falciger, chaetiger 45. H. Not. Homogomph spiniger, chaetiger 45. I. Chaetiger 2. J. Chaetiger 12. K. Chaetiger 26. L. Chaetiger 45. Scale bars. A, B, D, 0.5 mm; C, E, 1 mm; F–H, 0.01 mm; I–L, 0.1 mm.

253 Figure 12. Nereis arroyensis Treadwell, 1901. Syntype A–K (USNM 16005). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Pharynx, dorsal view. D. Same, ventral view. E. Same, close-up of area VII-VIII. F. Sub. Heterogomph falciger, chaetiger 12. G. Not. Homogomph falciger, chaetiger 62. H. Chaetiger 2. I. Chaetiger 12. J. Chaetiger 32. K. Chaetiger 63. Scale bars. A, B, 1 mm; C–E, 0.5mm; F, G, 0.01 mm; H–K, 0.2 mm.

254 Figure 13. Nereis panamensis Fauchald, 1972. Holotype A (USNM 53139); Paratypes B (USNM 53140); Non-types C–M (USNM 73683). A. Anterior end, dorsal view. B Whole specimens, dorsal view. C. Anterior end, dorsal view. D. Posterior end, dorsal view. E. Close- up of same, ventral view. F. Chaetiger 2. G. Chaetiger 10. H. Chaetiger 23. I. Chaetiger 45. J. Sup. Homogomph spiniger, chaetiger 23. K. Not. Homogomph falciger, chaetiger 23. L. Sub. Heterogomph falciger, chaetiger 23. M. Sup. Heterogomph falciger, chaetiger 23. Scale bars. A, C, 1mm; B, 2 mm; D, 0.5 mm; E–I, 0.2 mm; J–M, 0.01 mm.

255 Figure 14. Nereis cf. riisei 1. Non-type A–K (ECOSUR P0000). A. Anterior end dorsal view. B. Posterior end, dorsal view. C. Chaetiger 1. D. Chaetiger 11. E. Chaetiger 49. F. Chaetiger 88. G. Chaetiger 113. H. Not. Homogomph falciger, chaetiger 113. I. Sup. Homogomph spiniger, chaetiger 88. J. Sub. Heterogomph falciger, chaetiger 88. K. Sup. Heterogomph falciger, chaetiger 88. Scale bars. A, B, 1 mm; C–G, 0.2 mm; H–K, 0.01 mm.

256 Figure 15. Nereis cf. riisei 2. Non-type A, B (ECOSUR P0000); non-type C–K (ECOSUR Pilbs1196). A and C. Anterior ends, dorsal view. B and D. Posterior ends, dorsal view. E. Not. Homogomph falciger, chaetiger 52. F. Sup. Heterogomph falciger, chaetiger 28. G. Sub. Heterogomph falciger, chaetiger 28. H. Chaetiger 1. I. Chaetiger 12. J. Chaetiger 28. K. Chaetiger 52. Scale bars. A, C, 1 mm; B, D, 0.5 mm; E–G, 0.01 mm; H–K, 0.2 mm.

257 Figure 16. Nereis allenae Pettibone, 1956. Paratype A-D (USNM 27779). Kainonereis sp. 1 sp. n. Holotype E-H (USNM 0000). Profundilycastis profundus (Hartman, 1965). Holotype I, J (LACM-AHF 841). A. Anterior end, dorsal view. B. Posterior end, lateral view. C. Chaetiger 7 (arrow points position of ankylosed chaetae, blade missing). D. Ankylosed chaetae, chaetiger 6. E. Anterior end, dorsal view. F. Whole specimen, dorsal view. G. Closeup of calyx-like dorsal cirri, lateral view. H. Chaetiger 7. I. Whole specimen, dorsal view. J. Anterior end, dorsal view. Scale bars. A, B, E, I, 0.5 mm; C, H, 0.1 mm; D, 0.01 mm; F, J, 1 mm; G, 0.2 mm.

258 Figure 17. Nereis caecoides Hartman, 1965. Holotype A–O (LACM-AHF 818). A. Whole specimen. B. Anterior end, dorsal view. C. Posterior end, dorsal view. D. Right jaw, dorsal view. E. Squamose paragnaths, areas III or IV. F. Patch of squamose paragnaths. G. Not. Homgomph falciger, chaetiger 53. H. Not. Homogomph spiniger, chaetiger 53. I. Sub. Heterogomph falciger, chaetiger 4. J. Sup. Heterogomph falciger, chaetiger 53. K. Chaetiger 4. L. Chaetiger 23. M. Chaetiger 36. N. Chaetiger 53. O. Closeup of notopodial, showing homogomph falciger partially embeded, chaetiger 36. Scale bars. A–C, 0.5 mm; D, 0.2 mm; E, 0.01 mm; F, O, 0.25 mm; G–J, 0.01 mm; K–O, 0.1 mm. 259 Figure 18. Nereis caymanensis Fauchald, 1977. Holotype A–G (LACM-AHF 1206). Perinereis cariboea de León-González & Solís-Weiss, 1998. Non-type H–N (ECOSUR P0000). A. Whole specimens, dorsal view. B. Anterior end, dorsal view. C. Chaetiger 9. D. Sup. Homogomph spiniger, chaetiger 9. E. Not. Homogomph falciger, chaetiger 9. F. Sup. Heterogomph falcigers, chaetiger 9. G. Sub-Heterogomph falcigers, chaetiger 9. H. Anterior end, dorsal view. I. Posterior end, dorsal view. J. Sub. Heterogomph falciger, chaetiger 18. K. Chaetiger 10. L. Chaetiger 18. M. Chaetiger 31. N. Right jaw, dorsal view. Scale bars. A, 1 mm; B, H, I, 0.5 mm; C, K–M, 0.1 mm; D, 0.02 mm; E–G, J, 0.01 mm; N, 0.3 mm.

260 Figure 19. Perinereis bairdii (Webster, 1884). Syntypes of Nereis bairdii A–F (USNM 4786). Perinereis floridana (Ehlers, 1868). Syntypes of N. bairdii G–J (USNM 4786). A. Anterior end, dorsal view. B. Same, pharynx partially everted, ventral view. C. Chaetiger 2. D. Chaetiger 11. E. Chaetiger 34. F. Chaetiger 57. G and H. Anterior ends, dorsal view. I. Pharynx, dorsal view. J. Same, ventral view. Scale bars. A, I, J, 0.5 mm; B, 0.25 mm; C–F, 0.2 mm; G, H, 1 mm.

261 Figure 20. Perinereis bairdii (Webster, 1884). Non type A–G, J–N (ECOSUR P0000); Non- type H, I (ECOSUR P0000). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Chaetiger 2. D. Chaetiger 10. E. Chaetiger 24. F. Chaetiger 42. G. Chaetiger 56. H. Pharynx, dorsal view. I. Pharynx, ventral view. J. Left jaw, dorsal view. K. Sup. Heterogomph falciger, chaetiger 54. L. Sub. Heterogomph falciger, chaetiger 54. M. Not. Homogomph spiniger, chaetiger 54. N. Sub. Heterogomph spiniger, chaetiger 54. Scale bars. A, B, 1 mm; C–G, 0.2 mm; H, I, 0.5 mm; J, 0.3 mm; K, L, 0.01 mm; M, N, 0.05 mm.

262 Figure 21. Perinereis floridana (Ehlers, 1868). Non-type specimen A–L (ECOSUR P0000). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Chaetiger 2. D. Chaetiger 10. E. Chaetiger 28. F. Chaetiger 61. G. Chaetiger 91. H. Left jaw, dorsal view. I. Sub. Heterogomph falciger, chaetiger 61. J. Sup. Heterogomph falciger, chaetiger 61. K. Sub. Heterogomph spiniger, chaetiger 61. L. Sup. Homogomph spiniger, chaetiger 61. Scale bars. A, B, 1 mm; C–H, 0.25 mm; I, J, 0.01 mm; K, L, 0.05 mm.

263 Figure 22. Perinereis paradisea sp. n. Holotype A, B, D–O (ECOSUR P0000); paratype A insert (ECOSUR P0000); paratype C (ECOSUR P0000 Mahahual). A. Anterior end, dorsal view (insert: close-up of antennae). B. Posterior end, dorsal view. C. Anterior end, dorsal view. D. Right jaw, dorsal view. E. Pharynx, dorsal view. F. Pharynx, ventral view. G. Chaetiger 2. H. Chaetiger 10. I. Chaetiger 28. J. Chaetiger 44. K. Chaetiger 62. L. Sup. Heterogomph falciger, chaetiger 62. M. Sub. Heterogomph falciger, chaetiger 62. N. Sub. Heterogomph spiniger, chaetiger 62. O. Not. Homogomph spiniger, chaetiger 62. Scale bars. A–C, 1 mm; D, G–K, N, O, 0.25 mm; E, F, 0.5 mm, L, M, 0.01 mm.

264 Figure 23. Perinereis anderssoni Kinberg, 1865. Topotype specimen A–K (ECOSUR P0000). A. Anterior end. B. Posterior end. C. Posterodorsal tentacular cirri. D. Left jaw. E. Chaetiger 1. F. Chaetiger 10. G. Chaetiger 28. H. Chaetiger 45. I. Chaetiger 62. J. Sup. Heterogomph falciger, chaetiger 62. K. Sub. Heterogomph falciger, chaetiger 28. Scale bars. A–C, 1mm; D, 0.25 mm; E–I, 0.1 mm; J, K, 0.01 mm.

265 Figure 24. Platynereis antillensis (McIntosh, 1885). Non-types A–G (ECOSUR P0000). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Chaetiger 2. D. Chaetiger 12. E. Chaetiger 33. F. Chaetiger 61. G. Not. Homogomph falciger, chaetiger 61. Scale bars. A, B, 1 mm; C–F, 0.1 mm; G, 0.01 mm.

266 Figure 25. Platynereis hutchingsae de León-González et al. 2001. Holotype A–D, F–H (UANL 4281); paratype E, I (UANL 4282). A. Whole specimen, dorsal view. B. Whole specimens, dorsal view. C. Anterior end, dorsal view. D, Posterior end, dorsal view. E. Chaetiger 10. F. Chaetiger 64. G. Sup. Heterogomph falciger, chaetiger 64. H. Sub. Heterogomph falciger, same. I. Left jaws, dorsal view. Scale bars. A, B, 1 mm; C, D, 0.5 mm; E, I, 0.2 mm; F, 0.1 mm; G, H, 0.01 mm.

267 Figure 26. Platynereis lutea Treadwell, 1928. Non-types A–J (ECOSUR P0000). A. Anterior end, dorsal view. B. Same, showing the tube. C. Fragment of Sargassum lamine with a specimen attached. D. Not. Homogomph falciger, chaetiger 19. E. Same, chaetiger 70. F. Chaetiger 2. G. Chaetiger 11. H. Chaetiger 31. I. Chaetiger 51. J. Chaetiger 70. Scale bars. A-C, 2 mm; D, E, 0.01 mm; F-J, 0.2 mm.

268 Figure 27. Platynereis mucronata de León-González, Solís-Weiss & Valadez- Rocha, 2001. Non-types A–J (ECOSUR P0000). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Sup. Heterogomph falciger, chaetiger 77. D. Sub. Heterogomph falciger, chaetiger 77. E. Not. Homogomph falcigers, chaetiger 27. F. Chaetiger 1. G. Chaetiger 12. H. Chaetiger 39. I. Chaetiger 59. J. Chaetiger 77. Scale bars. A, B, 2 mm; C, D, 0.01 mm; E, 0.02 mm; F-J, 0.2 mm.

269 Figure 28. Platynereis sp. 1 sp. n. Holotype A, B, D–H, K–N (USNM 0000); Paratypes C, I, J (USNM 0000). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Right jaw. D. Chaetiger 1. E. Chaetiger 12. F. Chaetiger 29. G. Chaetiger 47. H. Chaetiger 66. I. Chaetiger 83. J. Same, closeup of notochaetae. K. Not. Homogomph spiniger, chaetiger 83. L. Sup. Heterogomph falciger, from same. M. Sub. Heterogomph falciger, from same. N. Not. Homogomph falciger, from same. Scale bars. A, B, 2 mm; C–I, 0.3 mm; J, 0.05 mm; K–N, 0.01 mm.

270 Figure 29. Pseudonereis citrina sp. n. Holotype A–C, F–P (ECOSUR P0000); paratypes D (ECOSUR P0000) and E (ECOSUR P0000). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Left jaw, dorsal view. D and E. Areas V and VI. F. Chaetiger 2. G. Chaetiger 10. H. Chaetiger 24. I. Chaetiger 38. J. Chaetiger 56. K. Sub. Heterogomph spiniger, chaetiger 56. L. Sup. Heterogomph falciger, chaetiger 56. M. Sub. Heterogomph falciger chaetiger 56. N. Sub. Heterogomph falciger, chaetiger 38. O. Not. Homogomph spiniger, chaetiger 38. P. Sub. Heterogomph spiniger, chaetiger 56. Scale bars. A, B, 1 mm; C, 0.3 mm; F–J, 0.2 mm; K–N, 0.01 mm; O, P, 0.03 mm.

271 Figure 30. Pseudonereis brunnea sp. n. Holotype A–H (ECOSUR P0000); paratypes I–L, M–O (ECOSUR P0000). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Chaetiger 2. D. Chaetiger 10. E. Chaetiger 26. F. Chaetiger 46. G. Chaetiger 65. H. Close up of anterior end. I. Maxillary ring, frontal view. J. Oral ring, frontal view. K. Sup. Heterogomph falciger, chaetiger 55. L. Sub. Heterogomph falciger, chaetiger 55. M. Sub. Heterogomph spiniger, chaetiger 55. N. Not. Homogomph spiniger, chaetiger 55. O. Sub. Heterogomph spiniger, chaetiger 55. Scale bars. A, B, 1 mm; C–G, 0.2 mm; H-J, 0.5 mm; K–M, 0.01 mm; N, O, 0.03 mm.

272 Figure 31. Namalycastis intermedia Glasby, 1999. Holotype A, B (USNM 178870); paratype C–I (USNM 31011). Namalycastis borealis Glasby, 1999. Holotype J–L (USNM 52926); paratype M (USNM 58129). A. Whole specimen, dorsal view. B. Anterior end, dorsal view. C. Right jaw, dorsal view. D. Sub. Heterogomph falciger, chaetiger 82. E. Sup. Heterogomph falciger, chaetiger 52. F. Sup. Heterogomph spiniger, chaetiger 52. G. Chaetiger 13. H. Chaetiger 52. I. Chaetiger 82. J. Anterior end, dorsal view. K. Posterior end, dorsal view. L. Abnormal dorsal cirrus from middle body, dorsal view. M. Right jaw, dorsal view. Scale bars. A, J, K, 1 mm; B, L, 0.5 mm; C, M, 0.25 mm; D, 0.01 mm; E, F, 0.03 mm, G–I, 0.1 mm.

273 Figure 32. Namanereis christopheri sp. n. Holotype A, B, F–O (LACM AHF POLY 1227); paratypes C–E (LACM AHF POLY 1229). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Anterior end, dorsal view. D. Areas V and VI. E. Right jaw, dorsal view. F. Chaetiger 5. G. Chaetiger 21. H. Chaetiger 49. I. Chaetiger 90. J. Sup. Sesquigomph spiniger, chaetiger 49. K. Sub. Heterogomph falciger, chaetiger 49. L. Sup. Heterogomph falciger, chaetiger 49. M. Sub. Heterogomph falciger, chaetiger 49. N. Sup. Sesquigomph spiniger, close-up of blade, chaetiger 49. O. Sup. Heterogomph falciger, close-up of blade, chaetiger 49. Scale bars. A–D 0.5 mm; E, 0.1 mm; F–I, 0.1 mm; J, K, 0.05 mm; L, M, 0.3 mm; N, O, 0.01 mm.

274 Figure 33. Ceratocephale oculata Banse, 1977. Holotype A (USNM 51079); paratypes B, D, F–I (USNM 52096); non-type C, I (USNM 1183121). A. Whole specimens, dorsal view. B. Whole fragments, dorsal view. C. Chaetigers 1–8, lateral view (arrow: chaetiger 1). D and E. Anterior ends, dorsal view. F. Chaetiger 2. G. Chaetiger 8. H. Chaetiger 19. I. Chaetiger 6 (VC missing). Scale bars. A, B, E, 1 mm; C, D, 0.3 mm; F–I, 0.1 mm.

275 Figure 34. Laeonereis culveri (Webster, 1879). Syntypes of Nereis culveri A–N (USNM 541). A and B. Anterior ends, dorsal view. C. Same, posterior end, dorsal view (insert: pygidium, lateral view). D. Anterior chaetigers, dorsal view. E. Same, ventral view. F. Middle chaetigers, dorsal view. G. Same, ventral view. H. Middle-posterior chaetigers, dorsal view. I. Same, ventral view. J. Chaetiger 2. K. Chaetiger 10. L. Chaetiger 41. M. Chaetiger 82. N. Chaetiger 102. Scale bars. A, B, 2 mm; C–I, 1 mm; J–N, 0.3 mm.

276 Figure 35. Nicon vossae sp. n. Holotype A, B, E–H (ECOSUR P0000); Paratype D, I–M (ECOSUR P0000); non-type C (ECOSUR Pilsb736). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Right jaw, dorsal view. D. Middle chaetigers, laterodorsal view. E. Chaetiger 1. F. Chaetiger 12. G. Chaetiger 35. H. Chaetiger 59. I. Sub. Heterogomph falciger, chaetiger 43. J, Same, close-up of tip. K. Sup. Heterogomph falciger, chaetiger 43. L. Sub. Heterogomph spiniger, chaetiger 43. M. Not. Homogomph spiniger, chaetiger 43. Scale bars. A–D, 1 mm; E–H, L, M, 0.3 mm; I, K, 0.01 mm.

277 Figure 36. Olganereis salazari sp. n. Holotype, atokous male A–O (ECOSUR P0000). A. Anterior end, dorsal view. B. Posterior end, dorsal view. C. Whole specimen, dorsal view. D. Pharynx, dorsolateral view (arrows point to papillae). E. Pharynx, lateroventral view (arrows point to papillae). F. Chaetiger 2. G. Chaetiger 10. H. Chaetiger 25. I. Chaetiger 44. J. Chaetiger 64. K. Left jaw, dorsal view. L. Sup. Heterogomph falciger, chaetiger 64. M. Sub. Heterogomph falciger, chaetiger 64. N. Not. Homogomph spiniger, chaetiger 64. O. Sub. Heterogomph spiniger, chaetiger 64. Scale bars. A, B, D, E, 0.5 mm; C, 1 mm; F-K, 0.1 mm; L–O, 0.01 mm.

278 Figure 37. Stenoninereis lackeyi (Hartman, 1958). Paratype A, C, D, F–H (AHF POLY 806); non-type B, E, I–M (USNM 45699). A, B. Whole specimens. C. Anterior end, dorsal view. D. Posterior end, dorsal view. E. Left jaw, dorsal view. F. Chaetiger 7. G. Chaetiger 19. H. Chaetiger 26. I. Chaetiger 2. J. Chaetiger 6. K. Chaetiger 17. L. Chaetiger 23. M. Chaetiger 29. Scale bars. A, B, 1 mm; C, D, 0.25 mm; E, 0.05 mm; F-M, 0.1 mm.

279 Figure 38. Stenoninereis tecolutlensis de León-González & Solís-Weiss, 1997. Holotype A–C (USNM 174870). Stenoninereis martini Wesenberg-Lund, 1958. Syntypes D–F (USNM 29726). A. Whole specimen, dorsal view. B. Anterior end, dorsal view. C. Posterior end, dorsal view. D, E. Whole specimens, dorsal view. F. Anterior end, dorsal view. Scale bars. A, 1 mm; B, C, F, 0.3 mm; D, E, 0.5 mm.

280 Figure 39. Stenoninereis elisae sp. n. Syntypes A–M (USNM 55366). A, B. Whole specimens, dorsal view. C. Anterior end, dorsal view. D. Posterior end, dorsal view. E. Chaetiger 6. F. Chaetiger 16. G. Chaetiger 18. H. Chaetiger 24. I. Sub. Heterogomph spinigers, chaetiger 18. J–L. Sub. Heterogomph falcigers, chaetiger 18. M. Not. Homogomph spiniger, chaetiger 49. Scale bars. A, B, 0.5mm; C, 0.25 mm; E–H, 0.2 mm; I–L 0.01 mm; M, 0.03 mm.

281 CONCLUSIONES

En el primer capítulo, se redescribió la especie Nereis oligohalina (Rioja, 1956) con base en material de Veracruz, costas mexicanas del Golfo de México, y N. garwoodi

González-Escalante & Salazar-Vallejo, 2003, del Caribe mexicano, proponiéndose un neotipo para la primera, y lectotipo y paralectotipos para la segunda. Material adicional del Pacífico mexicano e identificado como N. oligohalina, demostró tener características morfológicas suficientes para separarla de la especie del Golfo de México y fue descrita como nueva especie, N. confusa Conde-Vela & Salazar-Vallejo, 2015. El patrón de pigmentación del cuerpo resultó ser un carácter novedoso para la identificación, además de las diferencias morfológicas entre los organismos reproductivos. Además se concluyó, como otros autores, que son necesarias revisiones taxonómicas detalladas para delimitar la distribución de las especies o para constatar su estatus de amplia distribución, sobre todo por las implicaciones que tienen en trabajos biogeográficos o filogenéticos, entre otros.

Para el segundo capítulo, en total se examinaron 845 ejemplares, de los cuales 777 provinieron de las muestras de las Expediciones del Mar Profundo de la Universidad de

Miami, el resto se encuentra depositado en las colecciones de ECOSUR, USNM,

LACM-AHF y UANL.

En conclusión, para la región del Gran Caribe, se encontraron 108 nombres registrados en la literatura. De estos registros, 44 especies se consideraron válidas y 38 resultaron ser registros cuestionables, de estos últimos 8 especies tienen estatus de anfiamericanas. Las 26 especies faltantes se detallan a continuación.

282

7 especies se encontraron suficientes características que permitieron considerarlas como válidas y fueron restablecidas a partir de registros de amplia distribución.

Ceratonereis gracilis (Bermuda), Perinereis bairdii (Bermuda), y Stenoninereis lackeyi

(Florida) fueron restablecidas con base en la revisión de material tipo; sin embargo

Platynereis antillensis (Bermuda), P. lutea (Mar de Sargaso), P. trimaculosa (Golfo de

México), Namanereis beumeri (Cuba), fueron restablecidos de manera informal,

únicamente basados en la literatura.

Unas 13 especies del género Nereis y varias pertenecientes al grupo N. riisei, requieren redescripciones con material tipo y topotipos, principalmente porque la descripciones originales no permiten considerarlas válidas; de éstas, solo dos especies (Mastigonereis heterodonta Schmarda, 1861 y Nereis anodonta Schmarda, 1861) no cuentan con material original.

El género Profundilycastis Hartmann-Schröder, 1977, antes sinónimo de Rullierinereis, fue reinstalado ya que la especie tipo, P. profundus, presenta características distintas como la forma del prostomio y lígulas alargadas y delgadas.

Además, 5 especies válidas demostraron tener afinidades genéricas cuestionables, ya que poseen características que no corresponden con sus géneros actuales: Neanthes egregicirrata, Nereis allenae, N. caecoides, N. caymanensis, N. goajirana. Para éstas, se incluyen argumentos para el futuro establecimiento de nuevos géneros.

Finalmente, 11 formas fueron consideradas nuevas especies para la ciencia:

Ceratonereis setosa sp. n., Kainonereis chamberlini sp. n., Namanereis christopheri sp. n., Neanthes sp. 1 n. sp., Nicon vossae sp. n., Olganereis salazari sp. n., Perinereis

283 paradisea sp. n., Platynereis sp. 1 sp. n., Pseudonereis citrina sp. n., P. brunnea sp. n. y

Stenoninereis elisae sp. n. N. christopheri, P. citrina y P. brunnea se propusieron a partir de registros anfiamericanos.

Después de la revisión de glosarios especializados en terminología de poliquetos y literatura adicional como claves y descripciones, se encontró que el término lígula debiera restringirse a aquellas estructuras parapodiales que no portan las acículas

(excluyendo a los cirros parapodiales), y lóbulo a aquellas que sí. La implementación de la mejora en las mediciones parapodiales permitió mejorar las descripciones, aunque requiere futuros ajustes, como hacerla más sencilla. Por otro lado, la observación de los primeros dos setígeros dio como resultado una propuesta de clasificación en 5 tipos denominados de la A hasta E, basado en la presencia y posición de estructuras notopodiales (cirro dorsal, lígula notopodial y notoacicula). Así, la mayoría de los géneros demostraron tener el tipo C, con cirro dorsal y lígula notopodial, pero sin notoacícula, dudosamente presente en Alitta; el tipo A, caracterizado por la ausencia de estructuras notopodiales, está presente en los géneros Micronereis y Stenoninereis; el tipo B cuenta solamente con cirro dorsal y lóbulo neuropodial, pero tiene el cirro ventral duplicado, y se encuentra en los géneros Ceratocephale y Gymnonereis; el tipo D es similar al C, pero además tiene la notoacicula, presente en los géneros Ceratonereis y

Alitta; y el tipo E es único de los namaneréidinos, además de ser el único presente en todo el cuerpo, caracterizado por tener la notoacícula inmersa en el lóbulo neuropodial, resultado de la reducción del notopodio. Se espera que estas nuevas consideraciones morfológicas permitan realizar descripciones más precisas y eficientes y una delimitación más adecuada de las especies de esta familia.

284

REFERENCIAS

Amaral, A.C.Z., Nallin, S.A.H., Steiner, T.M., Forroni, T. de O., y Filho, D.G., 2013.

Catálogo das espécies de Annelida Polychaeta do Brasil.

http://www.ib.unicamp.br/museu_zoologia/sites/www.ib.unicamp.br.museu_zoolo

gia/files/Cat%C3%A1logo_Polychaeta_Brasil_Amaral_et_al_2013_1a.pdf, 141

pp. (Accessed in 24 November 2015).

Augener, H., 1906. Reports on the results of dredging, under the supervision of

Alexander Agassiz, in the Gulf of Mexico and the Caribbean Sea, and on the east

coast of the United States, 1877 to 1880, by the U.S.S. Coast Survey Steamer

‘Blake’, Lieut. Commander C.D. Sigsbe. Bulletin of the Museum of Comparative

Zoology, 43(4), pp.91–196.

Báez, D.P., y Ardila, N.E., 2003. Poliquetos (Annelida: Polychaeta) del Mar Caribe

colombiano. Biota Colombiana, 4(1), pp.89–109.

Bakken, T., y Wilson, R.S., 2005. Phylogeny of nereidids (Polychaeta, Nereididae) with

paragnaths. Zoologica Scripta, 34(5), pp.507–547.

Carrera-Parra, L.F., y Salazar-Vallejo, S.I., 2011. Redescriptions of Eunice filamentosa

and E. denticulata and description of E. tovarae n. sp. (Polychaeta: Eunicidae),

highlighted with morphological and molecular data. Zootaxa, 2880, pp.51–64.

Conde-Vela, V.M., y Salazar-Vallejo, S.I., 2015. Redescriptions of Nereis oligohalina

(Rioja, 1946) and N. garwoodi González-Escalante & Salazar-Vallejo, 2003 and

description of N. confusa sp. n. (Annelida, Nereididae). Zookeys, 518, pp.15–49.

285 de León-González, J.Á., 2009. Nereididae. En: J.Á. de León-González, J.R. Bastida-

Zavala, L.F. Carrera-Parra, M.E. García-Garza, A. Peña-Rivera, S.I. Salazar-

Vallejo y V. Solís-Weiss, eds., Poliquetos (Annelida: Polychaeta) de México y

América Tropical. Monterrey: Universidad Autónoma de Nuevo León, pp.325–

354. de León-González, J.Á., y Solís-Weiss, V., 1997. A new species of Stenoninereis

(Polychaeta: Nereididae) from the Gulf of Mexico. Proceedings of the Biological

Society of Washington, 110(2), pp.198–202. de León-González, J.Á., y Solís-Weiss, V., 1998. The genus Perinereis (Polychaeta:

Nereididae) from Mexican littoral waters, including the description of three new

species and the redescriptions of P. anderssoni and P. elenacasoae.

Proceedings of the Biological Society of Washington, 111(3), pp.674–693. de León-González, J.Á., Solís-Weiss, V., y Ochoa-Rivera, V., 1999. Nereidids

(Polychaeta) from the Caribbean Sea and adjacent coral islands of the Southern

Gulf of Mexico. Proceedings of the Biological Society of Washington, 112(4),

pp.667–681. de León-González, J.Á.,, Solís-Weiss, V., y Valadez-Rocha, V., 2001. Two new species

of Platynereis (Polychaeta: Nereididae) from Eastern Mexican shores.

Proceedings of the Biological Society of Washington, 114(2), pp.389–395.

Dean, H.K., 2012. A literature review of the Polychaeta of the Caribbean Sea. Zootaxa,

3596, pp.1–86.

Ehlers, E., 1887. Reports on the results of dredging, under the direction of L. F.

286

Pourtalès, during the years 1868-1870, and the Alexander Agassiz, in the Gulf of

Mexico (1877-78) and in the Caribbean Sea (1878-79), in the U.S. Coast Survey

Steamer ‘Blake’. XXI. Memoirs of the Museum of Comparative Zoology at

Harvard College, 15, pp.1–335.

Fitzhugh, K., 1987. Phylogenetic relationships within the Nereididae (Polychaeta):

implications at the subfamily level. Bulletin of the Biological Society of

Washington, 7, pp.174–183.

Glasby, C.J., 1991. Phylogenetic relationships in the Nereididae (Annelida: Polychaeta),

chiefly in the subfamily Gymnonereidinae, and the monophyly of the

Namanereidinae. Bulletin of Marine Science, 48(2), pp.559–573.

Glasby, C.J., 1999. The Namanereidinae (Polychaeta: Nereididae). Part 1, Taxonomy

and Phylogeny. Records of the Australian Museum, Supplement, [online] 25,

pp.1–129. Available at:

[Accessed 7 Dec. 2014].

Grube, A.E., 1857. Annulata örstediana. Enumeratio Annulatorum, quae in itinere per

Indiam occidentalem et Americam centralem annis 1845-1848 suscepto legit cl.

A.S. Örsted, adjectis speciebus nonnullis a cl. H. Kröyero in itinere ad Americam

meridionalem collectis. [1. Familia Aphroditea – F. Euniceae]. Videnskabelige

Meddelelser fra Dansk Naturhistorisk Forening for 1856, pp.44–62.

Grube, A.E., 1858. Annulata örstediana. Enumeratio Annulatorum, quae in itinere per

Indiam occidentalem et Americam centralem annis 1845-1848 suscepto legit cl.

A.S. Örsted, adjectis speciebus nonnullis a cl. H. Kröyero in itinere ad Americam

287

meridionalem collectis, (Fortsættelse [continued]) [2. Familia Euniceae - F.

Syllidea]. Videnskabelige Meddelelser fra Dansk Naturhistorisk Forening for

1857, pp.158–186.

Grube, A.E., 1859. Annulata örstediana. Enumeratio Annulatorum, quae in itinere per

Indiam occidentalem et Americam centralem annis 1845-1848 suscepto legit cl.

A.S. Örsted, adjectis speciebus nonnullis a cl. H. Kröyero in itinere ad Americam

meridionalem collectis, (Slutning [conclusion]) [3. Familia Ariciaea - F.

Sipunculacea and Index]. Videnskabelige Meddelelser fra Dansk Naturhistorisk

Forening for 1858, pp.105–120.

Hartman, O., 1942. Report on the scientific results of the Atlantis Expeditions to the

West Indies under the joint auspices of the University of Havana and Harvard.

The Polychaetous Annelida. Memorias de la Sociedad Cubana de Historia

Natural, 16(2), pp.89–104.

Hartman, O., 1959a. Catalogue of the polychaetous annelids of the world. Part I. Allan

Hancock Foundation Publications. Occasional Paper., 23(1), pp.1–353.

Hartman, O., 1959b. Catalogue of the polychaetous annelids of the world. Part II. Allan

Hancock Foundation Publications. Occasional Paper., 23(2), pp.355–628.

Hummelinck, P.W., 1981. Land and fresh-water localities. Studies on the Fauna of

Curaçao and other Caribbean Islands, 192, pp.1–133.

Klautau, M., Russo, C.A.M., Lazoski, C., Boury-Esnault, N., Thorpe, J.P., y Solé-Cava,

A.M., 1999. Does cosmopolitanism result from overconservative systematics? A

case study using the marine sponge Chondrilla nucula. Evolution, 53(5),

288

pp.1414–1422.

Liñero-Arana, I., y Díaz-Díaz, Ó., 2007. Nuevas adiciones de Nereididae (Annelida:

Polychaeta) para las costas de Venezuela. Boletín del Instituto Oceanográfico de

Venezuela, 46(2), pp.149–159.

Liñero-Arana, I., y Reyes-Vásquez, G., 1979. Nereidae (Polychaeta, Errantia) del Golfo

de Cariaco, Venezuela. Boletín del Instituto Oceanográfico de la Universidad de

Oriente, 18(1,2), pp.3–12.

McIntosh, W.C., 1885. Report on the Annelida Polychaeta dredged by H.M.S.

Challenger during the years 1873-1876. Challenger Reports, 12, pp.1–554.

Perkins, T.H., y Savage, T., 1975. A bibliography and checklist of polychaetous anelids

of Florida, the Gulf of Mexico, and the Caribbean Region. Florida Marine

Research Publications, 14, pp.1–62.

Pettibone, M.H., 1971. Revision of Some Species Referred to Leptonereis, Nicon, and

Laeonereis (Polychaeta: Nereididae). Smithsonian Contributions to Zoology, 104,

pp.1–53.

Salazar-Vallejo, S.I., 1996. Lista de especies y bibliografía de poliquetos (Polychaeta)

del Gran Caribe. Anales del Instituto de Biología de la Universidad Nacional

Autónoma de México, Serie Zoología, 67(1), pp.11–50.

Salazar-Vallejo, S.I., 2000. Biogeografía marina del Gran Caribe. Interciencia, 25(1),

pp.7–12.

Salazar-Vallejo, S.I., y Jiménez-Cueto, M.S., 1997. Neréididos (Polychaeta) del Caribe

mexicano con una clave para las especies del Gran Caribe. Revista de Biología 289

Tropical, 44(45), pp.361–377.

Santos, C.S.G., 2007. Nereididae from Rocas Atoll (North-East, Brazil). Arquivos do

Museu Nacional do Rio de Janeiro, 65(3), pp.369–380.

Santos, C.S.G., y Lana, P. C., 2000. Nereididae (Annelida, Polychaeta) da costa

Nordeste do Brasil. I. Padrões regionais e zoogeográficos de distribução.

Iheringia, Série Zoologia, 88, pp.181–188.

Santos, C.S.G., y Lana, P. C., 2001. Nereididae (Annelida, Polychaeta) da Costa

Nordeste do Brasil. II. Gêneros Namalycastis, Ceratocephale, Laeonereis e

Rullierinereis. Iheringia, Série Zoologia, 91, pp.137–149.

Santos, C.S.G., y Lana, P. C., 2003. Nereididae (Polychaeta) Da Costa Nordeste Do

Brasil. III. Gêneros Ceratonereis e Nereis. Iheringia, Série Zoologia, 93(1), pp.5–

22.

Santos, C.S.G., Pleijel, F., Lana, P. C., y Rouse, G.W., 2005. Phylogenetic relationships

within Nereididae (Annelida: Polychaeta). Invertebrate Systematics, 19, pp.557–

576.

Treadwell, A.L., 1924. Polychaetous annelids collected by the Barbados-Antigua

Expedition from the University of Iowa in 1918. University of Iowa Studies in

Natural History, 10(4), pp.3–23.

Vanegas-Espinosa, V., Díaz-Díaz, Ó., y Liñero-Arana, I., 2007. Nereididae Lamarck,

1818 (Annelida: Polychaeta) de la Costa Occidental de Venezuela. Boletín del

Instituto Oceanográfico de Venezuela, 46(2), pp.119–127.

Villalobos-Guerrero, T.F., y Carrera-Parra, L.F., 2015. Redescription of Alitta succinea 290

(Leuckart, 1847) and reinstatement of A. acutifolia (Ehlers, 1901) n. comb. based

upon morphological and molecular data (Polychaeta: Nereididae). Zootaxa,

3919(1), pp.157–178.

Wesenberg-Lund, E., 1958. Lesser Antillean Polychaetes, chiefly from brackish water

with a Survey and a Bibliography of Fresh and Brackish – Water Polychaetes.

Studies on the Fauna of Curaçao and other Caribbean Islands, 8(1), pp.1–41.

Williams, D.D., 2004. Review of the polychaete genus Namanereis (Nereididae) in the

Caribbean Region, with a record of N. hummelincki from deep freshwater wells on

Barbados. Caribbean Journal of Science, 40, pp.401–407.

Wilson, R.S., 2000. Nereididae. En: P.L. Beesley, G.J. Ross and C.J. Glasby, eds.,

Polychaetes and Allies: The southern synthesis. Fauna of Australia. Vol. 4A

Polychaeta, Myzostomida, Pogonophora, Echiura, Sipuncula. Melbourne: CSIRO

Publishing, pp.139–141.

Yáñez-Rivera, B., y Carrera-Parra, L.F., 2012. Reestablishment of Notopygos megalops

McIntosh, description of N. caribea sp. n. from the Greater Caribbean and

barcoding of ‘amphiamerican’ Notopygos species (Annelida, Amphinomidae).

Zookeys, 223, pp.69–84.

Yáñez-Rivera, B., y Salazar-Vallejo, S.I., 2011. Revision of Hermodice Kinberg, 1857

(Polychaeta: Amphinomidae). Scientia Marina, 75(2), pp.251–262.

291

ANEXOS

Información para los autores

292

Preparing final files for publication in Zootaxa

Please consult the Information for Authors: www.mapress.com/zootaxa/author.html. If possible, use the common font Times New Roman for the main text and use as little formatting as possible (use only bold and italics where necessary; and indent the paragraphs from the second of each section). Special symbols (e.g. male or female sign) should be avoided because they are likely to be altered when files are converted. You are recommended to code males as m# and females as f#. We can replace them easily. Please provide the following information when submitting the final files to enable us to typeset and index your paper: a) name of corresponding author and email; b) family names of authors and preferred short running title (<60 alphabets); c) num- ber of plates and cited references; and d) high taxon (as in Zootaxa sections) and number of new taxa described in the paper.

Specific points to note: 1) Text of the manuscript. Please save the file containing the text and tables as a ms word or RTF file. Do not double space or use footer/header. Captions of plates/figures should be inserted where you want your figures to be inserted, or listed at the end of the manuscript. Vector diagrams/charts generated in programs such as excel can be embedded in the text file as well. 2) Tables. Please use the table function in your word processor to build tables so that the cells can be easily re-sized to fit the page by the typesetters. Never use the Tab key to type tables, nor use space bar to adjust space. If you did that, please cor- rect them. 3) Figures. Please note that the journal has a matter size of 25 cm x 17 cm (printable area of the main text and plates) and is printed on A4 paper. If the final size of your plates is greater than this, please crop extra white areas around plates or reduce it to this size; this will reduce file size. Plates containing only black&white lines/dots are line art works and must be scanned as such (i.e. 1 bit, monochrome line art) at 600 dpi (maximum 1200 dpi). They should be saved in tiff file and LZW compression is recommended to be used to reduce file size for easy sub-mission by e-mail. Plates containing greyscale drawings and photographs should be scanned at 300 dpi and saved in TIFF (use LZW compression) or jpeg at the highest quality. Please do not modify photographs in the jpeg files; the print quality would be severely altered if you did that. You can modify files in the TIFF file and when completed, convert to high quality jpeg for submission. If you have colour figures, it is best to group them together in plates, which will save cost. 4) Reference list. This must be formatted using Zootaxa style (see below for examples of main types of papers): Please note that (1) journal titles must be written in full (not abbreviated); (2) journal titles and volume numbers are followed by a “,”; (3) page ranges are connected by a “n dash”, not a hyphen “-”, which is used to connect two words. For websites, it is important to include the last date when you see that site, as it can be moved or deleted from that address in the future. 5) Submission of files. Please send final files by e-mail (or ftp) to your subject editor who will then forward files of accepted versions to Zootaxa office. Plates (if many) may be sent directly to Zootaxa office if the subject editor agrees. year in brackets, not followed by a period Journal paper journal title in italics, spelt in full (never abbreviate), followed by comma volume/issue number followed by comma Gusarov, V.I. (2002) A revision of Nearctic species of the genus Geostiba Thomson, 1858 (Coleoptera: Staphylinidae: Aleochar- inae). Zootaxa, 81, 1–88.

connect author names by “&”, not “and” Book chapterpage ranges connected by n-dash

Newton, A.F., Thayer, M.K., Ashe, J.S. & Chandler, D.S. (2000) Staphylinidae Latreille, 1802. In: Arnett, R.H. & Thomas, M.C. (Eds.), American Beetles. Vol.1. Archostemata, Myxophaga, Adephaga, Polyphaga: Staphyliniformia. CRC Press, Boca Raton, pp. 272– 418. author family name first, followed by initials of given names (no space between initials) book title in italics publisher name first, followed by location Book

Baker, E.W. & Tuttle, D.M. (1994) A Guide to Spider Mites (Tetranychidae) of the United States. Indira Publishing House, West Bloomfield, Michigan, 347 pp.

Internet resources

O’Connor, R.J., Dunn, E, Johnson, D.H., .Jones, S.L., Petit, D., Pollock, K., Smith, C.R., Trapp, J.L. & Welling, E. (2000) A programmatic review of the North American Breeding Bird Survey: report of a peer review panel. U.S. Geological Survey Patuxent Wildlife Research Center, Laurel, Maryland. Available from http://www.mp2-pwrc.usgs.gov/bbs//bbsreview/ (accessed 3 April 2003)

include the last date when you see that site, as it can be moved or deleted from that address in the future