Open Access
How to translate text using browser tools
18 April 2024 Phylogenetic Classification of Living and Fossil Ray-Finned Fishes (Actinopterygii)
Thomas J. Near, Christine E. Thacker
Author Affiliations +
Abstract

Classification of the tremendous diversity of ray-finned fishes (Actinopterygii) began with the designation of taxonomic groups on the basis of morphological similarity. Starting in the late 1960s morphological phylogenetics became the basis for the classification of Actinopterygii but failed to resolve many relationships, particularly among lineages within the hyperdiverse Percomorpha. The introduction of molecular phylogenetics led to a dramatic reconfiguration of actinopterygian phylogeny. Refined phylogenetic resolution afforded by molecular studies revealed an uneven diversity among actinopterygian lineages, resulting in a proliferation of redundant group names in Linnean-ranked classifications. Here we provide an unranked phylogenetic classification for actinopterygian fishes based on a summary phylogeny of 830 lineages of ray-finned fishes that includes all currently recognized actinopterygian taxonomic families and 287 fossil taxa. We provide phylogenetic definitions for 90 clade names and review seven previously defined names. For each of the 97 clade names, we review the etymology of the clade name, clade species diversity and constituent lineages, clade diagnostic morphological apomorphies, review synonyms, and provide a discussion of the clade's nomenclatural and systematic history. The new classification is free of redundant group names and includes only one new name among the 97 clade names we review and describe, yielding a comprehensive classification that is based explicitly on the phylogeny of ray-finned fishes that has emerged in the 21st century and rests on the foundation of the previous 200 years of research on the systematics of ray-finned fishes.

Introduction

There are currently more than 35,085 described species of ray-finned (Actinopterygii) fishes (Fricke et al. 2023), comprising nearly half of the total living species diversity of vertebrates. The first classifications of the immense diversity of Actinopterygii were the culmination of several important and ambitious surveys of ray-finned and teleost fishes on the basis of comparative anatomy (e.g., Müller 1845a; Cope 1871a, 1871b; T. N. Gill 1872; Goodrich 1909; Jordan 1923; Regan 1929; Garstang 1931; Berg 1940; Greenwood et al. 1966) and morphological studies that were among the first to use cladistic methods (G. J. Nelson 1968, 1969c, 1973; Patterson 1973; Rosen 1973). These early efforts provided support for the monophyly of major clades of Actinopterygii still recognized today, including groups such as sturgeons, gars, tarpons and eels, catfishes, salmons, anglerfishes, tunas, gobies, and flatfishes. However, prior to the application of molecular data, the relationships among many of the major lineages of ray-finned fishes remained unresolved and specific phylogenetic hypotheses relied on the interpretation of a few key morphological characters (e.g., Patterson 1973; Rosen 1973, 1985; Lauder and Liem 1983; Patterson 1993, 1996).

The introduction of molecular data to phylogenetics revolutionized the inference of the tree of life and brought astounding insights, including the paraphyly of Prokaryota (Woese and Fox 1977), the discovery of the inclusive placental mammal lineage Afrotheria (Stanhope et al. 1998), and the resolution of ctenophores as the sister lineage of all other metazoans (C. W. Dunn et al. 2008). In a similar way, molecular data have had an astonishing effect on the resolution of the phylogenetic relationships of Actinopterygii (Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Miya and Nishida 2015; Hughes et al. 2018; Ghezelayagh et al. 2022; Mu et al. 2022), with nearly every part of the ray-finned fish phylogeny modified as a result of molecular analyses (Dornburg and Near 2021). In the first years of the 21st century, morphology alone was the basis for review papers and authoritative reference texts on the relationships and classification of Actinopterygii (e.g., A. C. Gill and Mooi 2002; Stiassny et al. 2004; J. S. Nelson 2006), indicating the application of molecular data to the phylogenetics of fishes has lagged behind the study of other groups of vertebrates. Given the enormous diversity of Actinopterygii, their ecological divergence throughout nearly every available aquatic habitat, and the variety and extent of their phenotypic disparity, it is unsurprising that morphological studies have been unable to resolve many of the phylogenetic relationships within Actinopterygii.

Over the past 10 years molecular phylogenetics has significantly influenced the classification of Actinopterygii (Near, Eytan, et al. 2012; Wainwright et al. 2012; Near et al. 2013; W. L. Smith et al. 2016; Betancur-R et al. 2017; Dornburg and Near 2021). Studies based on nuclear and mitochondrial gene sequences are now complemented by those using comprehensive datasets of genomic sequences (e.g., Malmstrøm et al. 2016; Arcila et al. 2017; Hughes et al. 2018; Ghezelayagh et al. 2022; Melo, Sidlauskas, et al. 2022; Glass et al. 2023). Phylogenomic data mitigate the issues that may mislead morphological studies; in particular, the data are extremely abundant and there are strategies to detect and accommodate incomplete lineage sorting, introgression, and paralogous loci (Bravo et al. 2019; Simion et al. 2020; M. L. Smith and Hahn 2022). The sheer size of genomic datasets is likely to compensate for random and systematic errors that affect phylogenetic inferences, simply by amplifying a consistent phylogenetic signal over any noise (Simion et al. 2020). Empirical support for this theory would be the repeated inference of the same phylogenetic relationships from different molecular datasets.

As phylogenetic studies of Actinopterygii using larger molecular datasets with inclusive taxonomic sampling became practical, a remarkable result has been the extent to which the molecular phylogenies of ray-finned fishes agree with one another (e.g., Miya et al. 2005; Near, Eytan, et al. 2012; Betancur-R et al. 2017; Chakrabarty et al. 2017; Hughes et al. 2018; Ghezelayagh et al. 2022; Melo, Sidlauskas, et al. 2022). The results of independent phylogenetic analyses are not congruent in every respect, but an overall highly supported phylogeny of Actinopterygii has emerged from analysis of molecular data in the 21st century (Dornburg and Near 2021). The new consensus phylogeny supports traditional relationships such as the resolution of Ostariophysi, Siluriformes, Esocidae, Acanthomorpha, Atheriniformes, Pleuronectoidei, Lophioidei, and Tetraodontoidei as monophyletic groups, but includes relationships not inferred from traditional morphological studies across the entire phylogeny of Actinopterygii (Dornburg and Near 2021). The molecular consensus crucially provides unprecedented resolution in portions of the actinopterygian phylogeny that have been historically difficult to resolve, in particular among lineages of Percomorpha that formerly comprised the largest polytomy in vertebrate phylogenetics (Figure 1; G. J. Nelson 1989; A. C. Gill and Mooi 2002; Dornburg and Near 2021). Molecular phylogenies are also amenable to calibration with fossils to estimate divergence times and evolutionary rates, allowing insight into the mechanisms that generate biodiversity. The known fossil record for Actinopterygii is continually improving (Appendix 1). Fossil-calibrated phylogenies provide estimates of the timing of diversification of ray-finned fishes, placing the origin of Actinopterygii in the Carboniferous (Giles et al. 2017, 2023) and highlighting the Eocene (56.0–33.9 Ma) as an important time in the diversification of percomorph fishes that dominate marine habitats (Ghezelayagh et al. 2022). Such inferences on the timing of lineage diversification would be impossible to resolve with morphological data alone. Instead, we may now use the time-calibrated phylogenies to understand the tempo and patterns of species diversification (e.g., Rincon-Sandoval et al. 2020; Troyer et al. 2022; Friedman and Muñoz 2023) and revisit the abundant, detailed morphological data available and interpret its evolution in the context of evolutionary patterns revealed by genomic-scale phylogenies (e.g., Nakae and Sasaki 2010; Chanet et al. 2013; M. G. Girard et al. 2020).

FIGURE 1.

Comparison of phylogenies and classifications of Acanthomorpha. In the three phylogenies the colors of the branches indicate traditional classifications: red branches are non-percomorph acanthomorphs, orange branches are non-perciform Percomorpha, and blue branches are Perciformes (sensu lato). The phylogeny from A. C. Gill and Mooi (2002) is a summary hypothesis based on morphology. The phylogeny used to show the classifications of Betancur-R et al. (2017) and Dornburg and Near (2021) are based primarily on molecular studies. Numbers in parentheses indicate the number of taxonomic families.

img-z3-1_03.jpg

An important application of a robust phylogeny is to provide the framework for a classification. In the case of Actinopterygii, many of the lineages resolved in the 21st century molecular phylogeny had already been known and named in taxonomies based primarily on morphological inferences (e.g., Bleeker 1859; T. N. Gill 1872; Greenwood et al. 1966; J. S. Nelson 2006). Linnaean ranked classification requires the use of primary taxonomic categories: Actinopterygii is a Linnaean class, containing the ranks of order, family, genus, and species, each of which must be assigned for every taxon. In the Linnaean-ranked classification system of Actinopterygii, 28% of the approximately 515 taxonomic families are monotypic or monogeneric. In these cases, the family-group and genus names are redundant: both names refer to the same group of taxa. Consider the Salamanderfish (Lepidogalaxias salamandroides), which is the sister lineage of a clade containing more than 21,400 species of euteleost fishes (J. Li, Xia, et al. 2010; McDowall and Burridge 2011; Burridge et al. 2012; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Campbell, López, et al. 2013; Davis et al. 2016; W. L. Smith et al. 2016; Campbell, Alfaro, et al. 2017; Hughes et al. 2018; Straube et al. 2018; Rosas Puchuri 2021; Mu et al. 2022). In the Linnaean rank-based classification, L. salamandroides is classified as the only species in the family Lepidogalaxiidae, which is the only family in the order Lepidogalaxiiformes, which is the only order in the subcohort Lepidogalaxii (Betancur-R et al. 2017). In this taxonomy, Lepidogalaxias, Lepidogalaxiidae, Lepidogalaxiiformes, and Lepidogalaxii all have the same composition. Using these nested ranks to include the single species Lepidogalaxias salamandroides conveys no information and only lists several redundant group names.

An alternative to the Linnaean system is the unranked phylogenetically-based taxonomy outlined in the PhyloCode (Cantino and de Queiroz 2020). Use of the PhyloCode system prevents the proliferation of unnecessary and redundant group names and avoids the unsupported preconception that ranked categories have meaning apart from their exclusivity (de Queiroz and Gauthier 1990, 1992, 1994). In other words, it is easy to overlook that family-ranked taxa are not comparable to one another in any biologically or evolutionarily significant way; all a ranked taxon indicates is that any species within it are not included in any other taxon of equivalent rank. The PhyloCode is also strictly phylogenetic, a desirable characteristic that gives meaning to group names by explicitly tying them to clades. Clades in the PhyloCode are defined phylogenetically, differing from traditional Linnaean group names in they are defined in terms of ancestry and descent rather than being defined in terms of ranks and types. Each clade name is defined by at least two reference points on a phylogeny, either two taxa or a taxon and an apomorphy. The formulation of such phylogenetic definitions requires a comprehensive phylogenetic hypothesis. For Actinopterygii, such a hypothesis is now available, allowing for a transformation of the traditional classification of ray-finned fishes into a strictly phylogenetic framework that is as free as possible from redundant group names.

A landmark and ambitious Linnaean-ranked classification of Actinopterygii based on a phylogeny inferred from mtDNA and nuclear genes led to a proliferation of taxonomic orders and redundant group names (Betancur-R, Broughton, et al. 2013; Betancur-R et al. 2017). The proliferation of group names in Betancur-R et al. (2017) was a consequence of an effort to preserve traditional ordinal ranks for percomorph clades such as Pleuronectiformes, Tetraodontiformes, Mugiliformes, and Cyprinodontiformes. Because of their morphological disparity, these lineages were traditionally classified as taxonomic orders (e.g., A. C. Gill and Mooi 2002), set apart from the wastebin taxon Perciformes in morphology-based efforts (Figure 1). Molecular phylogenies resulted in the dramatic reallocation of lineages traditionally classified as Perciformes into nearly every major clade of Percomorpha (Figure 1), pushing traditional taxonomic orders such as Tetraodontiformes, Gobiesociformes, and Synbranchiformes from deeply nested positions into more apical resolutions in the phylogeny of Percomorpha. Within Percomorpha, the Betancur-R et al. (2017) classification delimits 34 taxonomic orders, each containing an average of only 7.4 taxonomic families; 13 of the 34 taxonomic orders contain only one or two families and only 10 of the orders have 10 or more families. In addition to delimiting less inclusive groups, the Betancur-R et al. (2017) classification treats 10% of all percomorph families as incertae sedis (Figure 1). The phylogenetic rank-free classification presented here delimits 16 major clades in Percomorpha, 13 of which are consistent with traditional taxonomic orders and contain an average of 21.8 lineages each that are treated as taxonomic families in rank-based classifications (Figure 1, Appendix 2; Dornburg and Near 2021; Ghezelayagh et al. 2022). The effort to maintain a handful of traditional taxonomic orders in Percomorpha in the Betancur-R et al. (2017) classification has resulted in a proliferation of “-iformes” group names that are neither inclusive nor phylogenetically informative (Figure 1).

Our goal in constructing a new rank-free classification of Actinopterygii is to build on and unify the punctuated progress made in the phylogenetics of ray-finned fishes in the 21st century (Dornburg and Near 2021). In this monograph, we consolidate and review the history of systematics and phylogenetics of the primary clades of ray-finned fishes, provide phylogenetic definitions for the names of 97 actinopterygian clades, introduce a summary phylogeny of 830 ray-finned fish lineages that includes 287 fossil taxa (Appendix 1), review information on species diversity in each clade, and provide a comprehensive list of constituent lineages for every major actinopterygian clade. We explicitly incorporate available phylogenies and whenever possible list diagnostic morphological apomorphies for each named clade. The new rank-free classification avoids redundant group names and attempts to preserve the exclusivity of clade names with -iformes, -oidei, and -oidea suffixes. For instance, clade names with an -iformes suffix are not nested in any other clade with a name ending in -iformes. In the phylogenetic trees, we list the genus name or the species binomial if a taxonomic family contains a single genus or species. In the clade accounts, we acknowledge the long history of the use of taxonomic families in ichthyology, listing all recognized taxonomic families but indicating those that are monotypic or monogeneric by identifying them with an asterisk as a redundant group name.

This new rank-free classification of Actinopterygii consolidates and reviews the systematic ichthyology literature of the past two centuries, builds on a consensus phylogenetic hypothesis of actinopterygian relationships, and constructs an explicit phylogenetic-based taxonomy that aims to be useful and flexible for researchers now and in the future. With this comprehensive phylogeny and classification, it is possible to investigate and communicate the overarching patterns of evolution within ray-finned fishes, which are rich in morphological complexity, ecological diversity, and biogeographic range. Combined with advances in comparative analyses that use time-calibrated molecular phylogenies, we are beginning to understand the tempo and characteristics of vertebrate evolution in aquatic habitats, across oceans and rivers, at the poles and the tropics, on coral reefs, and in environments from shallow shores to abyssal depths (e.g., Tedesco et al. 2017; Rabosky et al. 2018; Rincon-Sandoval et al. 2020; Melo, Sidlauskas, et al. 2022; E. C. Miller et al. 2022; Friedman and Muñoz 2023).

Materials and Methods

We develop a phylogeny-based classification of Actinopterygii following the principles of phylogenetic nomenclature outlined in the PhyloCode (de Queiroz and Gauthier 1990, 1992, 1994; Cantino and de Queiroz 2020), except where indicated. Articles (Art.), examples (Ex.), and recommendations (Rec.) are referred to as outlined in the International Code of Phylogenetic Nomenclature (PhyloCode) ver. 6 (Cantino and de Queiroz 2020). Following Rec. 6.1A, all scientific names of clades are italicized. This differs from the customary practice of only italicizing the genus and species names. Most of the clades presented and reviewed in this monograph are defined as minimum-crown-clades that have at minimum two internal specifiers that are both extant (Arts. 9.5 and 9.9). If there is uncertainty about the early branching history of a well-established clade, more than two specifiers are used (Art. 9.5). In a few instances, external specifiers (Art. 11.13, Ex. 1) are used to prevent the use of a clade name under specific phylogenetic hypotheses.

In following the requirements for establishing clade names (Art. 7), we provide a protologue (Art. 7.2, N. 7.2.1) for each clade name that provides everything associated with the name as it is established according to the requirements of the PhyloCode. The terms protologue and clade account are used interchangeably in this monograph. In this classification of Actinopterygii each protologue contains 10 sections.

The definition is the statement that explicitly identifies a clade as the referent of the taxon name and includes at least two specifiers (Art. 9.4). Original author citations are provided for each specifier.

Etymology is an attempt to trace the linguistic origin of clade names. Most of the clade names originate in ancient Greek, and we provide the original spelling following reference texts (D. W. Thompson 1947; Liddell et al. 1968). When the original spelling is ancient Greek, we provide a phonetic spelling of the word using the International Phonetic Alphabet (IPA 1999).

The registration number is the product of the required submission of the clade name to the official registration database (Art. 8.1). All the clade names and associated information tied to the clade definitions were submitted to the online RegNum database (Cellinese and Dell 2020), which is the official registry of clade names in PhyloCode. No registration number is given for the 11 clades that are not defined using the PhyloCode.

The reference phylogeny is a specific phylogenetic hypothesis that provides the basis and context for applying a clade name in the phylogenetic definition (Art. 7.2). The reference phylogenies were selected on the basis of taxonomic coverage and the inclusion of appropriate specifiers. Phylogenies resulting from an explicit and reproduceable analysis were the only ones considered (Rec. 9.13A). The reference phylogenies come from a total of 34 phylogenetic studies. Among the 97 clade accounts, the reference phylogeny for 46 clades are based on analysis of genomic data, 33 are based on analyses of Sanger-sequenced molecular data, 11 clades are defined using phylogenies inferred from combined molecular and morphological datasets, and seven clades are defined on the basis of phylogenies inferred from morphological characters alone. A synthetic phylogeny of 830 lineages of Actinopterygii was constructed using published phylogenetic trees in an agglomerative procedure (Beaulieu et al. 2012). All the phylogenetic studies used to construct the synthetic tree are cited among the clade accounts. The tree file in Newick is available at the Dryad data repository (Near and Thacker 2023). In the reference phylogeny section, we refer to the figure number in this monograph where the relationships of a given clade are shown and citations are provided to justify the placement of any fossil taxa in the phylogenies (Appendix 1). The absolute age intervals of the epochs, ages, and stages of the fossil record follow the Geologic Time Scale 2020 (Gradstein and Ogg 2020). In the phylogenetics section we provide a brief history of the systematics of the clade. Often this is the longest section of the clade account.

The composition of the clade includes a statement as to the current recognized species diversity and a listing of all the named major subclades of the named clade. There are no redundant group names listed in this section. If a taxonomic family in the Linnaean ranked system is monotypic or monogeneric, the species binomial or the genus name is provided. The names of fossil taxa within each named clade that are not nested in a subclade not defined in the classification are included in the phylogenetic trees and listed in the clade composition. We also highlight recent biodiversity discoveries by listing the number of new species described over the past 10 years (2013–2023).

Diagnostic apomorphies lists morphological traits that investigators have offered as diagnostic for the clade. While not required to establish a clade name in PhyloCode, we acknowledge the rich history of morphological phylogenetics in ichthyology that has resulted in hypothesized morphological synapomorphies for many of the clades reviewed here. In providing this information we make no judgment on the veracity of the characters but rely on dozens of studies that list morphological characters as diagnostic for the clades named and reviewed here.

A synonym is a name that has a spelling that is different from another name that refers to the same taxon (Art. 14.1). We differentiate three types of synonyms. Ambiguous synonyms are two names that are spelled differently for the same clade with the same taxa contained in that clade but were not given explicit phylogenetic definitions. Approximate synonyms are very close to the same clade and the content may slightly differ. Partial synonyms could be names for paraphyletic groups that exclude a part of the crown or other examples where some portion of the defined clade content is not included in the group delimited by the partial synonym.

The comments section provides space to discuss aspects of the phylogenetics or biology of a clade that merit highlighting. In addition, we attempt to list the earliest fossil occurrences of the clade and provide information on any molecular age estimate for the lineage.

The constituent lineages section provides a tabulation of all the major taxa comprising the defined clade. Any taxonomic families listed that are monotypic or monogeneric are marked with an asterisk as redundant group names. All names that are defined as clade names or listed in a protologue that have the suffix of -oidea, -idae, -inae, or -ini are valid family-group names under the International Code of Zoological Nomenclature (Van der Laan et al. 2014).

Our approach to constructing a rank-free classification of Actinopterygii necessitated a slight deviation from the principles and rules of the PhyloCode. While committed to maximizing the benefits of a classification that avoids redundant names, we have chosen a tempered approach that aims to accommodate traditional aspects of systematic ichthyology. Our classification is fully rank-free, but we use names with suffixes that include -formes, -oidei, and -oidea that are traditionally used for ranks of order, suborder, and superfamily. In avoiding the nesting of group names with the same suffixes, we maintain the exclusivity of those names, which requires replacement of the suffixes of several names in current usage. For example, we use Lophioidei and Tetraodontoidei in favor of Lophiiformes and Tetraodontiformes to avoid nesting these groups in Acanthuriformes. Because this is counter to PhyloCode's Principle 4 on stability, we do not use the PhyloCode in defining Gadoidei, Atheriniformes, Atherinoidei, Belonoidei, Cyprinodontoidei, Pleuronectoidei, Lophioidei, and Tetraodontoidei. We also do not use the PhyloCode in defining Salmoniformes, Esocidae, and Gadiformes because our delimitations of these groups would require the application of new group names.

In this classification we aim to preserve the nomenclatural history of actinopterygian systematics by retaining preexisting names for clades as much as possible. Among the 97 clade names in this classification, only one (Oseanacephala) is new and only seven other clade names date to the 21st century (Acropomatiformes, Apogonoidei, Cithariniformes, Eupercaria, Ovalentaria, Stomiatii, and Zoarcoidea). Forty-five of the group names were introduced from 1700 to 1900 CE, 19 names date from 1901 to 1950 CE, 25 group names were introduced between 1951 and 2000 CE, and 8 group names date from 2001 to 2022 CE. Seven of the 97 clade definitions were initially published in the PhyloCode companion volume (de Queiroz et al. 2020b; Lundberg 2020a, 2020b, 2020d; Moore and Near 2020a, 2020b, 2020c, 2020f) and are included here with any additional information to make the accounts uniform with the 90 new clade accounts.

Clade Accounts

Actinopterygii A. S. Woodward 1891:423
[J. A. Moore and T. J. Near 2020b]

  • Definition. Defined as a minimum-crown-clade by Moore and Near (2020b) as: “The least inclusive crown clade that contains Polypterus bichir Lacépède 1803, Acipenser sturio Linnaeus 1758, Psephurus gladius (Martens 1862), Lepisosteus osseus (Linnaeus 1758), Amia calva Linnaeus 1766, and Perca fluviatilis Linnaeus 1758.”

  • Etymology. From the ancient Greek ἀκτίς (̍æktIs), meaning ray or beam, and πτερὀν (t̍εɹαːn), meaning fin or wing.

  • Registration number. 206.

  • Reference phylogeny. Diogo (2007, figs. 3, 4) was designated as the primary reference phylogeny by Moore and Near (2020b). See Figures 2 and 3 for a summary phylogeny of major clades in Actinopterygii. The placement of †Scanilepiformes is supported in phylogenetic analyses of morphological characters (Giles et al. 2017, 2023; Latimer and Giles 2018).

  • Phylogenetics. The earliest phylogenetic investigations of Actinopterygii involved the secondary mapping of morphological character state changes onto tree topologies that placed Polypteridae (bichirs and ropefish) as the sister group of Actinopteri (e.g., Rosen et al. 1981; Patterson 1982; Lauder and Liem 1983; Gardiner 1984). The earliest phylogenetic analyses of morphological data matrices resolved Polypteridae as an actinopterygian and placed several Devonian fossil taxa (e.g., †Mimia, †Howqualepis, †Moythomasia, and †Kentuckia) as crown lineage Actinopterygii (Gardiner and Schaeffer 1989; Coates 1999; Gardiner et al. 2005; Xu, Gao, et al. 2014; Caron et al. 2023). The status of these Devonian taxa as crown clade actinopterygians was dramatically overturned by more recent morphological phylogenetic analyses that resolve numerous Devonian-Triassic taxa as stem lineage actinopterygians and place Polypteridae as nested within the Triassic-aged pan-scanilepiforms or as the sister group of †Scanilepiformes (Giles et al. 2017, 2023; Argyriou et al. 2018, 2022; Latimer and Giles 2018; Ren and Xu 2021). From the first molecular phylogenetic studies of ray-finned fishes to the most recent phylogenomic studies (e.g., Normark et al. 1991; Hughes et al. 2018), Actinopterygii is resolved as monophyletic with Polypteridae as the sister lineage of Actinopteri (Inoue et al. 2003a; Kikugawa et al. 2004; Alfaro, Santini, et al. 2009; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; M.-Y. Chen et al. 2015; Hughes et al. 2018; Vialle et al. 2018; Wcisel et al. 2020; Bi et al. 2021). In contrast to the consistent resolution of Polypteridae as the sister lineage of all other living Actinopterygii in molecular studies, some morphological phylogenetic analyses that include fossil taxa resolve a clade with low node support containing Polypteridae, †Scanilepiformes, pan-acipenseriforms, and Acipenseriformes (Argyriou et al. 2018; Latimer and Giles 2018; Caron et al. 2023; Giles et al. 2023).

  • Composition. Actinopterygii includes more than 35,085 living species (Fricke et al. 2023) classified in Polypteridae and Actinopteri. Fossil taxa within Actinopterygii include †Scanilepiformes (Appendix 1; Sytchevskaya 1999; Xu and Gao 2011; Giles et al. 2017). Appendix 1 provides details of the ages and locations of the fossil scanilepiforms. Over the past 10 years 3,657 new living species of Actinopterygii have been described (Fricke et al. 2023), comprising 10.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Actinopterygii include (1) cerebellum with corpus cerebelli, auricle, and valvula (Gardiner 1973; Løvtrup 1977:175), (2) teeth with apical cap of acrodin (Ørvig 1978; Rosen et al. 1981; Patterson 1982), (3) absence of superficial constrictors on gill arches (Wiley 1979), (4) presence of obliqui ventrales branchial muscle (Wiley 1979), (5) origin of coracomandibularis on branchial arch 3 (Wiley 1979), (6) adductor operculi continuous with adductor hyomandibulae (Lauder 1980), (7) adductor arcus palatini absent (Lauder 1980), (8) pelvic plate and two series of radials present (Patterson 1982), (9) anterodorsal process on scales (Patterson 1982), (10) a slender peg-and-socket articulation between scales (Patterson 1982), (11) autosphenotic ossified in postorbital process, autosphenotic and dermosphenotic fused (Patterson 1982), (12) single hyomandibular articulation above jugal canal (Patterson 1982), (13) postcleithrum present (Patterson 1982; Coates 1998), (14) prismatic ganoine on scales (Gardiner and Schaeffer 1989; Coates 1999), (15) three or more supraorbitals (Giles et al. 2017), (16) one or two infradentaries (Giles et al. 2017), (17) coronoid process of lower jaw present (Giles et al. 2017), (18) palatoquadrate with separate centers of ossification (Giles et al. 2017), (19) palate with flat dorsal margin (Giles et al. 2017), (20) narrow interorbital septum (Giles et al. 2017), (21) roof of posterior myodome perforated by palatine branch of facial nerve (Giles et al. 2017), (22) median posterior myodome present (Giles et al. 2017), (23) dermal component to basipterygoid process present (Giles et al. 2017), (24) parasphenoid extends to basioccipital (Giles et al. 2017), (25) ascending process of parasphenoid process present (Giles et al. 2017), (26) proximal segments of pectoral fin elongate with terminal segmentation (Giles et al. 2017); (27) proximal radials of dorsal fin enlarged (Giles et al. 2017); (28) constrictor mandibularis dorsalis attaches to the hyoid arch (Datovo and Rizzato 2018); and (29) constrictor mandibularis has an insertion on the lateral face of the palatoquadrate (Datovo and Rizzato 2018).

  • Synonyms. There are no synonyms of Actinopterygii.

  • Comments. Actinopterygii represents one of the major lineages of living vertebrates and along with Sarcopterygii comprises Osteichthyes (Rosen et al. 1981; Stiassny et al. 2004; Bertrand and Escrivá 2014; Moore and Near 2020d). When Actinopterygii was first introduced as a group name it excluded Polypteridae and had a composition identical to Actinopteri (Woodward 1891:423). Citing evidence from the morphology of scales, dermal bones of the head, the skull, nostrils, median fins, and paired fins and girdles, Goodrich (1928) considered Polypteridae as a group within Actinopterygii. By the 1980s the concept of Actinopterygii as comprising Polypteridae and Actinopteri was solidified in studies and reviews of morphological evidence (Rosen et al. 1981; Patterson 1982).

  • The earliest actinopterygian fossil taxon is †Platysomus superbus from the Visean (346.7–330.0 Ma) in the Carboniferous of Scotland, UK (C. D. Wilson et al. 2021). The inferred phylogenetic relationships of †Platysomus vary among morphological studies, but the taxon is consistently resolved as a lineage of Actinopterygii (Giles et al. 2017, 2023; Argyriou et al. 2018, 2022; Latimer and Giles 2018). Bayesian relaxed molecular clock age estimates for the crown age of Actinopterygii range between 333.5 and 384.1 million years ago, extending across the Devonian-Carboniferous boundary (Giles et al. 2017).

  • img-z11-7_03.gif

    FIGURE 2.

    Phylogenetic relationships of the major living lineages of Actinopterygii, Actinopteri, Neopterygii, Teleostei, Oseanacephala, Clupeocephala, Otocephala, Ostariophysi, Otophysi, Euteleostei, Salmoniformes, Stomiatii, Neoteleostei, Ctenosquamata, Acanthomorpha, Paracanthopterygii, Gadiformes, Acanthopterygii, Percomorpha, Ovalentaria, and Eupercaria. Filled circles identify the common ancestor of clades, with formal names defined in the clade accounts.

    img-z9-1_03.jpg

    FIGURE 3.

    Phylogenetic relationships of the major living lineages and fossil taxa of Actinopterygii, Actinopteri, Neopterygii, Pan-Teleostei, and Teleostei. Filled circles identify the common ancestor of clades, with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1. The clade description of Pan-Teleostei is presented in Moore and Near (2020e). The illustration of †Leptolepis coryphaenoides is reproduced with permission from Arratia (1996b).

    img-z10-1_03.jpg

    Polypteridae C. L. Bonaparte 1835
    [in Bonaparte 1840]:188–189
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive clade that contains Erpetoichthys calabaricus J. A. Smith 1865:2 and Polypterus bichir Lacepède 1803. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek πoυλύς (p̍uːlәs), meaning many, and πτερόν (tˈɛɹɑːn), meaning wing.

  • Registration number. 851.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of eight concatenated Sanger-sequenced nuclear genes (Near, Dornburg, Tokita, et al. 2014, fig. 1). A phylogeny of all species of Polypteridae is shown in Figure 4.

  • Phylogenetics. All species of Polypteridae are included in phylogenies inferred from mtDNA and Sanger-sequenced nuclear genes (Figure 4; Suzuki et al. 2010; Near, Dornburg, Tokita, et al. 2014).

  • Composition. There are currently 14 living species of Polypteridae that includes Erpetoichthys calabaricus and 13 species of Polypterus (Moritz and Britz 2019). Over the past 10 years no new living species of Polypteridae have been described (Fricke et al. 2023).

  • Diagnostic apomorphies. Morphological apomorphies for Polypteridae include (1) larvae with external gills that originate outside the branchial cavity (Daget 1950; Stundl et al. 2019), (2) single basibranchial (Jarvik 1980; Carvalho et al. 2013), (3) separate dorsal finlets (Daget 1950; Jarvik 1980; Gardiner and Schaeffer 1989; Coelho et al. 2018), (4) putative dorsal ribs (Britz and Bartsch 2003), (5) occipital bone that articulates posteriorly with centrum of second vertebra (Britz and Johnson 2010), (6) spiracular canal absent (Gardiner et al. 2005), (7) ascending process of parasphenoid fused to otic region and not related to spiracle (Gardiner et al. 2005), (8) parasphenoid with aortic canal (Gardiner et al. 2005), (9) parietals absent, dermopterotics meet (Gardiner et al. 2005), (10) maxilla with superimposed infraorbital canal and dorsal arm of preopercular greatly expanded (Gardiner et al. 2005), (11) coronoid process of lower jaw composed exclusively of prearticular (Gardiner et al. 2005; Giles et al. 2017), (12) optic foramen adjacent to dorsal margin of parasphenoid (Giles et al. 2017), (13) broad interorbital septum (Giles et al. 2017), (14) lateral process present on ectopterygoid (L. Grande 2010; Giles et al. 2017), (15) four ceratobranchials (Britz and Johnson 2003; Giles et al. 2017), (16) loss of fulcra of caudal fin (Patterson 1982; Giles et al. 2017), (17) three pairs of extrascapulars (Gardiner and Schaeffer 1989; Giles et al. 2017), and (18) constrictor mandibularis dorsalis, levator arcus palatini is differentiated into partes interna and externa (Datovo and Rizzato 2018).

  • Synonyms. Brachiopterygii (G. J. Nelson 1969a, fig. 25), Cladistia (Rosen et al. 1981, fig. 62; Betancur-R et al. 2017:9), and Polypteriformes (J. S. Nelson et al. 2016:116; Betancur-R et al. 2017:9) are ambiguous synonyms of Polypteridae.

  • Comments. Bonaparte (1840) applied the group name Polypterini as a subfamily of Lepidosteidae, which is a synonym of Lepisosteidae. The delimitation of Polypteridae as containing Polypterus and Erpetoichthys calabaricus presented here was frequently used by ichthyologists in the second half of the 19th though the early 20th century (Günther 1870:326–331, 1880:364; Bridge 1904:481–485; Boulenger 1909:4; Goodrich 1909:300). We selected the name Polypteridae as the clade name over its synonyms because it is the name most frequently applied to a taxon approximating the named clade. Polypteridae is the living sister lineage of all other actinopterygians (Actinopteri) and results from relaxed molecular clock analyses that estimate the common ancestry of these two lineages dates to an interval between 333.5 and 384.1 million years ago (Giles et al. 2017).

  • In contrast to the ancient divergence of Polypteridae and Actinopteri, the earliest pan-polypterid fossils date to the Cenomanian (100.5–93.9 Ma) of the Upper Cretaceous (Daget et al. 2001; Gayet et al. 2002; Near, Dornburg, Tokita, et al. 2014), implying a gap in the fossil record of polypterids that spans at least 240 million years. All extant species of Polypteridae live in the freshwaters of western and central Africa, although pan-polypterid fossils are known from both Africa and South America (Gayet and Meunier 1991, 1992; Meunier and Gayet 1996; Daget et al. 2001; Otero et al. 2009). Time-calibrated multi-species coalescent analyses estimate a relatively recent time to common ancestry for the living species of Polypteridae, spanning the Miocene and early Oligocene between 13.6 and 24.9 million years ago (Near, Dornburg, Tokita, et al. 2014). Polypteridae is a valid family-group name under the International Code of Zoological Nomenclature (Van der Laan et al. 2014:27).

  • img-z13-6_03.gif

    FIGURE 4.

    Phylogenetic relationships of the species of Polypteridae. Filled circles identify the common ancestor of clades, with formal names defined in the clade accounts.

    img-z12-1_03.jpg

    Actinopteri E. D. Cope 1871:587
    [J. A. Moore and T. J. Near 2020]

  • Definition. Defined as a minimum-crown-clade by Moore and Near (2020a) as: “The least inclusive crown clade that contains Acipenser sturio Linnaeus 1758, Psephurus gladius (Martens 1862), Lepisosteus osseus (Linnaeus 1758), Amia calva Linnaeus 1766, and Perca fluviatilis Linnaeus 1758.”

  • Etymology. From the ancient Greek πουλύς (̍æktIs), meaning ray or beam, and πτερόν (tˈɛɹɑːn), meaning fin or wing.

  • Registration number. 208.

  • Reference phylogeny. Diogo (2007, figs. 3, 4) was designated as the primary reference phylogeny by Moore and Near (2020a). See Figures 2 and 3 for summary phylogenies of major clades in Actinopteri. The placements of the stem-acipenseriforms †Pycnodontiformes, †Guildayichthyidae, †Bobasatraniidae, †Australosomus, †Redfieldiidae, †Platysiagidae, †Dipteronotus, †Peltopleuridae, †Thoracopteridae, †Venusichthys, and †Habroichthys are on the basis of phylogenetic analyses of morphological characters (L. Grande and Bemis 1991, 1996; Bemis et al. 1997; Lund 2000; Hilton and Forey 2009; Mickle et al. 2009; Hilton et al. 2011; Xu et al. 2012; Poyato-Ariza 2015; Xu and Ma 2016; Xu and Zhao 2016; Giles et al. 2017, 2023; Xu 2021; Shedko 2022; Yuan et al. 2022).

  • Phylogenetics. The earliest phylogenetic investigations of Actinopteri involved the secondary mapping of morphological character state changes onto tree topologies that placed chondrosteans (Acipenseriformes) and Neopterygii (Holostei and Teleostei) as sister lineages (Rosen et al. 1981; Patterson 1982; Lauder and Liem 1983; Gardiner 1984). Phylogenetic analyses of morphological data matrices corroborate the monophyly of Actinopteri (Coates 1999; Gardiner et al. 2005; Xu, Gao, et al. 2014; Poyato-Ariza 2015; Giles et al. 2017; Latimer and Giles 2018). Several molecular studies ranging from analyses of whole mtDNA genomes, samples of Sanger-sequenced nuclear genes, and phylogenomic analyses resolve Actinopteri as a monophyletic group (Inoue et al. 2003a; Kikugawa et al. 2004; Alfaro, Santini, et al. 2009; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; M.-Y. Chen et al. 2015; Hughes et al. 2018; Vialle et al. 2018; Wcisel et al. 2020; Bi et al. 2021; Mu et al. 2022).

  • Composition. Actinopteri includes 35,075 living species classified in the subclades Acipenseriformes and Neopterygii. Fossil taxa within Actinopteri include the pan-acipenseriforms †Boreosomus, †Chondrosteus, and †Peipiaosteus; and the pan-neopterygians †Australosomus, †Bobasatraniidae, †Dipteronotus, †Guildayichthyidae, †Habroichthys, †Peltopleuridae, †Platysiagidae, †Pycnodontiformes, †Redfieldiidae, †Thoracopteridae, and †Venusichthys. Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years 3,675 new living species of Actinopteri have been described (Fricke et al. 2023), comprising 10.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Actinopteri include (1) perforated propterygium (Patterson 1982), (2) bases of marginal rays or pectoral fin embracing propterygium (Patterson 1982), (3) basal fulcra on dorsal margin of caudal fin (Patterson 1982), (4) fringing fulcra on median fins (Patterson 1982), (5) supra-angular bone present on lower jaw (G. J. Nelson 1973; Patterson 1982), (6) presence of spiracular canal in braincase (Patterson 1982), (7) swim bladder with dorsal connection to foregut (Patterson 1982), (8) hemopoietic organ above medulla (Patterson 1982), (9) diffuse pancreas (Patterson 1982), (10) olfactory rosette (Patterson 1982), (11) supratemporal fused with intertemporal forming dermopteric (Coates 1999), (12) fewer than 12 or 13 branchiostegal rays or plates (Coates 1999), (13) posterior parasphenoid expanded to cover ventral otic fissure (Coates 1999), (14) post-temporal fossa (Xu, Gao, et al. 2014), (15) basipterygoid process absent (Xu, Gao, et al. 2014), (16) quadratojugal overlaying quadrate (Xu, Gao, et al. 2014), (17) loss of presupracleithrum (Xu, Gao, et al. 2014), (18) dorsal aorta open in groove (Giles et al. 2017), (19) cerebellar corpus undivided (Giles et al. 2017), (20) cerebellar corpus with median anterior projecting portion (Giles et al. 2017), (21) hour-glass shaped medial constriction of anterior ossification of ceratohyal (Giles et al. 2017), and (22) uncinate processes on epibranchials (Giles et al. 2017).

  • Synonyms. Actinopterygii as delimited in Boulenger (1891:10), Woodward (1891:423), Dean (1895:8), McAllister (1968:18–20), G. J. Nelson (1969a:534), J. S. Nelson (1976:58; 1984:77–78) Løvtrup (1977:170–176), and Forey (1980:378) excluded Polypteridae and is therefore an approximate synonym of Actinopteri. Garstang (1931:255–256) introduced the group name Epipneusta, containing Acipenseriformes, Holostei, and Teleostei; Epipneusta is an approximate synonym of Actinopteri.

  • Comments. Cope's (1871b) first delimitation of Actinopteri included Acipenseriformes, Holostei, and Teleostei and is identical to the composition of the clade described here. Later Cope (1877b) modified Actinopteri to include only Holostei and Teleostei, but subsequent classifications used Cope's (1871b) initial concept of Actinopteri to include Acipenseriformes, Holostei, and Teleostei (Jordan 1905; Gregory 1907). After the application of phylogenetic systematics to the study of ray-finned fishes, Actinopteri was reintroduced to include all living actinopterygians except Polypteridae (Patterson 1982). The earliest fossils of Actinopteri are the pan-neopterygian guildayichthyids †Discoserra pectinodon and †Guildayichthys carnegiei from the Serpukhovian (330.3–323.4 Ma) in the Carboniferous of Montana, USA (Lund 2000; Mickle et al. 2009). Bayesian relaxed molecular clock analyses estimate a crown age of Actinopteri between 309 and 357 million years ago (Giles et al. 2017).

  • img-z15-2_03.gif

    Acipenseriformes L. S. Berg 1940:408–409
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Acipenser sturio Linnaeus 1758 and Polyodon spathula (Walbaum 1792). This is a minimum-crown-clade definition.

  • Etymology. Acipenser is the Latin name for sturgeon, which is derived from the ancient Greekἀκκιπἠσιoς (D. W. Thompson 1947). The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 879.

  • Reference phylogeny. A phylogeny inferred from a dataset of combined molecular and morphological characters (Shedko 2022, fig. 1). See Figures 2 and 3 for the relationship of Acipenseriformes among the major lineages of Actinopterygii. See Figure 5A for a summary phylogeny of the major lineages of Acipenseriformes. Placements of the fossil acipenseriform taxa in the phylogeny are based on the results of morphological phylogenetic analyses (L. Grande and Bemis 1991, 1996; Bemis et al. 1997; Hilton and Forey 2009; Hilton et al. 2011).

  • Phylogenetics. The earliest phylogenetic trees that show a monophyletic Acipenseriformes were inferred from a distribution of derived character states without an analysis of a coded character data matrix using an explicit optimality criterion (e.g., G. J. Nelson 1969a; Lauder and Liem 1983). Morphological and molecular phylogenies consistently resolve Acipenseriformes as monophyletic (e.g., L. Grande and Bemis 1996; Inoue et al. 2003a; Artyukhin 2006; Alfaro, Santini, et al. 2009; Hilton and Forey 2009; Broughton 2010; Hilton et al. 2011; Near, Eytan, et al. 2012; Giles et al. 2017, 2023; Hughes et al. 2018; Y. Shen et al. 2020; Shedko 2022).

  • Composition. Acipenseriformes includes 28 living species (Fricke et al. 2023) classified in Acipenseridae and Polyodontidae. The number of living species does not include the recently declared extinct Chinese Paddlefish, Psephurus gladius (H. Zhang et al. 2020). The fossil taxa †Protopsephurus and †Paleopsephurus are resolved in morphological phylogenies as pan-polyodontids, that is, outside of the crown clade Polyodontidae (L. Grande and Bemis 1991, 1996). †Priscosturion is resolved as either a pan-acipenserid or nested within Acipenseridae as the sister lineage of Scaphirhynchus (L. Grande and Hilton 2006; Hilton et al. 2011; Shedko 2022; Murray et al. 2023). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years no new living species of Acipenseriformes have been described (Fricke et al. 2023).

  • Diagnostic apomorphies. Morphological apomorphies for Acipenseriformes include (1) loss of opercle (L. Grande and Bemis 1996; Bemis et al. 1997), (2) fewer than four branchiostegal rays (L. Grande and Bemis 1996; Bemis et al. 1997), (3) endocranium with extensive rostrum (L. Grande and Bemis 1996; Bemis et al. 1997), (4) dorsal and ventral rostral bones (L. Grande and Bemis 1996; Bemis et al. 1997; Hilton et al. 2011), (5) posttemporal bone with a ventral process (L. Grande and Bemis 1996; Bemis et al. 1997), and (6) absence of constrictor mandibularis dorsalis connection to palatoquadrate (Datovo and Rizzato 2018).

  • Synonyms. Acipenseroidei (L. Grande and Bemis 1991:113, 1996:107; Bemis et al. 1997:51–53) is an ambiguous synonym of Acipenseriformes.

  • Comments. Berg (1940) originally included Acipenseridae, Polyodontidae, and †Chondrostei in Acipenseriformes, which we selected as the clade name over its synonyms because it is the name most frequently applied to a taxon approximating the named clade. Morphological phylogenies resolve †Chondrostei and †Peipiaosteidae as pan-acipenseriforms and are not included in Acipenseriformes as delimited here (L. Grande and Bemis 1991, 1996; Hilton and Forey 2009; Hilton et al. 2011). The earliest fossil Acipenseriformes is the pan-polyodontid †Protopsephurus liui from the Barremian (126.5–121.4 Ma) in the Cretaceous of China (Appendix 1). Bayesian relaxed molecular clock analyses of Acipenseriformes result in an average posterior crown age estimate of 126.8 million years ago, with the credible interval ranging between 120.9 and 144.5 million years ago (Hughes et al. 2018).

  • img-z17-2_03.gif

    FIGURE 5.

    Phylogenetic relationships of the major living lineages and fossil taxa of (A) Acipenseriformes and (B) Holostei. Filled circles identify the common ancestor of clades, with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z16-1_03.jpg

    Neopterygii C. T. Regan 1923b:458
    [J. A. Moore and T. J. Near 2020]

  • Definition. Defined as a minimum-crown-clade by Moore and Near (2020c) as: “The least inclusive crown clade containing Lepisosteus osseus (Linnaeus 1758), Amia calva Linnaeus 1766, and Perca fluviatilis Linnaeus 1758.”

  • Etymology. From the ancient Greek νἐος (n̍iːo͡Ʊz), meaning new, and πτερὀν (t̍εɹαːn), meaning fin or wing.

  • Registration number. 210.

  • Reference phylogeny. Diogo (2007, figs. 3, 4) was designated as the primary reference phylogeny by Moore and Near (2020c). See Figures 2 and 3 for summary phylogenies of the major clades in Neopterygii. The placements of the pan-holosteans and Pan-Teleostei fossil taxa in the phylogeny are on the basis of inferences from analyses of morphological characters (Patterson 1977; Patterson and Rosen 1977; Arratia 1991, 1997, 1999, 2000a, 2001, 2008, 2013, 2016, 2017; Arratia and Thies 2001; Arratia and Tischlinger 2010; Taverne 2013; Sferco et al. 2015; Giles et al. 2017, 2023; Latimer and Giles 2018; Bean and Arratia 2020; Veysey et al. 2020; Arratia et al. 2021; Bean 2021; C. Shen and Arratia 2021).

  • Phylogenetics. The earliest phylogenetic investigations of Neopterygii resulted in tree topologies that depicted Holostei (Lepisosteidae and Amia) plus Teleostei as a monophyletic group (e.g., G. J. Nelson 1969a; Patterson 1973; Wiley 1976; Lauder and Liem 1983; Wiley and Schultze 1984; Maisey 1986). Phylogenetic analyses of morphological data matrices resolve Neopterygii as monophyletic (Olsen 1984; Olsen and McCune 1991; Gardiner et al. 1996; Coates 1999; Cavin and Suteethorn 2006; Hurley et al. 2007; Arratia and Tischlinger 2010; L. Grande 2010; Xu and Gao 2011; Xu and Wu 2012; Xu et al. 2012; Arratia 1999, 2013; Xu, Gao, et al. 2014; Poyato-Ariza 2015; Xu and Ma 2016; Xu and Zhao 2016; Giles et al. 2017, 2023; Argyriou et al. 2018, 2022; Latimer and Giles 2018; López-Arbarello and Sferco 2018; Ren and Xu 2021; Xu 2021; Mu et al. 2022; Yuan et al. 2022).

  • One of the earliest molecular phylogenetic studies of ray-finned fishes used DNA sequences of small fragments of three mtDNA protein coding genes and failed to resolve Neopterygii as monophyletic (Normark et al. 1991). Phylogenetic analysis of complete mtDNA genome sequences strongly resolves Neopterygii as paraphyletic, with Acipenseriformes and Holostei as sister lineages (Inoue et al. 2003a). Molecular phylogenetic analyses of Sanger-sequenced nuclear genes, combinations of nuclear and mtDNA genes, and phylogenomic studies all resolve Neopterygii as monophyletic (Kikugawa et al. 2004; Alfaro, Santini, et al. 2009; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Broughton et al. 2013; M.-Y. Chen et al. 2015; Betancur-R et al. 2017; Hughes et al. 2018; Vialle et al. 2018; Wcisel et al. 2020).

  • Composition. Neopterygii includes more than 35,045 living species classified in Holostei and Teleostei (Near, Eytan, et al. 2012; Fricke et al. 2023). Fossil neopterygian lineages classified as Pan-Teleostei include †Ankylophoriformes, †Ascalabos, †Aspidorhynchidae, †Atacamichthys, †Catervariolus, †Dorsetichthys, †Ichthyodectiformes, †Ichthyokentema, †Leptolepis, †Pachycormidae, †Pholidophoridae, †Prohalecites, †Tharsis, and †Varasichthyidae (Patterson and Rosen 1977; J. Gaudant 1978; Arratia 1981; Arratia and Tischlinger 2010; Taverne 2011b, 2013; Arratia 2013; Arratia and Schultze 2024). Fossil lineages of pan-holostean neopterygians include †Dapediidae and †Hulettia (Schaeffer and Patterson 1984; Latimer and Giles 2018; López-Arbarello and Sferco 2018). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years 3,657 new living species of Neopterygii have been described (Fricke et al. 2023), comprising approximately 10.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Neopterygii include (1) number of fin-rays in the anal and dorsal fins equal in number to endoskeletal supports (Patterson 1973; Patterson and Rosen 1977; Lauder and Liem 1983), (2) premaxilla with interior process that lines front part of nasal pit (Patterson 1973; Patterson and Rosen 1977), (3) vomer attached to underside of ethmoid (Patterson 1973), (4) coronoid process on articular (Patterson 1973; Patterson and Rosen 1977), (5) vertically oriented suspensorium (Patterson 1973), (6) dorsal limb of preopercle narrow (Patterson 1973), (7) symplectic present and is an outgrowth of the hyomandibular cartilage (Patterson 1973; Patterson and Rosen 1977), (8) enhanced upper pharyngeal dentition (Patterson 1973; Patterson and Rosen 1977; Lauder and Liem 1983), (9) clavicle lost or reduced to small plate lateral to cleithrum (Patterson 1973; Wiley 1976; Patterson and Rosen 1977; Lauder and Liem 1983), (10) basipterygoid process entirely composed of parasphenoid (Wiley 1976), (11) posterior commissure between the supraorbital and infraorbital canals (Wiley 1976), (12) uncinate process on first and second infrapharyngobranchials (Wiley 1976), (13) infrabranchials laterally supported (Wiley 1976), (14) differentiated dorsal gill arch musculature (Wiley 1976), (15) four basibranchial copulae (Wiley 1976), (16) quadratojugal braces the quadrate (Gardiner 1984), (17) antorbitals present (Gardiner 1984), (18) palatoquadrate disconnected from dermal cheek bones dorsally and posteriorly (Gardiner 1984), (19) hyoid facet directed posteroventrally (Gardiner and Schaeffer 1989; Coates 1999), (20) maxilla elongate and shallow (L. Grande and Bemis 1998; Hurley et al. 2007; López-Arbarello and Sferco 2018), (21) maxilla detached from preopercle (Gardiner and Schaeffer 1989; Xu, Gao, et al. 2014; López-Arbarello and Sferco 2018), (22) upper-most hypaxial caudal rays with a bundle of elongate fin ray bases that extend over several hypurals (Gardiner et al. 1996; Hurley et al. 2007), (23) ventral cranial-otic fissure closed by bone (Coates 1999), (24) canal for dorsal aorta secondarily absent (Coates 1999), (25) cerebellar corpus arches above fourth ventricle (Coates 1999), (26) presence of one or more accessory postcleithra (Arratia 1999; Hurley et al. 2007), (27) rostral-postrostral and frontal contact wholly or partially separating nasal bones (Xu, Gao, et al. 2014), (28) nasal process on premaxilla (Xu, Gao, et al. 2014), (29) four or more infraorbitals between antorbital and dermosphenotic (Xu, Gao, et al. 2014), (30) presence of mobile maxilla in cheek (Xu, Gao, et al. 2014), (31) interopercle present (Xu, Gao, et al. 2014; López-Arbarello and Sferco 2018), (32) presence of medial gular bones (Xu, Gao, et al. 2014), (33) presence of peg-like anterior process of maxilla (Xu, Gao, et al. 2014), (34) infraorbitals and suborbitals broadly overlap preopercle (Giles et al. 2017), (35) postrostral bone absent (López-Arbarello and Sferco 2018), (36) supramaxilla present (López-Arbarello and Sferco 2018), (37) subopercle with ascending process (López-Arbarello and Sferco 2018), (38) absence of a distinguishable spiracularis of the constrictor mandibularis (Datovo and Rizzato 2018), (39) posterior end of maxilla ends behind orbit (Xu 2021), and (40) posttemporals broad, nearly as wide as extrascapular (Xu 2021).

  • Synonyms. Cope's (1877b:293–294) revised definition of Actinopteri, Regan's (1904b:331–332, 1909b:76–82) delimitation of Teleostei, and Goodrich's (1930:xvii) composition of Holostei were limited to Amia calva, Lepisosteidae, and Teleostei and are all approximate synonyms of Neopterygii (Moore and Near 2020c). While not a formal taxonomic name, the term “crown Neopterygii” is an ambiguous synonym of Neopterygii.

  • Comments. In a study of the morphology of Lepisosteidae, Regan (1923b:458) justified the use of a new group name (italics added to clade names): “Holostei and Teleostei, therefore are one group, for which it seems better to use the name Neopterygii, rather than to use Holostei or Teleostei in a new and extended sense.” Among the earliest studies of actinopterygian relationships after the introduction of phylogenetic systematics, Neopterygii is resolved as the clade containing Holostei and Teleostei (G. J. Nelson 1969a; Patterson 1973) and was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade. The oldest fossil taxon of Neopterygii is the pan-amiiform †Watsonulus eugnathoides from the Induan (251.9–249.9 Ma) in the Triassic of Madagascar (Olsen 1984; Giles et al. 2017, 2023). The crown age of Neopterygii estimated from Bayesian relaxed molecular clock analyses ranges from the Permian to the Carboniferous between 278 and 318 million years ago (Giles et al. 2017).

  • img-z19-3_03.gif

    Holostei J. Müller 1845a:420
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Amia calva Linnaeus 1766 and Lepisosteus osseus (Linnaeus 1758), but not Perca fluviatilis Linnaeus 1758. This is a minimum-crown-clade definition with an external specifier.

  • Etymology. From the ancient Greek ὃλoς (h̍o͡Ʊlo͡Ʊz), meaning whole, entire, or complete, and ὀστέoν (̍αːstIәn), meaning bone.

  • Registration number. 881.

  • Reference phylogeny. A phylogeny inferred from concatenated DNA sequences of 1,105 exons (Hughes et al. 2018, fig. S2). See Figures 2 and 3 for the relationship of Holostei among the major lineages of Actinopterygii. Phylogenetic relationships among the lineages of Holostei are shown in Figure 5B. Placements of the fossil holostean taxa in the phylogeny are on the basis of inferences from morphological analyses (Olsen 1984; Schultze and Wiley 1984; Lambers 1995; Gardiner et al. 1996; Wenz 1999; Xu and Gao 2011; López-Arbarello 2012; Xu and Wu 2012; Xu et al. 2012, 2018; Cavin et al. 2013; López-Arbarello et al. 2014, 2019, 2020; Xu. Zhao, et al. 2014; Poyato-Ariza 2015; Xu and Shen 2015; Brito et al. 2017; Sun et al. 2017; Ebert 2018; Latimer and Giles 2018; López-Arbarello and Sferco 2018; Ren and Xu 2021; Brownstein 2022; Brownstein and Lyson 2022; Brownstein et al. 2023).

  • Phylogenetics. The monophyly of Holostei was supported in one of the earliest phylogenetic systematic perspectives on the relationships of vertebrates (G. J. Nelson 1969a), reflecting pre-cladistic hypotheses that grouped Amia and Lepisosteidae (Regan 1923b; Goodrich 1930). An assessment of skeletal morphology led to the conclusion that Holostei is paraphyletic, with Amia calva as the sister lineage of Teleostei (Patterson 1973). Nearly every molecular phylogenetic analysis from the earliest efforts using partial-gene DNA sequences to phylogenomic analyses resolves Holostei as monophyletic (Normark et al. 1991; Inoue et al. 2003a; Broughton 2010; Near, Eytan, et al. 2012; Faircloth et al. 2013; Braasch et al. 2016; Hughes et al. 2018; Mu et al. 2022). In addition, a phylogenetic analysis of 70 morphological character state changes resolved Holostei as monophyletic (Hurley et al. 2007). A critical examination of morphology in bowfin, gars, teleosts, and several fossil lineages demonstrated that nearly all of the proposed characters supporting the hypothesis that Amia and teleosts share common ancestry are also present in gars (L. Grande 2010). Subsequent morphological phylogenetic analyses consistently resolve Holostei as monophyletic (Hurley et al. 2007; L. Grande 2010; Xu and Gao 2011; Xu and Wu 2012; Xu et al. 2012; Cavin et al. 2013; Xu, Gao, et al. 2014; Xu, Zhao, et al. 2014; Poyato-Ariza 2015; Xu and Shen 2015; Xu and Ma 2016; Xu and Zhao 2016; Giles et al. 2017, 2023; Sun et al. 2017; Argyriou et al. 2018, 2022; Latimer and Giles 2018; López-Arbarello and Sferco 2018; Xu et al. 2018, 2019; López-Arbarello et al. 2020; Ren and Xu 2021; Xu 2021; Yuan et al. 2022; Feng et al. 2023). As such, Holostei exemplifies one of the first conflicts in ichthyological systematics between morphological and molecular phylogenetic analyses (Patterson 1994:65–70) that was reconciled through continued morphological and genomic phylogenetic studies, in this case offering overwhelming support for the monophyly of Holostei (e.g., Hurley et al. 2007; L. Grande 2010; Near, Eytan, et al. 2012; Hughes et al. 2018; López-Arbarello and Sferco 2018; A. W. Thompson et al. 2021).

  • Composition. There are nine living species of Holostei, two species of Amia, and seven species of Lepisosteidae (Suttkus 1963; L. Grande 2010; Brownstein et al. 2022). There are several extinct pan-amiiform taxa that include †Amiopsis, †Caturus, †Cyclurus, †Ionoscopus, †Panxianichthys, †Sinamia, †Solnhofenamia, †Vidalamia, and †Watsonulus. Extinct pan-lepisosteiform lineages include †Araripelepidotes, †Cuneatus, †Fuyuanichthys, †Lepidotes, †Macrosemius, †Masillosteus, †Nhanulepisosteus, †Obaichthyidae, †Pliodetes, †Semionotus, †Thaiichthys, and †Ticinolepis (L. Grande and Bemis 1998; Wenz 1999; L. Grande 2010; López-Arbarello 2012; Xu and Wu 2012; Cavin et al. 2013; Xu, Zhao, et al. 2014; Xu and Shen 2015; López-Arbarello et al. 2016; Brito et al. 2017). Details of the ages and locations of fossil holosteans are presented in Appendix 1. Over the past 10 years no new living species of Holostei have been described (Fricke et al. 2023), but one species was elevated from synonymy with Amia calva (Brownstein et al. 2022).

  • Diagnostic apomorphies. Morphological apomorphies for Holostei include (1) posterior extent of median rostral bone in adults reduced (L. Grande 2010), (2) anterior arm on antorbital with a tube-like canal (L. Grande 2010; Xu et al. 2018), (3) adults with two vertebral centra fused into the occipital condyle (L. Grande 2010), (4) pterotic bone absent (L. Grande 2010), (5) adults with paired vomer (L. Grande 2010), (6) coronoid process of mandibula involves more than one bone (L. Grande 2010), (7) supraangular bone present (L. Grande 2010), (8) caudal region with both paired and median neural spines (L. Grande 2010), (9) normally all primary rays in caudal fin branched (L. Grande 2010), (10) fringing fulcra present on upper and lower margins of caudal fin (L. Grande 2010), (11) presence of anterior and posterior clavicle elements (L. Grande 2010), (12) four hypobranchials present (L. Grande 2010), (13) long nasal process that is tightly sutured to the frontals attaches immovable premaxilla to braincase (L. Grande 2010; Xu et al. 2018), (14) anterior portion of premaxilla pierced by olfactory foramen and lies in nasal pit (L. Grande 2010), (15) sphenotic with dermal component (L. Grande 2010; Xu et al. 2018), and (16) presence of a larval attachment organ that is a compound super-organ located at the front of the snout (Pinion et al. 2023).

  • Synonyms. There are no synonyms of Holostei.

  • Comments. Müller (1845a) delimited Holostei as including Polypteridae and Lepisosteidae. Later definitions of Holostei limited the group to Amia calva, pan-amiiforms, pan-lepisosteiforms, and Lepisosteidae (Regan 1923b; Goodrich 1930; L. Grande 2010). The alternative phylogenetic hypothesis that A. calva and Teleostei are sister lineages to the exclusion of Lepisosteidae was introduced by Patterson (1973). If future phylogenetic analyses find support for this hypothesis, the use of an external specifier in the clade definition would render Holostei inapplicable and Halecostomi would be an appropriate name for the smallest clade containing A. calva and Teleostei, but not Lepisosteidae. We were motivated to include an external specifier because the situation with Holostei and Halecostomi was used as an example in the Phylonyms volume of how to create a definition that will make a name inapplicable in the context of some phylogenies (de Queiroz et al. 2020a:xxvii). The earliest holostean is the pan-amiiform †Watsonulus eugnathoides from the Induan (251.9–249.9 Ma) in the Triassic of Madagascar (Olsen 1984; Giles et al. 2017, 2023). Fossil-calibrated Bayesian relaxed molecular clock analyses place the crown age of Holostei between 248 and 312 million years ago (Near, Eytan, et al. 2012, tbl. S1), which spans the Lower Triassic, Permian, and Upper Pennsylvanian (Carboniferous).

  • img-z21-2_03.gif

    Teleostei J. Müller 1845b:129
    [J. A. Moore and T. J. Near 2020]

  • Definition. Defined as a minimum-crown-clade by Moore and Near (2020f) as: “The least inclusive crown clade that contains Hiodon tergisus Lesueur 1818 (Osteoglossomorpha), Elops saurus Linnaeus 1766 (Elopomorpha), Engraulis encrasicolus (Linnaeus 1758) (Otocephala/ Clupeomorpha), and Perca fluviatilis Linnaeus 1758 (Euteleostei).”

  • Etymology. From the ancient Greek τέλειoς (t̍εlƗᵻo͡Ʊz), meaning perfect or complete, and ὀστέoν (̍αːstIәn), meaning bone.

  • Registration number. 212.

  • Reference phylogeny. Diogo (2007, figs. 3, 4) was designated as the primary reference phylogeny by Moore and Near (2020f). See Figures 2 and 3 for summary phylogenies of major clades that comprise Teleostei. The phylogenetic placement of the fossil taxon †Tselfatiiformes in Figure 3 is on the basis of analysis of morphological characters (Cavin 2001).

  • Phylogenetics. The first phylogenetic analyses supporting the monophyly of Teleostei were inferred from a distribution of derived character states without an analysis of a coded character data matrix using an explicit optimality criterion (Patterson 1977; Patterson and Rosen 1977). Subsequent phylogenetic analyses of morphological characters consistently resulted in teleost monophyly (Arratia 1991, 1997, 1999, 2001, 2008, 2013, 2017; Diogo 2007; Gouiric-Cavalli and Arratia 2022). Many of these morphological studies did not include a broad sampling of teleost diversity as they were aimed at resolving relationships among Teleostei and stem lineages that comprise the more inclusive Pan-Teleostei (Moore and Near 2020e).

  • The earliest molecular phylogenetic studies of ray-finned fishes used DNA sequences from small fragments of mtDNA and nuclear ribosomal RNA genes and did not resolve teleosts as monophyletic or did so with low node support (Normark et al. 1991; Lê et al. 1993). Starting in the early 21st century, molecular phylogenetic analyses ranging from the use of whole mtDNA genomes to phylogenomic analyses consistently resolve Teleostei as monophyletic (Inoue et al. 2003a; Hurley et al. 2007; Alfaro, Santini, et al. 2009; Broughton 2010; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Austin et al. 2015; M.-Y. Chen et al. 2015; Bian et al. 2016; Betancur-R et al. 2017; Hughes et al. 2018; Vialle et al. 2018; Roth et al. 2020; Wcisel et al. 2020; Mu et al. 2022; Parey et al. 2023).

  • Composition. Teleostei contains more than 35,035 living species (Fricke et al. 2023) classified in Oseanacephala and Clupeocephala. Fossil teleosts include the †Tselfatiiformes (Cavin 2001). Details of the age and location of the fossil tselfatiiform taxon are presented in Appendix 1. Over the past 10 years 3,657 new living species of Teleostei have been described (Fricke et al. 2023), composing 10.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Teleostei include (1) presence of endoskeletal basihyal (G. J. Nelson 1969a; Patterson 1977; Arratia 1999, 2000c, 2008), (2) absence of a structure on ventral surface of basioccipital for cranial attachment of aortic ligament (Patterson 1975; Patterson and Rosen 1977; De Pinna 1996), (3) three hypobranchials (Patterson 1977), (4) four pharyngobranchials (Patterson 1977; De Pinna 1996), (5) seven hypurals in caudal skeleton (Patterson 1977; Patterson and Rosen 1977; De Pinna 1996), (6) base of fin rays on upper lobe of caudal fin attaches to or overlies no more than one hypural (Patterson and Rosen 1977; Arratia 1996b, 1997), (7) craniotemporal muscle present (Stiassny 1986; De Pinna 1996; Arratia 1999, 2000c; Wiley and Johnson 2010), (8) hypurals 1 and 2 laterally fused in adults (Arratia 1991), (9) absence of dorsal processes of the bases of the innermost primary caudal rays of upper lobe (Arratia 1991, 1996b, 2000c, 2008), (10) lateral forebrain bundle composed of myelinated fibers (De Pinna 1996; Wiley and Johnson 2010), (11) presence of accessory nasal sacs (X. Y. Chen and Arratia 1994; De Pinna 1996; Arratia 1999, 2000c, 2008; Wiley and Johnson 2010), (12) hyoidean artery pierces either both hypohyals or ventral hypohyal (Arratia 1999, 2000c, 2008), (13) pharyngobranchials with three ossified elements and a tooth plate-bearing cartilaginous element (Arratia 1999, 2000c, 2008; Wiley and Johnson 2010), (14) five or fewer ural neural arches modified as uroneurals (Arratia 1999, 2008; Wiley and Johnson 2010), (15) absence of notch in deep dorsal ascending margin of dentary (Arratia 2008, 2013, 2017), (16) many developed epipleural intermuscular bones in abdominal and caudal region (Arratia 2008), (17) parahypural haemal arch in adults not fused laterally to autocentrum (Arratia 2008), (18) uroneural 1 reaches anterior to preural centrum 2 (Wiley and Johnson 2010), (19) presence of an independent endoskeletal basihyal (Wiley and Johnson 2010), (20) absence of segmentum buccalis of adductor mandibulae (Datovo and Rizzato 2018), (21) presence of dilatator process on opercle (Datovo and Rizzato 2018), (22) presence of adductor crest (Datovo and Rizzato 2018), and (23) autocentrum of vertebrae with thickened lateral wall and series of ornaments including crests, grooves, and pits (Arratia 1997, 1999, 2013; Peskin et al. 2020).

  • Synonyms. Teleocephala (de Pinna 1996:159; Wiley and Johnson 2010:129–130; J. S. Nelson et al. 2016:132–133) is an ambiguous synonym of Teleostei. Many authors (Patterson 1977; Patterson and Rosen 1977; de Pinna 1996; Arratia 2001, 2013; Hilton 2022) use Teleostei as the name for a more inclusive clade that includes several stem fossil lineages (e.g., †Ichthyodectiformes, †Leptolepis, †Pholidophorus, and †Varasichthyidae), which is synonymous with Pan-Teleostei (Moore and Near 2020e).

  • Comments. Müller (1845b) introduced, named, and diagnosed Teleostei with a composition that is nearly identical to the delimitation presented here. Teleosts are an iconic lineage of vertebrates, but evidence for their monophyly, identification of major lineages of teleosts, and resolution of teleost phylogeny did not come into focus until the second half of the 20th century (Gosline 1965; Greenwood et al. 1966; G. J. Nelson 1969a, 1969c; Patterson 1977). The results of this research inaugurated a dramatic shift in ichthyology regarding Teleostei, a situation described by Patterson (1997:201) as: “An analogy is to imagine the situation in mammalogy if monotremes, marsupials, and placentals were not distinguished until 1966.”

  • Bayesian relaxed molecular clock analyses of Teleostei result in an average posterior crown age estimate of 239.6 million years ago, with the credible interval ranging between 224.5 and 256.5 million years ago (Giles et al. 2017).

  • img-z22-5_03.gif

    Oseanacephala C. E. Thacker and T. J. Near, new clade name

  • Definition. The least inclusive crown clade that contains Anguilla rostrata (Lesueur 1817) and Osteoglossum bicirrhosum (Cuvier 1829), but not Engraulis encrasicolus (Linnaeus 1758) or Perca fluviatilis Linnaeus 1758. This is a minimum-crown-clade definition with external specifiers.

  • Etymology. Oseanacephala is a partial acronym composed of the first two letters of Osteoglossomorpha and the first letter from the remaining lineages that comprise the clade: Elopiformes, Albulidae, Notacanthiformes, and Anguilliformes. The suffix is from the ancient Greek ϰεϕαλή (kεf̍αːlә), meaning the head of a human or other animal.

  • Registration number. 882.

  • Reference phylogeny. A phylogeny inferred from concatenated DNA sequences of 1,105 exons (Hughes et al. 2018, fig. S2). See Figures 2 and 3 for the resolution of living lineages of Oseanacephala in the phylogeny of Actinopterygii. See Figure 6 for the phylogenetic relationships of living and fossil lineages of Oseanacephala. Phylogenetic placements of the pan-osteoglossomorphs †Jiuquanichthys and †Lycoptera are on the basis of inferences from morphological analyses (G.-Q. Li and Wilson 1996, 1999; J.-Y. Zhang 1998, 2006; Hilton 2003; Murray et al. 2018).

  • Phylogenetics. In contrast to the substantial support for teleost monophyly from morphological and molecular phylogenetic analyses, uncertainty has remained regarding the relationships among the teleost lineages Clupeocephala, Elopomorpha, and Osteoglossomorpha (Hilton and Lavoué 2018; Dornburg and Near 2021; Takezaki 2021). Taeniopaedia was introduced as a name for the group that included Elopomorpha and Clupeomorpha (Greenwood et al. 1967), which was presented as “Division I” in the classification of teleosts (Greenwood et al. 1966). The morphological phylogeny presented in Patterson and Rosen (1977) resolved Clupeocephala and Elopomorpha as a clade supported with two traits: the presence of two uroneurals in the caudal skeleton that extend beyond ural centrum 2 and the presence of well-developed epipleural intermuscular bones. Arratia (1997) inferred a phylogeny of Pan-Teleostei using parsimony analyses of 131 character state changes coded from living and fossil taxa, resulting in a hypothesis where Elopomorpha is the sister lineage of all other teleosts.

  • Molecular phylogenetic analyses have resulted in all three possible relationships among Clupeocephala, Elopomorpha, and Osteoglossomorpha (Inoue et al. 2001; Hurley et al. 2007; Broughton 2010; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Faircloth et al. 2013; J. N. Chen et al. 2014; Bian et al. 2016; Betancur-R et al. 2017). The earliest molecular phylogenetic studies of teleosts resolved Elopomorpha and Osteoglossomorpha as sister lineages (Lê et al. 1993), which is a frequent result in phylogenomic analyses (M.-Y. Chen et al. 2015; Bian et al. 2016; Lin et al. 2016; Hughes et al. 2018; Vialle et al. 2018; Hao et al. 2020; Roth et al. 2020; Wcisel et al. 2020; Takezaki 2021; Parey et al. 2023). Evidence from genomic organization in the form of the conservation of gene adjacency and the proportion of shared chromosomal breakpoints support monophyly of Oseanacephala (Parey et al. 2023).

  • Composition. There are currently 1,361 living species of Oseanacephala (Fricke et al. 2023) classified in Elopomorpha and Osteoglossomorpha. Fossil lineages of Oseanacephala include the pan-osteoglossomorphs †Jiuquanichthys and †Lycoptera. Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years 139 new living species of Oseanacephala have been described, comprising 10.2% of the living species diversity in the clade.

  • Diagnostic apomorphies. All lineages of Oseanacephala share a chromosomal rearrangement, where duplicated chromosomes 1a and 2a are fused and lineages of Clupeocephala are characterized by an independent fusion of duplicated chromosomes 1b and 2b (Parey et al. 2023). There are no known morphological apomorphies for Oseanacephala; however, fusion of the retroarticular with the angular or the articular is shared by Elopomorpha, Hiodon, and Mormyridae (Parey et al. 2023). The absence of this trait in Gymnarchus niloticus, Notopteridae, Osteoglossidae, and Pantodon may represent a secondary loss in these lineages of Osteoglossomorpha (Parey et al. 2023).

  • Synonyms. Eloposteoglossocephala (Parey et al. 2023) is an ambiguous synonym of Oseanacephala.

  • Comments. The resolution of Oseanacephala is completely driven by molecular phylogenetic analyses and consideration of genomic organization. From the start of phylogenetic investigations of teleosts using morphology, the hypothesis that Elopomorpha and Osteoglossomorpha are sister lineages was never proposed (Patterson and Rosen 1977; Arratia 1997, 1999, 2000c). It is not clear what insight on the evolutionary diversification of teleosts is gained through the resolution of the relationships among Clupeocephala, Elopomorpha, and Osteoglossomorpha, but at minimum it may motivate a reexamination of jaw anatomy and bite kinematics between the bony-tongued Osteoglossomorpha and the complex pharyngeal jaw morphology in Anguilliformes. This resolution also invites investigation of potential commonalities between the robust larvae and juveniles of osteoglossomorph species and the leptocephalus larvae characteristic of Elopomorpha.

  • The earliest Oseanacephala fossil is the pan-elopiform †Anaethalion zapporum from the Kimmeridgian (154.8–149.2 Ma) in the Jurassic of Germany (Arratia 2000c). A Bayesian relaxed molecular clock analysis of Oseanacephala resulted in an average posterior crown age estimate of 223.98 million years ago, but no credible interval was reported (Vialle et al. 2018).

  • img-z25-3_03.gif

    FIGURE 6.

    Phylogenetic relationships of the major living lineages and fossil taxa of Oseanacephala, Osteoglossomorpha, Osteoglossiformes, Osteoglossidae, Elopomorpha, Elopiformes, Albulidae, Notacanthiformes, Anguilliformes, Synaphobranchoidei, Anguilloidei, Muraenoidei, and Congroidei. Asterisk identifies lineages currently classified as Congridae. Lineages currently classified as Congridae are highlighted with an asterisk. Filled circles identify the common ancestor of clades, with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z24-1_03.jpg

    Elopomorpha P. H. Greenwood, D. E. Rosen,
    S. H. Weitzman, and G. S. Meyers 1966:350, 354–358, 393–394
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Elops saurus Linnaeus 1766, Albula vulpes (Linnaeus 1758), and Anguilla rostrata (Lesueur 1817). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek ϰλλoΨ (Ɨl̍αːps), an epithet for fish that may mean either scaly or dumb, for example, “dumb as a fish” (D. W. Thompson 1947:62; Liddell et al. 1968:537) and µoρϕή (m̍ͻ͡ɹ fiː), meaning form or shape.

  • Registration number. 883.

  • Reference phylogeny. A phylogeny inferred from a concatenated dataset of DNA sequences of mitochondrial and nuclear genes and morphological characters (Dornburg et al. 2015, fig. 3). Phylogenetic relationships among living and fossil lineages of Elopomorpha are shown in Figure 6. The resolutions of fossil taxa in the phylogeny are on the basis of inferences from morphological characters (Arratia 1991, 1997, 1999, 2000b, 2000c, 2008, 2010b; Forey et al. 1996; Belouze 2002; Gallo and de Figueiredo 2002; Arratia and Tischlinger 2010; Forey and Maisey 2010; Mayrinck et al. 2010; de Figueiredo, Gallo, and Leal 2012; Pfaff et al. 2016; Guinot and Cavin 2018; Alves et al. 2020; Bean and Arratia 2020; Bean 2021; Hernández-Guerrero et al. 2021).

  • Phylogenetics. The shared presence of specialized leptocephalus larvae was the primary character that led to the delimitation of Elopomorpha to include Elopiformes (including Albulidae), Notacanthiformes, and Anguilliformes (Greenwood et al. 1966). The monophyly of Elopomorpha was challenged in several morphological and molecular inferences that included a de-emphasis on the importance of the leptocephalus larvae (Gosline 1971:100; Nybelin 1971; Hulet and Robins 1989; Filleul and Lavoué 2001; Obermiller and Pfeiler 2003); however, surveys of osteological traits, explicit phylogenetic analyses of morphological character state changes, and molecular phylogenetic analyses consistently resolve elopomorphs as monophyletic (Forey 1973a; G. J. Nelson 1973; Greenwood 1977; Patterson and Rosen 1977; Forey et al. 1996; Inoue et al. 2004; Forey and Maisey 2010; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; J. N. Chen et al. 2014).

  • While analyses of morphological and molecular characters consistently resolve Elopomorpha as monophyletic, relationships among the elopomorph subclades vary among phylogenetic studies. Albulidae, containing Albula, Pterothrissus, and the recently described Nemoossis (Hidaka et al. 2017), is resolved as paraphyletic in some morphological and molecular studies (Forey 1973b; Inoue et al. 2004; Dornburg et al. 2015), but is monophyletic in others (Forey et al. 1996; de Figueiredo, Gallo, and Leal 2012; Alves et al. 2020). Notacanthiformes is resolved as either sharing common ancestry with Albulidae (G. J. Nelson 1973; Greenwood 1977; Patterson and Rosen 1977; C. R. Robins 1989; Inoue et al. 2004; G. D. Johnson et al. 2012) or Anguilliformes (Forey 1973a; Forey et al. 1996; J. N. Chen et al. 2014; Dornburg et al. 2015). There are several morphological character state changes that support Albulidae as the sister lineage of a clade containing Notacanthiformes and Anguilliformes (Forey et al. 1996; Datovo and Vari 2014).

  • Composition. Elopomorpha currently contains 1,107 living species (Fricke et al. 2023) classified in Albulidae, Anguilliformes, Elopiformes, and Notacanthiformes. Fossil taxa of Elopomorpha include the pan-elopiforms †Anaethalion, †Daitingichthys, and †Paraelops (Arratia 1987a; Fielitz and Bardack 1992; de Figueiredo, Gallo, and Leal 2012); the pan-albulids †Baugeichthys, †Brannerion, †Bullichthys, †Farinichthys, †Lebonichthys, and †Osmeroides (Forey et al. 1996, 2003; Filleul 2000; Gallo and de Figueiredo 2002; Forey and Maisey 2010; Mayrinck et al. 2010); and the pan-anguilliforms †Abisaadia, †Anguillavus, †Enchelurus, †Hayenchelys, †Luenchelys, and †Urenchelys (Belouze 2002; Belouze et al. 2003; Pfaff et al. 2016; Guinot and Cavin 2018). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years 124 new living species of Elopomorpha have been described (Fricke et al. 2023), comprising 11.2% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Elopomorpha include (1) presence of the leptocephalus larval stage (Greenwood et al. 1966; Forey 1973a, 1973b; Forey et al. 1996; Inoue et al. 2004; Wiley and Johnson 2010), (2) fusion between angular and retroarticular bones of lower jaw (G. J. Nelson 1973), (3) presence of prenasal ossicles (Forey 1973a, 1973b; Forey et al. 1996; Wiley and Johnson 2010), (4) presence of pectoral splint (Forey 1973a, 1973b; Forey et al. 1996), (5) sternohyoides originates primarily on cleithrum (Greenwood 1977; Wiley and Johnson 2010), (6) spermatozoa flagellum with 9+0 axoneme arrangement and proximal centriole divided into two elongate bundles of four- and five-triplet structure (Matthei and Matthei 1975; Jamieson 1991; Wiley and Johnson 2010), (7) compound neural arch forms in a mass of cartilage over first preural and first ural centrum (Schultze and Arratia 1988; Arratia 1996a, 1997; Forey and Maisey 2010; Wiley and Johnson 2010), and (8) presence of a branchial spiracle (Forey and Maisey 2010).

  • Synonyms. Elopoidei (Gosline 1960:357) and Elopocephalai (Arratia 1999; Betancur-R, Broughton, et al. 2013; Betancur-R et al. 2017:13) are ambiguous synonyms of Elopomorpha.

  • Comments. Greenwood et al. (1966) introduced Elopomorpha as the name for a group that includes Albulidae, Anguilliformes, Elopiformes, and Notacanthiformes, and it is recognized in all subsequent classifications of Teleostei (e.g., G. J. Nelson 1969a; Patterson and Rosen 1977; Wiley and Johnson 2010; Betancur-R et al. 2017). Elopomorpha is an ancient lineage with the pan-elopid †Elopsomolos frickhingeri and pan-elopomorph †Anaethalion zapporum as the earliest known fossil taxa, both of which date from the Kimmeridgian (154.8–149.2 Ma) in the Jurassic (Arratia 2000c; Guinot and Cavin 2018). Bayesian relaxed molecular clock estimates of the crown age of Elopomorpha range between 157 and 200 million years ago in the Jurassic (Dornburg et al. 2015).

  • img-z26-7_03.gif

    Elopiformes P. H. Greenwood, D. E. Rosen,
    S. H. Weitzman, and G. S. Meyers 1966:354, 393
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Elops saurus Linnaeus 1766 and Megalops cyprinoides (Broussonet 1782), but not Albula vulpes (Linnaeus 1758). This is a minimum-crown-clade definition with an external specifier.

  • Etymology. From the ancient Greek ἔλλoΨ (Ɨl'αːps), an epithet for fish that may mean either scaly or dumb, for example, “dumb as a fish” (D. W. Thompson 1947:62; Liddell et al. 1968:537). The suffix is from the Latin word forma, meaning form, figure, or appearance.

  • Registration number. 884.

  • Reference phylogeny. A phylogeny inferred from a concatenated dataset of DNA sequences of mitochondrial and nuclear genes and morphological characters (Dornburg et al. 2015, fig. 3). Phylogenetic relationships of Elopiformes are presented in Figure 6. The placements of the fossil taxa in the phylogeny are on the basis of inferences from morphological characters (Arratia 2000c; de Figueiredo, Gallo, and Leal 2012; Alves et al. 2020; Hernández-Guerrero et al. 2021).

  • Phylogenetics. Elopiformes is consistently resolved as monophyletic in morphological and molecular phylogenetic studies (Forey 1973b; Demartini and Donaldson 1996; Forey et al. 1996; Filleul and Lavoué 2001; Obermiller and Pfeiler 2003; C. Wang et al. 2003; Inoue et al. 2004; Forey and Maisey 2010; de Figueiredo, Gallo, and Leal 2012; G. D. Johnson et al. 2012; K. L. Tang and Fielitz 2013; J. N. Chen et al. 2014; Dornburg et al. 2015; Poulsen et al. 2018; Alves et al. 2020; de Sousa et al. 2021; Hernández-Guerrero et al. 2021). Analyses of mtDNA sequences indicate there are multiple undescribed species masquerading as Elops smithi(McBrideetal.2010;Willifordetal.2022).

  • Composition. There are currently nine living species of Elopiformes classified in Elops and Megalops (Fricke et al. 2023). Fossil lineages of Elopiformes include the pan-megalopid †Elopoides and the pan-elopids †Elopsomolos and †Ichthyemidion (Forey 1973b; Poyato-Ariza 1995; Arratia 2000c). Details of the ages and locations of the fossil taxa are given in Appendix 1. Over the past 10 years no new living species of Elopiformes have been described (McBride et al. 2010; Fricke et al. 2023).

  • Diagnostic apomorphies. Morphological apomorphies for Elopiformes include (1) medial position of posterior opening of mandibular sensory canal within lower jaw (Forey et al. 1996; Wiley and Johnson 2010), (2) presence of posteriorly expanded preopercle (Arratia 2000c), (3) presence of posteriorly expanded opercles and subopercles (Arratia 2000c), (4) presence of well-developed process on mesethmoid (Arratia 2000c), (5) presence of lateral rostral bone (Arratia 2000c), (6) presence of elongated antorbital placed anterior to infraorbital (Arratia 2000c), (7) posterior margins of infraorbitals 3 and 4 do not reach anterior margin of preopercle (Arratia 2000c), (8) anterior portion of ceratohyal not fenestrated (Arratia 2000c), (9) first ossified pleural rib occurring on the fourth or more posterior centrum (Forey and Maisey 2010), and (10) presence of constrictor mandibularis dorsalis, levator arcus palatinia, and pars temporalis (Datovo and Rizzato 2018).

  • Synonyms. Elopoidei (Greenwood et al. 1966:393) is an approximate synonym of Elopiformes.

  • Comments. Greenwood et al. (1966) applied the name Elopiformes to a lineage that included Albulidae, Elopidae, and Megalopidae, a grouping proposed by Gosline (1960) on the basis of morphology of the caudal skeleton. Subsequent phylogenetic studies consistently resolve a clade that accords with our delimitation of Elopiformes as the sister lineage to all other elopomorphs (Forey et al. 1996; Inoue et al. 2004). Elopiformes is an ancient lineage dating to the Jurassic and the pan-elopid †Elopsomolos frickhingeri from the Kimmeridgian (154.8–149.2 Ma) of Germany is the earliest known fossil taxon (Arratia 2000c; Dornburg et al. 2015; Guinot and Cavin 2018). Bayesian relaxed molecular clock crown age estimates for Elopiformes range between 82 and 175 million years ago (Near, Eytan, et al. 2012).

  • img-z27-10_03.gif

    Albulidae P. Bleeker 1849:6, 12
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Albula vulpes (Linnaeus 1758), Nemoossis belloci (Cadenat 1937), and Pterothrissus gissu Hilgendorf 1877. This is a minimum-crown-clade definition.

  • Etymology. Albulae is a Latin name for the Tiber River in Italy (Livy 1919:14–15).

  • Registration number. 885.

  • Reference phylogeny. A phylogeny resulting from a phylogenetic analysis of morphological character state changes (Forey and Maisey 2010, fig. 13). Nemoossis belloci (Longfin Bonefish) is not included in any phylogenetic analyses, but it is assumed it will resolve as the sister species of Pterothrissus gissu (Japanese Gissu) (Hidaka et al. 2017). Phylogenetic relationships of Albulidae (bonefishes) are shown in Figure 6. The placements of fossil taxa in the phylogeny are on the basis of inferences from morphological characters (Fielitz and Bardack 1992; Gallo and de Figueiredo 2002; de Figueiredo, Gallo, and Leal 2012; Guinot and Cavin 2018; Alves et al. 2020; Hernández-Guerrero et al. 2021; L-Recinos et al. 2023).

  • Phylogenetics. Classifications of Teleostei from the early to mid-20th century grouped Albula and Pterothrissus in either Albulidae, Albuloidae, or Albuloidei (Boulenger 1904b:547–549; Goodrich 1909:387–388; Berg 1940:420; Greenwood et al. 1966). Reflecting alternative classifications that grouped Albula and Pterothrissus in separate and unrelated family rank taxonomic groups (Jordan 1905:44, 46–48), it was proposed that Pterothrissus is the sister lineage of a clade containing Notacanthiformes and Anguilliformes on the basis of shared similarities of an elongate snout, subterminal mouth, reduced ossification of the braincase, and inwardly turned head of the maxilla (Forey 1973a). Subsequent morphological studies consistently resolve Albulidae as monophyletic (Greenwood 1977; Forey et al. 1996), and several morphological phylogenetic analyses incorporated fossil taxa that are either more closely related to Albula or to the Pterothrissus-Nemoossis clade (Forey and Maisey 2010; de Figueiredo, Gallo, and Leal 2012; Guinot and Cavin 2018; Alves et al. 2020; Hernández-Guerrero et al. 2021; L-Recinos et al. 2023). Molecular phylogenies differ in their support for the monophyly of Albulidae. Analyses of mitochondrial DNA and concatenated nuclear genes each result in paraphyly of Albulidae, with Pterothrissus resolved as the sister lineage of Notacanthiformes (Inoue et al. 2004; G. D. Johnson et al. 2012; Dornburg et al. 2015); however, other phylogenetic studies using mitochondrial DNA result in the resolution of a monophyletic Albulidae (C. Wang et al. 2003; K. L. Tang and Fielitz 2013; Poulsen et al. 2018).

  • Composition. Albulidae currently contains 13 living species classified in Albula, Nemoossis, and Pterothrissus (Hidaka et al. 2017; Fricke et al. 2023). Fossil taxa of Albulidae include †Deltaichthys, †Hajulia, †Istieus, †Macabi, and †Nunaneichthys (Forey and Maisey 2010; de Figueiredo, Gallo, and Leal 2012; Guinot and Cavin 2018; Alves et al. 2020; Hernández-Guerrero et al. 2021; L-Recinos et al. 2023). Details of the ages and locations of the fossil taxa are presented in Appendix 1. There were no new living species of Albulidae described over the past 10 years, but there remains at least one undescribed species of Albula (Pickett et al. 2020).

  • Diagnostic apomorphies. Morphological apomorphies for Albulidae include (1) presence of subepiotic fossa (Forey et al. 1996; Wiley and Johnson 2010), (2) ectopterygoid with dorsal process (Forey et al. 1996; Wiley and Johnson 2010), (3) presence of fenestration within hyomandibular and metapterygoid suture that allows levator arcus palatine to pass through and insert on medial surface of palate (Forey et al. 1996; Wiley and Johnson 2010), and (4) sternohyoideus originating mainly on cleithrum (Forey et al. 1996).

  • Synonyms. Albuloidae (Berg 1940:420), Albuloidei (Greenwood et al. 1966:393; Forey 1973a:94), and Albuliformes (Forey et al. 1996:184; J. S. Nelson et al. 2016; Betancur-R et al. 2017:14) are all ambiguous synonyms of Albulidae.

  • Comments. When Bleeker (1849) introduced the name Albulidae, there was only one taxon, Albula, classified in the group (Günther 1868:468). Shortly after the description of Pterothrissus (Hilgendorf 1877), several classifications of teleosts grouped Albula and Pterothrissus in Albulidae (Boulenger 1904b:547–549; Goodrich 1909:387–388). Albulidae was selected as the clade name over its synonyms because they are redundant group names relative to Albulidae in ranked taxonomies. Albulidae is a valid family-group name under the International Code of Zoological Nomenclature (Van der Laan et al. 2014:64).

  • The earliest fossil taxa in Albulidae is †Nunaneichthys mexicanus from the Albian-Cenomanian (100.5–93.9 Ma) in the Cretaceous from Mexico (Hernández-Guerrero et al. 2021). There are no molecular divergence time estimates for Albulidae.

  • img-z29-4_03.gif

    Notacanthiformes E. S. Goodrich 1909:416
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Notacanthus chemnitzii Bloch 1788 and Halosaurus ovenii J. Y. Johnson 1864. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek νῶτoν (no͡Ʊtәn), meaning of the back, and ἇκανθα (æk'ænθә), meaning thorn or spine. The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 886

  • Reference phylogeny. A phylogeny resulting from analysis of a concatenated dataset of DNA sequences from mitochondrial and nuclear genes and morphological characters (Barros-García et al. 2018, fig. 1b). Phylogenetic relationships of Notacanthiformes are presented in Figure 6. The placement of the fossil lineage †Echidnocephalus in the phylogeny is based on inferences from morphological characters (Forey et al. 1996; Arratia 2010b; Guinot and Cavin 2018).

  • Phylogenetics. Classifications of teleost fishes from the early 20th century grouped Notacanthidae (deep-sea spiny eels) and Halosauridae (halosaurs) with the pan-aulopiform †Dercetidae and Fierasfer, which is a synonym of the ophiid Carapus (Boulenger 1904b; Goodrich 1909:416–419). Regan (1909b) removed †Dercetidae and Fierasfer and limited the group Heteromi to Notacanthidae and Halosauridae. Notacanthiformes, comprising Notacanthidae and Halosauridae, was identified as one of the major lineages of Elopomorpha (Greenwood et al. 1966) and subsequent phylogenetic analyses have supported notacanthiform monophyly (Forey et al. 1996; Inoue et al. 2004; J. N. Chen et al. 2014; Dornburg et al. 2015; Barros-García et al. 2018; Poulsen et al. 2018). There is less certainty on the relationships of Notacanthiformes among major lineages of Elopomorpha. Some phylogenetic analyses of morphological and molecular characters place Notacanthiformes as the sister lineage of Albulidae (Greenwood 1977; Patterson and Rosen 1977; C. R. Robins 1989; C. Wang et al. 2003; Inoue et al. 2004; G. D. Johnson et al. 2012; Near, Eytan, et al. 2012), but other studies resolve Notacanthiformes and Anguilliformes as sister lineages (Forey 1973a; Forey et al. 1996; Santini, Kong, et al. 2013; K. L. Tang and Fielitz 2013; J. N. Chen et al. 2014; Dornburg et al. 2015). Forey et al. (1996) identified 14 morphological synapomorphies for a clade containing Notacanthiformes and Anguilliformes; many of these traits are character losses in the context of their evolution within Elopomorpha (Wiley and Johnson 2010).

  • Composition. There are currently 28 living species of Notacanthiformes (Fricke et al. 2023) classified in Notacanthidae and Halosauridae. Fossil lineages of Notacanthiformes include the pan-halosaurid †Echidnocephalus (Forey et al. 1996; Arratia 2010b). Details of the age and location of †Echidnocephalus are presented in Appendix 1. Over the past 10 years one new species of Notacanthiformes has been described (Fricke et al. 2023).

  • Diagnostic apomorphies. Morphological apomorphies for Notacanthiformes include (1) complete separation of parmalaris from remaining muscles of adductor mandibulae (Greenwood 1977; Datovo and Vari 2014), (2) nodule between maxillary head and palatine (Greenwood 1977; Wiley and Johnson 2010), (3) presence of posteriorly directed spine on maxilla (Forey et al. 1996; Wiley and Johnson 2010), and (4) pelvic fins connected by membrane (Forey et al. 1996; Wiley and Johnson 2010).

  • Synonyms. Heteromi (T. N. Gill 1893:133; Regan 1909a:82–83) and Halosauri (Garstang 1931:258) are approximate synonyms of Notacanthiformes.

  • Comments. Greenwood et al. (1966) limited Notacanthiformes to Notacanthidae and Halosauridae. The name Notacanthiformes was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest notacanthiform fossil taxon is the pan-halosaurid †Echidnocephalus troscheli from the Campanian (83.2–72.2 Ma) in the Cretaceous of Germany (Forey et al. 1996; Arratia 2010b). Bayesian relaxed molecular clock age estimates of Notacanthiformes result in an average posterior crown age estimate between 70 and 125 million years ago (Dornburg et al. 2015).

  • img-z30-6_03.gif

    Anguilliformes E. S. Goodrich 1909:403
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Myroconger compressus Günther 1870, Gymnothorax formosus Bleeker 1864b, Protanguilla palau Johnson, Ida, and Sakaue in G. D. Johnson et al. 2012, Synaphobranchus kaupi J. Y. Johnson 1862, Conger oceanicus (Mitchill 1818a), and Anguilla anguilla (Linnaeus 1758). This is a minimum-crown-clade definition.

  • Etymology. From the Latin Anguilla, meaning eel, and forma, meaning form, figure, or appearance.

  • Registration number. 887.

  • Reference phylogeny. A phylogeny inferred from a concatenated dataset of DNA sequences from three nuclear genes and two mitochondrial protein coding genes (Santini, Kong, et al. 2013, fig. 2). Phylogenetic relationships of Anguilliformes are presented in Figure 6.

  • Phylogenetics. Species classified as Anguilliformes were initially grouped with unrelated eel-like species in Apodes of Linnaeus (1758:242). By the middle of the 19th century, a taxonomic group comprising the modern Anguilliformes was established (Bleeker 1864c). Greenwood et al. (1966) delimited two groups within Anguilliformes, Anguilloidei for the typical eels and Saccopharyngoidei containing the morphologically bizarre deep-sea lineages that included Saccopharynx (swallowers), Eurypharynx pelecanoides (Pelican Eel), and Monognathus (onejaw gulpers). The saccopharyngoids are so morphologically unique that it has been proposed they were a divergent lineage not closely related to any living osteichthyans (Tchernavin 1947). The saccopharyngoid traits include the absences of ventral fins, pelvic girdle, opercular bones, and branchiostegals (Böhlke 1966; Bertelsen et al. 1989). The saccopharyngoids were included with all other eels in Boulenger's (1904b:599–605) Apodes and in Goodrich's (1909:403–408) Anguilliformes. In a classification of teleosts, Regan (1909b) grouped anguilloids and saccopharyngoids in Apodes, but he later put Saccopharynx in T. N. Gill and Ryder's (1883) Lyomeri, established to accommodate the saccopharyngoid Eurypharynx pelecanoides (Regan 1912a, 1912e). On the basis of comparative morphology, C. R. Robins (1989) countered the classification of Anguilliformes presented in Greenwood et al. (1966) and vigorously promoted the hypothesis that anguilloids and saccopharyngoids are distantly related. The delimitation of the saccopharyngoids was later expanded to include the bobtail snipe eels Cyema atrum and Neocyema erythrosoma (Raju 1974; Castle 1977).

  • Subsequent to the delimitation of Elopomorpha (Greenwood et al. 1966), there is broad support for the monophyly of Anguilliformes in morphological and molecular phylogenetic studies (Forey 1973a; Forey et al. 1996; Inoue et al. 2003b, 2004, 2010; Obermiller and Pfeiler 2003; C. Wang et al. 2003; López et al. 2007; G. D. Johnson et al. 2012; Santini, Kong, et al. 2013; K. L. Tang and Fielitz 2013; J. N. Chen et al. 2014; Dornburg et al. 2015; Poulsen et al. 2018). The use of morphological characters to investigate phylogenetic relationships of Anguilliformes is challenged by difficulties in constructing inclusive data matrices due to limited knowledge on anguilliform anatomy and the reductive nature of eel skeletons (Forey et al. 1996). The situation is improving with detailed studies of gill arch musculature (Springer and Johnson 2015), the pharyngeal jaw apparatus (G. D. Johnson 2019), and the pectoral girdle (da Silva and Johnson 2018). A recent phylogenetic analysis of Congroidei using 42 coded characters from the pectoral girdle demonstrates the potential for explicit phylogenetic analysis of morphological traits in resolving relationships within Anguilliformes (da Silva et al. 2019).

  • Despite the historic challenges of using morphology to investigate anguilliform phylogeny, parsimony analysis of a data matrix of morphological character state changes resulted in the nesting of saccopharyngoids within the anguilloids (Forey et al. 1996). The paraphyly of anguilloids relative to saccopharyngoids is reflected in several molecular phylogenetic analyses (Inoue et al. 2003b, 2004, 2010; Santini, Kong, et al. 2013; J. N. Chen et al. 2014; Dornburg et al. 2015; Poulsen et al. 2018). The issue of the phylogenetic affinities of saccopharyngoids is effectively settled, as evidenced by a proposed set of taxonomic groupings in Anguilliformes that do not include Saccopharyngiformes or Saccopharyngoidei, classifying them with the anguilloid lineages Anguillidae (freshwater eels), Moringuidae (spaghetti eels), Nemichthyidae (snipe eels), and Serrivomeridae (sawtooth eels) (K. L. Tang and Fielitz 2013). Molecular phylogenetic analyses resolve Anguilliformes into four clades: Synaphobranchoidei, Muraenoidei, Congroidei, and Anguilloidei (K. L. Tang and Fielitz 2013). Several currently recognized taxa within Anguilliformes are non-monophyletic in molecular phylogenetic analyses, including Chlopsidae (false morays), Coloconger (shottail eels), Congridae (conger eels), Cyematidae (bobtail snipe eels), Derichthyidae (narrowneck eels), and Nettastomatidae (Santini, Kong, et al. 2013; K. L. Tang and Fielitz 2013; Poulsen et al. 2018; Lü et al. 2019; K. Zhang, Zhu, et al. 2021; Huang et al. 2023).

  • Composition. Anguilliformes currently contains 1,057 species (Fricke et al. 2023) classified in Anguilloidei, Chlopsidae, Congroidei, Muraenoidei, and Synaphobranchoidei (K. L. Tang and Fielitz 2013). Over the past 10 years 122 new living species of Anguilliformes have been described, comprising 11.5% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Anguilliformes include (1) symplectic fused with quadrate (Forey et al. 1996; Wiley and Johnson 2010), (2) absence of first pharyngobranchial, gill arches displaced posteriorly and free from the neurocranium (Forey et al. 1996; Wiley and Johnson 2010; Espíndola et al. 2023), (3) absence of pelvic girdle and pelvic fins (Forey et al. 1996; Wiley and Johnson 2010), (4) body scales absent or embedded with a basket-weave pattern (C. R. Robins 1989; Wiley and Johnson 2010; G. D. Johnson et al. 2012; Espíndola et al. 2023), (5) ceratohyal with elongated anterior end (C. R. Robins 1989; Wiley and Johnson 2010), (6) anterior branchiostegals curve behind and above opercle (C. R. Robins 1989; Wiley and Johnson 2010; G. D. Johnson et al. 2012; Espíndola et al. 2023), (7) endopterygoid absent (G. D. Johnson et al. 2012; Espíndola et al. 2023), (8) single hypohyal or hypohyal absent (G. D. Johnson et al. 2012; Espíndola et al. 2023), (9) dorsal and anal fins confluent with caudal fin (G. D. Johnson et al. 2012; Espíndola et al. 2023), (10) fewer than eight caudal fin rays in each lobe (G. D. Johnson et al. 2012; Espíndola et al. 2023), (11) posttemporal absent (G. D. Johnson et al. 2012; Espíndola et al. 2023), (12) epurals absent (G. D. Johnson et al. 2012; Espíndola et al. 2023), (13) absence of levator internus 3 (Springer and Johnson 2015; Espíndola et al. 2023), (14) presence of musculus pharyngobranchialis 2-epibranchialis 1 (Springer and Johnson 2015; Espíndola et al. 2023), (15) presence of a single pharyngoclavicularis (Springer and Johnson 2015; Espíndola et al. 2023), (16) presence of rectus ventralis 3 and 4 (Springer and Johnson 2015; Espíndola et al. 2023), (17) absence of rectus communis (Springer and Johnson 2015; Espíndola et al. 2023), (18) levator internus 2 insertion includes upper tooth plate 4 (Springer and Johnson 2015; Espíndola et al. 2023), (19) hypobranchial 3 either absent or entirely cartilaginous (Springer and Johnson 2015; Espíndola et al. 2023), (20) absence of accessory element at distal end of ceratobranchial 4 (Springer and Johnson 2015; Espíndola et al. 2023), (21) adductor mandibulae originates on parietal (Espíndola et al. 2023), and (22) adductor mandibulae lacks segmentum mandibularis (Espíndola et al. 2023).

  • Synonyms. Apodes (Kaup 1856:1; Boulenger 1904b:600–605; Regan 1912e:377–379; Jordan 1923:130; Trewavas 1932:655–656) and Muraeni (Bleeker 1864c:113) are approximate synonyms of Anguilliformes. Encheli is a partial synonym of Anguilliformes (Garstang 1931:257).

  • Comments. The composition of Anguilliformes in Goodrich (1909:370, 403–404) is very close to that delimited here and in Greenwood et al. (1966), the differences being the addition of lineages discovered after these important studies (e.g., Castle 1977; G. D. Johnson et al. 2012). The name Anguilliformes was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • While there are several fossil lineages of pan-anguilliforms from the Cretaceous, the earliest fossil Anguilliformes are from the Ypresian (56.0–48.1 Ma) in the Eocene of Italy (Bannikov 2014b; Carnevale et al. 2014; Pfaff et al. 2016). Relaxed molecular clock analyses estimate the crown age of Anguilliformes between 84 and 116 million years ago (Santini, Kong, et al. 2013).

  • img-z32-6_03.gif

    Synaphobranchoidei P. Bleeker 1864a:13
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Synaphobranchus kaupi J. Y. Johnson 1862, Simenchelys parasitica Gill in Bean and Goode 1879, and Protanguilla palau Johnson, Ida, and Sakaue in G. D. Johnson et al. 2012. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek σύν (s̍In), meaning together; ἁϕἠ (ɐf̍ε), meaning a joint or a fastening; and βραγχίoν (bɹ̍æɡki͡әn), Latinized as branchium, meaning a fish gill.

  • Registration number. 888.

  • Reference phylogeny. A phylogeny inferred from a concatenated dataset of DNA sequences from three nuclear genes and two mitochondrial protein coding genes (Santini, Kong, et al. 2013, fig. 2). Phylogenetic relationships of Synaphobranchoidei are presented in Figure 6.

  • Phylogenetics. Molecular phylogenetic analyses consistently resolve Protanguilla palau (Cave Eel) and species of Synaphobranchidae (cutthroat eels) as a monophyletic group (Santini, Kong, et al. 2013; K. L. Tang and Fielitz 2013; Poulsen et al. 2018). Some investigators contend that morphological character state changes support Protanguilla as the sister lineage of all other Anguilliformes (G. D. Johnson et al. 2012; Espíndola et al. 2023), but this inference is on the basis of the distribution of several key morphological traits and not the result of an explicit phylogenetic analysis of coded character state changes.

  • Composition. Synaphobranchoidei currently contains 53 species (Fricke et al. 2023) classified in Synaphobranchidae and Protanguilla (K. L. Tang and Fielitz 2013). Over the past 10 years 15 new living species of Synaphobranchoidei have been described (Fricke et al. 2023), comprising 28.3% of the living species diversity in the clade.

  • Diagnostic apomorphies. The leptocephalus larvae of species of Synaphobranchidae are unique among all other lineages of Anguilliformes in possessing vertically or diagonally elongated eyes (C. H. Robins and Robins 1989). The morphology of the larvae of Protanguilla palau is not known.

  • Synonyms. There are no synonyms of Synaphobranchoidei.

  • Comments. Bleeker (1864a,13) applied Synaphobranchoidei as a ranked taxonomic family to classify Synaphobranchus kaupi. Given the resolution of Synaphobranchidae and Protanguilla as sister lineages in molecular phylogenies (Santini, Kong, et al. 2013; Poulsen et al. 2018), K. L. Tang and Fielitz (2013, tbl. II) revised the classification of Anguilliformes with a new delimitation of Synaphobranchoidei that is the basis of the definition presented here. Bayesian relaxed molecular clock analysis estimates the crown age of Synaphobranchoidei between 55 and 108 million years ago (Santini, Kong, et al. 2013).

  • img-z33-5_03.gif

    Anguilloidei P. Bleeker 1859:xxxiii
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Anguilla anguilla (Linnaeus 1758), Moringua microchir Bleeker 1853a, and Serrivomer beanii T. N. Gill and Ryder 1883. This is a minimum-crown-clade definition.

  • Etymology. From the Latin Anguilla, meaning eel.

  • Registration number. 889.

  • Reference phylogeny. A phylogeny inferred from a concatenated dataset of DNA sequences from three nuclear genes and two mitochondrial protein coding genes (Santini, Kong, et al. 2013, fig. 2). See Figure 6 for a phylogeny of the major lineages of Anguilloidei.

  • Phylogenetics. Several molecular phylogenetic analyses result in the monophyly of Anguilloidei, with Moringuidae resolved as the sister lineage of all other anguilloids (Santini, Kong, et al. 2013; K. L. Tang and Fielitz 2013; Poulsen et al. 2018). The lineages previously classified as Lyomeri, Saccopharyngiformes, or Saccopharyngoidei are nested in Anguilloidei, but analyses differ on the monophyly of a group containing Monognathidae, Cyematidae, Saccopharyngidae, Neocyematidae, and Eurypharyngidae (Santini, Kong, et al. 2013; Poulsen et al. 2018).

  • Composition. Anguilloidei currently contains 81 species (Fricke et al. 2023) classified in Anguillidae, Cyema, Eurypharynx, Monognathidae, Moringuidae, Nemichthyidae, Neocyema, Saccopharyngidae, and Serrivomeridae. Over the past 10 years no new living species of Anguilloidei have been described (Fricke et al. 2023).

  • Diagnostic apomorphies. There are no known morphological apomorphies for Anguilloidei.

  • Synonyms. Lyomeri (T. N. Gill and Ryder 1883:263–264; Jordan 1923:134; Garstang 1931:257; Böhlke 1966:603–610), Saccopharyngiformes (Berg 1940:439–440; McAllister 1968:88–89; C. R. Robins 1989:13–15), and Saccopharyngoidei (Greenwood et al. 1966:393; J. S. Nelson 2006:125; J. S. Nelson et al. 2016:149–150) are all partial synonyms of Anguilloidei.

  • Comments. Bleeker (1864c) applied Anguilloidei as a ranked taxonomic family to a group containing Anguilla anguilla and the fossil taxon †Paranguilla tigrina. Greenwood et al. (1966) classified all Anguilliformes that were not in their Saccopharyngoidei into Anguilloidei. On the basis of the resolution of clades in molecular phylogenetic analyses, K. L. Tang and Fielitz (2013) revised the classification of Anguilliformes with a new delimitation of Anguilloidei that is the basis of the definition presented here. Bayesian relaxed molecular clock analysis estimates the crown age of Anguilloidei between approximately 64 and 90 million years ago (Santini, Kong, et al. 2013).

  • img-z34-2_03.gif

    Muraenoidei L. J. F. J. Fitzinger 1832:332
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive clade that contains Muraena helena Linnaeus 1758, Myroconger compressus Günther 1870, and Pythonichthys microphthalmus (Regan 1912d). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek µύραινα (mjƱɹɹ̍e͡Inә) that is the name of the Mediterranean Moray, Muraena helena (D. W. Thompson 1947:162–165).

  • Registration number. 890.

  • Reference phylogeny. A phylogeny inferred from a concatenated dataset of the mitochondrial 12S and 16S rRNA genes (K. L. Tang and Fielitz 2013, fig. 1). Phylogenetic relationships among the lineages of Muraenoidei are presented in Figure 6.

  • Phylogenetics. Molecular phylogenetic analyses resolve Muraenoidei as monophyletic, with Myroconger (thin eels) and Muraenidae (moray eels) as sister clades relative to Heterenchelyidae (Inoue et al. 2010; G. D. Johnson et al. 2012; Santini, Kong, et al. 2013; K. L. Tang and Fielitz 2013; Poulsen et al. 2018). Alternatively, a morphological analysis results in paraphyly of Muraenoidei, with the chlopsid Xenoconger fryeri (Fryer's False Moray) resolved as the sister lineage of Muraenidae relative to Myroconger (D. G. Smith 1984).

  • Composition. Muraenoidei currently contains 238 species (Fricke et al. 2023) classified in Heterenchelyidae (mud eels), Muraenidae, and Myroconger (K. L. Tang and Fielitz 2013). Over the past 10 years 23 new living species of Muraenoidei have been described (Fricke et al. 2023), composing 9.7% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Muraenoidei include (1) frontals not fused (J. S. Nelson et al. 2016), (2) reduction in gill arch elements (J. S. Nelson et al. 2016), (3) reduction of lateral line (J. S. Nelson et al. 2016), and (4) normal-sized eyes (J. S. Nelson et al. 2016).

  • Synonyms. There are no synonyms of Muraenoidei.

  • Comments. Müller (1845b) used Muraenoidei as a family group name in his classification of Teleostei. Muraenoidei was later treated as a suborder containing Chlopsidae (false morays), Muraenidae, and Myrocongridae (C. R. Robins 1989). Given the resolution of a clade containing Heterenchelyidae, Muraenidae, and Myroconger in molecular phylogenies (Inoue et al. 2010; G. D. Johnson et al. 2012; Santini, Kong, et al. 2013; Poulsen et al. 2018), K. L. Tang and Fielitz (2013, table II) revised the classification of Anguilliformes with a new delimitation of Muraenoidei that is the basis of the definition presented here. Bayesian relaxed molecular clock analysis estimates the crown age of Muraenoidei ranges between 60 and 90 million years ago (Santini, Kong, et al. 2013).

  • img-z34-15_03.gif

    Congroidei P. Bleeker 1864a:18 [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Conger conger (Linnaeus 1758), Conger oceanicus (Mitchill 1818a), Derichthys serpentinus T. N. Gill 1884b, Heteroconger hassi (Klausewitz and Eibl-Eibesfeldt 1959), and Ophichthys zophochir Jordan and Gilbert 1882a. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek γόγγρoϛ (ɡ̍ͻŋɹo͡Ʊz), meaning conger eel, Latinized to conger (D. W. Thompson 1947:49–50).

  • Registration number. 891.

  • Reference phylogeny. A phylogeny inferred from a concatenated dataset of DNA sequences from three nuclear genes and two mitochondrial protein coding genes (Santini, Kong, et al. 2013, fig. 2). Although Conger conger is not included in the reference phylogeny, it resolves with other species of Conger in molecular phylogenetic analyses (J. N. Chen et al. 2014, figs. 1, 2). The relationships among the lineages of Congroidei are shown in Figure 6.

  • Phylogenetics. The lineages delimited here in Congroidei are consistently resolved as monophyletic in molecular phylogenetic analyses (Inoue et al. 2010; G. D. Johnson et al. 2012; Santini, Kong, et al. 2013; K. L. Tang and Fielitz 2013; Poulsen et al. 2018). Despite the strong support for monophyly of Congroidei, the phylogenetics is complicated by the nonmonophyly of Derichthyidae (narrowneck eels) and Congridae (conger eels) in the analyses. Molecular studies support monophyly of a lineage containing Colocongridae (shorttail eels), Congriscus (Congridae), and Derichthyidae (López et al. 2007; Santini, Kong, et al. 2013; Poulsen et al. 2018); however, Nessorhamphus (Derichthyidae) and Congriscus are sister lineages, and Derichthys is resolved as sister to Colocongridae (worm eels) (Santini, Kong, et al. 2013). There is poor support for many of these nodes in the molecular phylogenies, but a morphological analysis provides strong support for the monophyly of Derichthyidae and resolution of a clade containing Colocongridae, Congriscus, and Derichthyidae (da Silva et al. 2019). Species of Nettastomatidae (duckbill eels) and the Congridae subclade Congrinae form a clade (Santini, Kong, et al. 2013; Poulsen et al. 2018), but Nettastomatidae and Congrinae are both paraphyletic. The Congridae subclades Bathymyrinae and Heterocongrinae are resolved as a clade (Santini, Kong, et al. 2013; Poulsen et al. 2018), but Bathymyrinae is rendered paraphyletic by Heteroconger (Santini, Kong, et al. 2013). Muraenesocidae (pike congers) and Ophichthidae (snake eels) are both monophyletic and resolved as sister lineages (Santini, Kong, et al. 2013; Poulsen et al. 2018).

  • Composition. Congroidei currently contains 660 species (Fricke et al. 2023) classified in Congridae, Coloconger, Derichthyidae, Muraenesocidae, Nettastomatidae, and Ophichthidae. Over the past 10 years 81 new living species of Congroidei have been described (Fricke et al. 2023), comprising 12.3% of the living species diversity in the clade.

  • Diagnostic apomorphies. There are no known morphological apomorphies for Congroidei.

  • Synonyms. There are no synonyms for Congroidei.

  • Comments. Bleeker (1864a:18) applied Congroidei as a taxonomic family. Classifications of Anguilliformes in the 20th century delimited Congroidei as a more inclusive group than given here on the basis of the presence of fused frontal bones in the skull (C. R. Robins 1989; J. S. Nelson 1994). Given the results of molecular phylogenetic analyses (Santini, Kong, et al. 2013; Poulsen et al. 2018), K. L. Tang and Fielitz (2013, table II) revised the classification of Congroidei to include Chlopsidae, Congridae, Derichthyidae, Muraenesocidae, Nettastomatidae, and Ophichthidae, which is not followed here. Bayesian relaxed molecular clock analysis estimates the crown age of Congroidei between 64 and 90 million years ago (Santini, Kong, et al. 2013).

  • img-z35-10_03.gif

    Osteoglossomorpha P. H. Greenwood,
    D. E. Rosen, S. H. Weitzman, and G. S. Meyers
    1966:350, 354–358, 393–394
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Hiodon tergisus Lesueur 1818, Pantodon buchholzi Peters 1876, Notopterus notopterus (Pallas 1769), and Osteoglossum bicirrhosum (Cuvier 1829). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek ὀστέoν (̍αːstIәn), meaning bone; γλῶσσα (ɡl̍ͻsә), meaning tongue; and µoρϕή (m̍ͻ͡ɹfiː), meaning form or shape.

  • Registration number. 892.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 546 exons (Peterson et al. 2022, fig. 1e). Phylogenetic relationships of the major living lineages and fossil taxa of Osteoglossomorpha are shown in Figure 6. The placements of the fossil lineages in the phylogeny are based on analyses of morphological characters (J.-Y. Zhang 1998, 2006; G.-Q. Li and Wilson 1999; Xu and Chang 2009).

  • Phylogenetics. Several studies prior to the mid-1960s hinted at a close relationship among what are now considered the lineages of Osteoglossomorpha (Ridewood 1904, 1905; Garstang 1931; Gregory 1933:161–175; Gosline 1960, 1961; Greenwood 1963), but it was Greenwood et al. (1966) that named the group and solidified evidence for its monophyly. The monophyly of Osteoglossomorpha is supported in several morphological and molecular phylogenetic analyses and Hiodon (mooneyes) and Osteoglossiformes are consistently resolved as sister groups (G.-Q. Li and Wilson 1996, 1999; G.-Q. Li, Wilson, et al. 1997; J.-Y. Zhang 1998; Hilton 2003; Inoue et al. 2003a, 2009; Lavoué and Sullivan 2004; M. V. H. Wilson and Murray 2008; Santini et al. 2009; Xu and Chang 2009; Lavoué, Miya, Arnegard, et al. 2012; Betancur-R, Broughton, et al. 2013; Hilton and Lavoué 2018; Murray et al. 2018; Brito et al. 2020; Peterson et al. 2022).

  • Composition. Osteoglossomorpha currently contains 254 living species (Fricke et al. 2023) classified in Hiodon and Osteoglossiformes. There are several fossil lineages of Osteoglossomorpha that include pan-hiodontids †Plesiolycoptera and †Yanbiania (G.-Q. Li and Wilson 1996, 1999; J.-Y. Zhang 1998) and the pan-osteoglossiforms †Paralycoptera, †Jinanichthys, †Huashia, and †Kuntulunia (Jiangyong 1990; G.-Q. Li and Wilson 1996, 1999; J.-Y. Zhang 1998, 2006; Xu and Chang 2009; Murray et al. 2018). Details of the ages and locations of the fossil taxa are given in Appendix 1. Over the past 10 years 15 new living species of Osteoglossomorpha have been described (Fricke et al. 2023), comprising 5.9% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Osteoglossomorpha include (1) primary bite between parasphenoid and basihyal; however, this trait is an apomorphy for a more inclusive pan-osteoglossomorphs (Greenwood et al. 1966; G.-Q. Li and Wilson 1996), (2) supramaxilla absent (G.-Q. Li and Wilson 1996, 1999; J.-Y. Zhang 2006; Xu and Chang 2009), (3) fourth and fifth infraorbitals fused (G.-Q. Li and Wilson 1996, 1999; J.-Y. Zhang 1998), (4) last uroneural much shorter than first uroneural (J.-Y. Zhang 1998), (5) rectangular-shaped infraorbital bone (G.-Q. Li and Wilson 1999), (6) seven pelvic-fin rays (J.-Y. Zhang 2006), (7) nasal bones tubular and strongly curved (Hilton 2003), (8) supraorbital sensory canal ending in frontal bone (Hilton 2003; M. V. H. Wilson and Murray 2008), (9) ascending process of premaxilla not developed or slightly developed (Hilton 2003; M. V. H. Wilson and Murray 2008), (10) autopalatine bone absent (M. V. H. Wilson and Murray 2008), (11) supraorbital absent (Mirande 2017), (12) complete absence of epurals (Mirande 2017), (13) bony epipleurals absent (Mirande 2017), and (14) intestine coils to the left of the stomach (Mirande 2017).

  • Synonyms. Osteoglossi is a partial (Garstang 1931:256–257) and an ambiguous (Gosline 1960:358) synonym of Osteoglossomorpha. Osteoglossoidei (Gosline 1960:358) is an ambiguous synonym of Osteoglossomorpha.

  • Comments. The earliest fossil osteoglossomorphs include the pan-osteoglossiforms †Paralycoptera, †Jinanichthys, and †Huashia that date from the Aptian (121.4–113.2 Ma) in the Cretaceous of China. Bayesian relaxed molecular clock analyses of Osteoglossomorpha result in an average posterior crown age estimate of 234.4 million years ago, with the credible interval ranging between 212.4 and 259.0 million years ago (Peterson et al. 2022).

  • img-z37-2_03.gif

    Osteoglossiformes P. H. Greenwood 1963:408
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Pantodon buchholzi Peters 1876, Notopterus notopterus (Pallas 1769), and Osteoglossum bicirrhosum (Cuvier 1829). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek ὀστέoν ('αːstIәn), meaning bone, and γλῶσσα (ɡl'ͻsә), meaning tongue. The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 896.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 546 exons (Peterson et al. 2022, fig. 1e). Phylogenetic relationships among the major lineages of Osteoglossiformes are shown in Figure 6. The placements of the fossil taxa †Palaeonotopterus and †Laeliichthys in the phylogeny are based on analyses of morphological characters (Cavin and Forey 2001; Murray et al. 2018; Brito et al. 2020).

  • Phylogenetics. The monophyly of Osteoglossiformes is supported in several morphological and molecular phylogenetic analyses (Taverne 1979, 1998; G.-Q. Li and Wilson 1996, 1999; G.-Q. Li, Wilson, et al. 1997; Hilton 2003; Lavoué and Sullivan 2004; J.-Y. Zhang 2006; M. V. H. Wilson and Murray 2008; Inoue et al. 2009; Xu and Chang 2009; Lavoué et al. 2011; Lavoue, Miya, Arnegard, et al. 2012; Lavoué 2015, 2016; Murray et al. 2018; Brito et al. 2020; Peterson et al. 2022). Within Osteoglossiformes, there is consistent support for the monophyly of a lineage consisting of Notopteridae (featherfin knifefishes) and the clade Mormyroidea, which contains Mormyridae (elephantfishes) and Gymnarchus niloticus (Aba) (Taverne 1979, 1998; G.-Q. Li and Wilson 1996, 1999; G.-Q. Li, Wilson, et al. 1997; Lavoué and Sullivan 2004; M. V. H. Wilson and Murray 2008; Inoue et al. 2009; Lavoué et al. 2011; Lavoue, Miya, Arnegard, et al. 2012; Lavoué 2015, 2016; Murray et al. 2018; Peterson et al. 2022).

  • Morphological phylogenies differ on the resolution of Osteoglossidae (bonytongues) with Pantodon buchholzi (Butterflyfish) as either the sister lineage of all other osteoglossids (Bonde 1996; M. V. H. Wilson and Murray 2008; Xu and Chang 2009) or nested within Osteoglossidae as the sister lineage of a clade containing Osteoglossum and Scleropages (Taverne 1979, 1998; G.-Q. Li and Wilson 1996, 1999; G.-Q. Li, Grande, et al. 1997; G.-Q. Li, Wilson, et al. 1997; Hilton 2003; Brito et al. 2020). Most molecular phylogenies resolve Pantodon as distantly related to other Osteoglossidae as the sister lineage of all other Osteoglossiformes (Lavoué and Sullivan 2004; Inoue et al. 2009; Lavoué et al. 2011; Lavoue, Miya, Arnegard, et al. 2012; Lavoué 2015, 2016; Hughes et al. 2018; Peterson et al. 2022).

  • Mormyridae is the most species-rich lineage of Osteoglossiformes, with at least 227 species classified in 22 genera (Fricke et al. 2023). Biodiversity discovery is active in mormyrids, as 11% of the living species diversity in the clade was described over the past 10 years (Sullivan et al. 2016; Fricke et al. 2023). Molecular phylogenies inferred from Sanger-sequenced nuclear and mitochondrial genes do not confidently resolve relationships within Mormyridae, but do strongly indicate that Brienomyrus, Hippopotamyrus, Marcusenius, and Pollimyrus are paraphyletic (Sullivan et al. 2000, 2016; Levin and Golubtsov 2018). Phylogenomic analyses of Mormyridae result in resolved and well-supported phylogenies where Hippopotamyrus and Marcusenius are paraphyletic, but Pollimyrus is monophyletic (Peterson et al. 2022). The Cretaceous fossil taxon †Palaeonotopterus greenwoodi is resolved as the sister lineage of Mormyroidea (Mormyridae and Gymnarchus) in phylogenetic analyses of morphological characters (Hilton 2003; Murray et al. 2018; Brito et al. 2020).

  • Morphological and molecular phylogenetic analyses resolve a monophyletic Notopteridae (e.g., Inoue et al. 2009; Brito et al. 2020). Within notopterids, the Asian (Chitala and Notopterus) and African (Papyrocranus and Xenomystus) lineages are each monophyletic and resolved as sister clades (Inoue et al. 2009). Morphological phylogenetic analyses resolved the Cretaceous fossil taxon †Laeliichthys ancestralis from Brazil as the sister lineage of Arapaiminae (Heterotis and Arapaima) (G.-Q. Li and Wilson 1996, 1999; Taverne 1998); however, a more recent morphological analysis places †Laeliichthys as the sister lineage of Notopteridae (Brito et al. 2020).

  • Composition. Osteoglossiformes currently contains 254 living species (Fricke et al. 2023) that include Pantodon buchholzi, Gymnarchus niloticus, and species classified in Mormyridae, Notopteridae, and Osteoglossidae (Hilton 2003). Fossil taxa include the pan-mormyroid †Palaeonotopterus and pan-notopterid †Laeliichthys (Silva Santos 1985; Lundberg 1993; Forey 1997; Cavin and Forey 2001; Murray et al. 2018; Brito et al. 2020). The ages and locations of †Palaeonotopterus and †Laeliichthys are given in Appendix 1. Over the past 10 years 15 new living species of Osteoglossiformes have been described (Fricke et al. 2023), comprising 5.9% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Osteoglossiformes include (1) 15 or fewer primary branched caudal fin rays (G.-Q. Li and Wilson 1996; G.-Q. Li, Grande, et al. 1997; G.-Q. Li, Wilson, et al. 1997; Hilton and Britz 2010), (2) two or fewer uroneurals in caudal skeleton (G.-Q. Li and Wilson 1996; Wiley and Johnson 2010), (3) nasal bone gutterlike or subrectangular (G.-Q. Li and Wilson 1996; G.-Q. Li, Grande, et al. 1997; G.-Q. Li, Wilson, et al. 1997), (4) six or fewer hypurals in caudal skeleton (G.-Q. Li, Grande, et al. 1997; Xu and Chang 2009), (5) dorsal hypurals and ural centrum 2 fused (G.-Q. Li, Wilson, et al. 1997; M. V. H. Wilson and Murray 2008; Hilton and Britz 2010; Wiley and Johnson 2010), (6) epurals absent (Hilton 2003; M. V. H. Wilson and Murray 2008; Hilton and Britz 2010), (7) bony elements associated with second ventral gill arch present as processes on second hypobranchial (Xu and Chang 2009), (8) presence of one ossified pair of hypohyals (Xu and Chang 2009), and (9) palatine and ectopterygoid fused (Xu and Chang 2009).

  • Synonyms. There are no synonyms of Osteoglossiformes.

  • Comments. The first delimitation of Osteoglossiformes included Hiodon and excluded Mormyridae (Greenwood et al. 1966:394). The hypothesis that Hiodon was nested in the clade delimited here as Osteoglossiformes was supported in several studies (G. J. Nelson 1968; Greenwood 1973; Lauder and Liem 1983). The delimitation of Osteoglossiformes that includes all living species of Osteoglossomorpha except Hiodon tergisus and H. alosoides was first proposed by Taverne (1979), and this hypothesis is corroborated in nearly all subsequent morphological and molecular phylogenetic analyses (e.g., G.-Q. Li and Wilson 1996; Lavoué and Sullivan 2004; Peterson et al. 2022).

  • The earliest fossil taxon of Osteoglossiformes is the pan-notopterid †Laeliichthys ancestralis from the Barremian (126.5–121.4 Ma) in the Cretaceous of Brazil (Silva Santos 1985; Brito et al. 2020). Bayesian relaxed molecular clock analyses of Osteoglossiformes result in an average posterior crown age estimate of 197.7 million years ago, with the credible interval ranging between 174.4 and 221.6 million years ago (Peterson et al. 2022).

  • img-z38-8_03.gif

    Osteoglossidae C. L. Bonaparte 1845:387
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Osteoglossum bicirrhosum (Cuvier 1829) and Heterotis niloticus (Cuvier 1829), but not Pantodon buchholzi Peters 1876. This is a minimum-crown-clade definition with an external specifier.

  • Etymology. From the ancient Greek ὀστέoν (̍αːstIәn), meaning bone, and γλῶσσα (ɡl̍ͻsә), meaning tongue.

  • Registration number. 897.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of complete mitochondrial genomes (Lavoué 2015, fig. 2). Phylogenetic relationships among the living lineages and fossil taxa of Osteoglossidae are presented in Figure 6. The placements of the fossil taxa in the phylogeny are on the basis of inferences from morphological characters (G.-Q. Li and Wilson 1996, 1999; J.-Y. Zhang 1998, 2006; Hilton 2003; Xu and Chang 2009; Murray et al. 2018).

  • Phylogenetics. Morphological and molecular analyses differ on the phylogenetic resolution of Pantodon buchholzi and Osteoglossidae. All morphological analyses resolve Pantodon either as the sister lineage of all other Osteoglossidae (G. J. Nelson 1969b; Greenwood 1973; Bonde 1996; M. V. H. Wilson and Murray 2008; Murray et al. 2018) or as the sister lineage of Osteoglossinae (G. J. Nelson 1968; Taverne 1979, 1998; G.-Q. Li and Wilson 1996, 1999; G.-Q. Li, Grande, et al. 1997; G.-Q. Li, Wilson, et al. 1997; Taverne 1998; J.-Y. Zhang 2006; Murray et al. 2018), which is a clade containing Osteoglossum and Scleropages (Hilton and Lavoué 2018). Molecular analyses agree with morphological studies in resolving two sets of sister lineages within Osteoglossidae: Osteoglossinae (Osteoglossum and Scleropages) and Arapaiminae (Arapaima and Heterotis); however, most molecular phylogenies place Pantodon as the sister lineage of all other Osteoglossiformes, distantly related to Osteoglossidae (Lavoué and Sullivan 2004; Inoue et al. 2009; Lavoué et al. 2011; Lavoue, Miya, Arnegard, et al. 2012; Lavoué 2015, 2016; Hughes et al. 2018; Peterson et al. 2022).

  • Composition. Osteoglossidae currently contains 12 living species (D. J. Stewart 2013a, 2013b; Fricke et al. 2023) that include Heterotis niloticus (African Arowana) and species classified in Arapaima, Osteoglossum, and Scleropages. Fossil lineages of Osteoglossidae include the pan-arapaimines †Joffrichthys and †Sinoglossus and the pan-osteoglossines †Cretophareodus, †Phareodus, and †Singida (G.-Q. Li and Wilson 1996, 1999; J.-Y. Zhang 1998, 2006; Hilton 2003; Xu and Chang 2009; Murray et al. 2018). The ages and locations of the fossil osteoglossids are given in Appendix 1. Over the past 10 years no new living species of Osteoglossidae have been described (Fricke et al. 2023).

  • Diagnostic apomorphies. Morphological apomorphies for Osteoglossidae include (1) six hypurals in caudal skeleton (G.-Q. Li and Wilson 1996), (2) opercle oval or kidney- shaped (G.-Q. Li and Wilson 1996; G.-Q. Li, Grande, et al. 1997), (3) palatoquadrate area behind and below orbit completely covered by infraorbitals (G.-Q. Li and Wilson 1996; G.-Q. Li, Grande, et al. 1997; Hilton 2003; J.-Y. Zhang 2006; M. V. H. Wilson and Murray 2008; Forey and Hilton 2010), (4) ventral part of preopercle does not reach level of orbit (G.-Q. Li, Wilson, et al. 1997; M. V. H. Wilson and Murray 2008), (5) basipterygoid process present (G.-Q. Li, Wilson, et al. 1997; Hilton 2003), (6) no connection between swim bladder and ear (G.-Q. Li, Wilson, et al. 1997; Forey and Hilton 2010), (7) supraorbital canal ending in frontal (G.-Q. Li and Wilson 1999; Forey and Hilton 2010), (8) extrascapular bone reduced and irregularly shaped (Hilton 2003; M. V. H. Wilson and Murray 2008; Forey and Hilton 2010), (9) nasal bones flat and broad (Hilton 2003; Forey and Hilton 2010), tooth plates of basibranchial and basihyal continuous (Hilton 2003), (10) subopercle small and anterior to opercle (Hilton 2003; J.-Y. Zhang 2006; Forey and Hilton 2010), (11) scales with reticulate furrows present over entire scale (Hilton 2003; Forey and Hilton 2010), (12) infraorbitals 3 and 4 fused (J.-Y. Zhang 2006), (13) temporal fossa present, bordered by epioccipital and pterotic (Xu and Chang 2009), and (14) first parapophysis expanded (Forey and Hilton 2010).

  • Synonyms. There are no synonyms of Osteoglossidae.

  • Comments. Bonaparte's (1845) introduction of the name Osteoglossidae was in a list of taxonomic names in a classification of fishes with no comment. Günther's (1868:377–380) delimitation of Osteoglossidae is identical to that presented here.

  • Osteoglossidae is a valid family-group name under the International Code of Zoological Nomenclature (Van der Laan et al. 2014:64). The earliest fossil taxon included in Osteoglossidae is †Cretophareodus alberticus from the Campanian (83.2–72.2 Ma) in the Cretaceous of Canada (G.-Q. Li 1996; Arbour et al. 2009). Bayesian relaxed molecular clock analyses of Osteoglossidae result in an average posterior crown age estimate of 96.6 million years ago, with the credible interval ranging between 78.6 and 112.7 million years ago (Peterson et al. 2022).

  • img-z40-3_03.gif

    Clupeocephala P. H. Greenwood 1973:326
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Clupea harengus Linnaeus 1758 (Otocephala, Clupeiformes), Engraulis encrasicolus (Linnaeus 1758) (Otocephala, Clupeiformes), Cyprinus carpio Linnaeus 1758 (Otocephala, Cypriniformes), Lepidogalaxias salamandroides Mees 1961 (Euteleostei), and Perca fluviatilis Linnaeus 1758 (Euteleostei, Perciformes). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek ϰλoυπαῖα (kluːpi͡ә), a name with an obscure origin for an uncertain number of fish species used by ancient authors such as Plutarch (D. W. Thompson 1947:117–118) and ϰεϕαλή (kεfˈαːlә), meaning the head of a human or other animal.

  • Registration number. 898.

  • Reference phylogeny. A phylogeny inferred from DNA sequences sampled from 1,105 exons (Hughes et al. 2018, fig. S2). Phylogenetic relationships among the major living lineages and fossil taxa of Clupeocephala are shown in Figure 7. The resolutions of the fossil taxa in the phylogeny are on the basis of inferences from morphology (Taverne 1981; Gayet 1994; Fielitz 2002; de Figueiredo and Gallo 2004; de Figueiredo 2005; Gallo et al. 2009; de Figueiredo, Gallo, and Delarmelina, et al. 2012a; Guinot and Cavin 2018).

  • Phylogenetics. Clupeocephala was identified as the clade containing all living teleosts to the exclusion of Elopomorpha and Osteoglossomorpha (Patterson and Rosen 1977). Monophyly of Clupeocephala is consistently supported; however, the delimitation of lineages within the clade and hypotheses of their relationships vary among molecular and morphological phylogenetic analyses (Lê et al. 1993; Lecointre 1995; G. D. Johnson and Patterson 1996; Lecointre and Nelson 1996; Arratia 1997, 2018; Ishiguro et al. 2003; Lavoué et al. 2005; Poulsen et al. 2009; Near, Eytan, et al. 2012; Faircloth et al. 2013; Hughes et al. 2018; Straube et al. 2018; Musilova et al. 2019; Roth et al. 2020; Mu et al. 2022). In addition to morphological and molecular phylogenetic analyses, the conservation of gene adjacency in the genome and the proportion of shared chromosomal breakpoints support monophyly of Clupeocephala (Parey et al. 2023).

  • Composition. Clupeocephala currently contains more than 33,675 living species (Fricke et al. 2023) classified in Otocephala and Euteleostei (Near, Eytan, et al. 2012; Dornburg and Near 2021). Fossil lineages of Clupeocephala include the pan-euteleosts †Avitosmerus, †Beurlenichthys, †Erichalcis, †Gaudryella, †Ghabouria, †Helgolandichthys, †Parawenzichthys, †Santanasalmo, †Scombroclupeoides, †Wenzichthys, and †Tchernovichthys (Taverne 1981; Gayet 1994; Fielitz 2002; de Figueiredo 2005; Gallo et al. 2009; de Figueiredo, Gallo, and Delarmelina, et al. 2012; Guinot and Cavin 2018). Details of the ages and locations of fossil lineages are presented in Appendix 1. Over the past 10 years there have been 3,518 new living species of Clupeocephala described, comprising 10.5% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Clupeocephala include (1) autopalatine bone ossifies early in ontogeny (Arratia 2010a), (2) hypohyals pierced by hyoidean arteries (Arratia 2010a), (3) tooth plate of cartilaginous fourth pharyngobranchial element forms by the growth of only one tooth plate (Arratia 2010a), (4) uroneurals not inclined toward horizontal plane, but aligned at different angles (Arratia 2010a), (5) angular and articular bones fused (Arratia 2010a), (6) retroarticular bone excluded from articular facet of quadrate (Arratia 2010a), (7) absence of tooth plates on pharyngobranchial 1 (Arratia 2010a), (8) absence of tooth plates on pharyngobranchial 2 (Arratia 2010a), (9) absence of tooth plates on pharyngobranchial 3 (Arratia 2010a), (10) six or fewer hypurals (Arratia 2010a), and (11) fusion of duplicated chromosomes 2a and 2b (Parey et al. 2023).

  • Synonyms. There are no synonyms of Clupeocephala.

  • Comments. In defining Clupeocephala as all living teleosts to the exclusion of Elopomorpha and Osteoglossomorpha, Patterson and Rosen (1977) provided a resolution to the long-standing uncertainly regarding the relationships of Clupeiformes that was left unresolved in Greenwood et al. (1966). The composition of Clupeocephala has not changed subsequent to its introduction by Patterson and Rosen (1977). The earliest fossil taxon in Clupeocephala that is not an otocephalan is the pan-euteleost †Tchernovichthys exspectatum from the Hauterivian (132.6–126.5 Ma) in the Cretaceous of Israel (Gayet 1994). Bayesian relaxed molecular clock analyses of Clupeocephala result in an average posterior crown age estimate of 224.8 million years ago, with the credible interval ranging between 210.8 and 236.6 million years ago (Hughes et al. 2018).

  • img-z42-4_03.gif

    FIGURE 7.

    Phylogenetic relationships of the major living lineages and fossil taxa of Clupeocephala, Euteleostei, Argentiniformes, Salmoniformes, Esocidae, Stomiatii, Osmeriformes, and Stomiiformes. Filled circles identify the common ancestor of clades, with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z41-1_03.jpg

    Otocephala G. D. Johnson and C. Patterson 1996:315
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Engraulis encrasicolus (Linnaeus 1758) (Clupeiformes), Gonorynchus greyi (Richardson 1845) (Gonorynchiformes), and Cyprinus carpio Linnaeus 1758 (Cypriniformes). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek ώτός (h̍o͡Ʊt̍o͡Ʊz), meaning of the ear (the genitive declension of όυς), and ϰεϕαλή (kεf̍αːlә), meaning the head of a human or other animal.

  • Registration number. 899.

  • Reference phylogeny. A maximum likelihood phylogeny inferred from DNA sequences sampled from whole mitochondrial genomes (Poulsen et al. 2009, fig. 2). Phylogenetic relationships among the major lineages of Otocephala are shown in Figure 8. The placement of the fossil lineages †Ellimmichthyiformes, †Santanaclupea, and †Tischlingerichthys in the phylogeny reflects inferences based on morphology (Arratia 1997; Taverne 1997a; Forey 2004; Zaragüeta-Bagils 2004; Diogo 2007; Alvarado-Ortega et al. 2008; Mayrinck et al. 2015a; Vernygora et al. 2016; Vernygora 2020; Marramà et al. 2023).

  • Phylogenetics. The first phylogenies supporting monophyly of Clupeocephala resolved Clupeiformes and Euteleostei as sister groups, with Ostariophysi included in Euteleostei (Patterson and Rosen 1977). The monophyly of Otocephala as a group containing Ostariophysi and Clupeiformes to the exclusion of Euteleostei was a discovery resulting from early molecular phylogenetic analyses of gnathostomes (Lê et al. 1993), but subsequently supported in several morphological phylogenetic analyses and reviews of morphological synapomorphies (Arratia 1996b, 1997, 1999; G. D. Johnson and Patterson 1996; Lecointre and Nelson 1996). Phylogenetic analyses of DNA sequences from whole mitochondrial genomes resulted in an unexpected expansion of Otocephala to include the deep-sea Alepocephaliformes (Ishiguro et al. 2003; Lavoué et al. 2005, 2007; Lavoué, Miya, Kawaguchi, et al. 2008), historically classified within Euteleostei as Argentiniformes (e.g., Greenwood and Rosen 1971; A. C. Gill and Mooi 2002; J. S. Nelson 2006:192–194). The monophyly of the expanded Otocephala and the resolution of Alepocephaliformes and Ostariophysi as sister lineages is supported in molecular phylogenetic analyses of nuclear genes, combinations of mitochondrial and nuclear genes, and a phylogenomic analysis of DNA sequences sampled from more than 800 exons (Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; W.-J. Chen et al. 2013; Straube et al. 2018). Investigations of morphological characters identify apomorphies consistent with the delimitation of Otocephala presented here (Arratia 2018; Straube et al. 2018).

  • Composition. Otocephala currently contains 12,270 living species (Fricke et al. 2023) classified in Alepocephaliformes, Clupeiformes, and Ostariophysi. Fossil lineages of Otocephala include the pan-clupeiforms †Ellimmichthyiformes and †Santanaclupea and the pan-ostariophysan †Tischlingerichthys (L. Grande 1985; Maisey 1993; Arratia 1997; M.-M. Chang and Maisey 2003; Zaragüeta-Bagils 2004; Alvarado-Ortega et al. 2008, 2020; de Figueiredo 2009; Murray and Wilson 2013; Vernygora 2020; Marramà et al. 2023). Over the past 10 years 1,729 new living species of Otocephala have been described (Fricke et al. 2023), comprising 14.1% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Otocephala include (1) parietals fused with extrascapulars, with an uncertain distribution in Alepocephaliformes (Lecointre and Nelson 1996; Arratia 2018; Straube et al. 2018), (2) anterior part of swim bladder with silvery peritoneum (Fink and Fink 1996; Straube et al. 2018), but Alepocephaliformes lack a swim bladder (Arratia 2018; Straube et al. 2018), and (3) haemal spines anterior of preural centrum 2 fuse with their centra from an early point in development (Arratia 2018; Straube et al. 2018).

  • Synonyms. Otomorpha (Wiley and Johnson 2010:134; Betancur-R et al. 2017:14–15) and Ostarioclupeomorpha (Arratia 1997:155) are ambiguous synonyms of Otocephala.

  • Comments. G. D. Johnson and Patterson (1996) applied the group name Otocephala to the clade containing Clupeiformes and Ostariophysi, which was initially discovered in one of the earliest molecular data to investigate teleost phylogeny (Lê et al. 1993). Molecular phylogenetic analyses led to the expansion of Otocephala to include Alepocephaliformes (Ishiguro et al. 2003; Near, Eytan, et al. 2012; Straube et al. 2018). The name Otocephala was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest fossil otocephalan lineages include the pan-ostariophysan †Tischlingerichthys from the Tithonian (149.2–143.1 Ma) in the Jurassic of Germany (Arratia 1997, 2001). Bayesian relaxed molecular clock analyses of Otocephala result in an average posterior crown age estimate of 194.5 million years ago, with the credible interval ranging between 179.7 and 211.2 million years ago (Hughes et al. 2018).

  • img-z44-8_03.gif

    FIGURE 8.

    Phylogenetic relationships of the major living lineages and fossil taxa of Otocephala, Clupeiformes, Clupeoidei, Alepocephaliformes, Ostariophysi, Gonorynchiformes, Otophysi, Gymnotiformes, Pan-Siluriformes, and Cithariniformes. Filled circles identify the common ancestor of clades, with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1. The clade description of Pan-Siluriformes is presented in Lundberg (2020c).

    img-z43-1_03.jpg

    Clupeiformes E. S. Goodrich 1909:386
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Denticeps clupeoides Clausen 1959, Clupea harengus Linnaeus 1758, and Engraulis encrasicolus (Linnaeus 1758). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek ϰλoυπαα (kl̍uːpi͡ә), a name with an obscure origin for an uncertain number of fish species used by ancient authors such as Plutarch (D. W. Thompson 1947:117–118). The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 900.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 1,165 exons (Q. Wang et al. 2022, fig. 2). Phylogenetic relationships among the living and fossil lineages of Clupeiformes are shown in Figure 8. The fossil lineages †Cynoclupea and †Paleodenticeps are placed in the phylogeny on the basis of inferences from morphology (Greenwood 1960, 1968; Malabarba and Di Dario 2017; Vernygora 2020).

  • Phylogenetics. Greenwood et al. (1966) delimited Clupeiformes to include Denticeps clupeoides and Clupeoidei, which is reflected in subsequent classifications (G. J. Nelson 1970b; L. Grande 1985; Lavoué, Konstantinidis, et al. 2014; J. S. Nelson et al. 2016:164–172; Betancur-R et al. 2017). A consistent result in morphological and molecular phylogenetic analyses of Clupeiformes is the resolution of Clupeoidei and Denticeps as sister groups (G. J. Nelson 1967, 1970b; Patterson and Rosen 1977; L. Grande 1985; Lavoué et al. 2007; de Pinna and Di Dario 2010; Lavoué, Konstantinidis, et al. 2014; Straube et al. 2018; Vernygora 2020; Milec et al. 2022; Q. Wang et al. 2022).

  • Composition. Clupeiformes currently contains 448 living species that include Denticeps clupeoides and species classified in Clupeoidei (Lavoué, Konstantinidis, et al. 2014; Q. Wang et al. 2022; Fricke et al. 2023). Fossil lineages of Clupeiformes include the pan-clupeoid †Cynoclupea (Malabarba and Di Dario 2017) and the pan-denticipitid †Paleodenticeps (Greenwood 1960). Details of the ages and locations of †Cynoclupea and †Paleodenticeps fossil taxa are given in Appendix 1. Over the past 10 years 43 new living species of Clupeiformes have been described (Fricke et al. 2023), comprising 9.6% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Clupeiformes include (1) presence of abdominal scutes (Whitehead 1962; Patterson 1970; L. Grande 1985; Wiley and Johnson 2010), (2) diverticulum of swim bladder penetrates exoccipital, expanding to form ossified bulla in prootic or pterotic (Greenwood et al. 1966; L. Grande 1985; Wiley and Johnson 2010), (3) presence of recessus lateralis where infraorbital canal merges with preopercular canal (Greenwood et al. 1966; Greenwood 1968; L. Grande 1982, 1985; T. Grande and de Pinna 2004; Zaragüeta-Bagils 2004), (4) supraoccipital completely separates parietals (Whitehead 1962; Patterson 1970; L. Grande 1982, 1985; Zaragüeta-Bagils 2004), (5) absence of basipterygoid process of parasphenoid (Zaragüeta-Bagils 2004), (6) third preural centrum with thin haemal spine (Zaragüeta-Bagils 2004), and (7) presence of sensory cephalic canal branch that originates at junction between extrascapular bone and recessus lateralis (Di Dario and de Pinna 2006).

  • Synonyms. Clupeomorpha (Greenwood et al. 1966:358–-361) and Clupei (Wiley and Johnson 2010:134–135; Betancur-R et al. 2017:15) are ambiguous synonyms of Clupeiformes.

  • Comments. When first delimited, Clupeiformes was a “purely artificial assemblage of lowly organised (sic) families (Goodrich 1909:386)” and included clupeiforms as well as lineages now classified as Elopomorpha, Osteoglossomorpha, Salmonidae, Gonorynchiformes, Alepocephaliformes, and Stomiiformes. Greenwood et al. (1966) dismantled the groups Isospondyli and Malacopterygii (e.g., Boulenger 1904a; Bigelow 1963), limiting Clupeiformes to Clupeoidei and Denticeps. The name Clupeiformes was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest fossil Clupeiformes is the pan-clupeoid †Cynoclupea from the Barremian-Aptian (129.4–113.0 Ma) in the Cretaceous of Brazil, which was initially placed as the sister lineage of a clade containing Chirocentridae and Engraulidae (Malabarba and Di Dario 2017). However, Engraulidae is the sister lineage of all other Clupeoidei, and Chirocentridae shares common ancestry with Pristigasteridae (Vernygora 2020). The shared character states with both Engraulidae and Pristigasteridae indicate †Cynoclupea is best resolved as a pan-clupeoid (Malabarba and Di Dario 2017). Bayesian relaxed molecular clock analyses of Clupeiformes result in an average posterior crown age estimate of 130.8 million years ago, with the credible interval ranging between 125.5 and 138.9 million years ago (Q. Wang et al. 2022).

  • img-z46-1_03.gif

    Clupeoidei P. Bleeker 1849:6
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Spratelloides gracilis (Temminck and Schlegel 1846), Clupea harengus Linnaeus 1758, and Engraulis encrasicolus (Linnaeus 1758), but not Denticeps clupeoides Clausen 1959. This is a minimum-crown-clade definition with an external specifier.

  • Etymology. From the ancient Greek ϰλoυπαῖα (kl̍uːpi͡ә), a name with an obscure origin for an uncertain number of fish species used by ancient authors such as Plutarch (D. W. Thompson 1947:117–118).

  • Registration number. 901.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 1,165 exons (Q. Wang et al. 2022, fig. 2). See Figure 8 for a phylogeny of the living and fossil lineages comprising Clupeoidei. The placements of fossil lineages in the phylogeny are based on inferences from morphology (Taverne 2002, 2004, 2007a, 2007b, 2011a; Marramà and Carnevale 2018).

  • Phylogenetics. Greenwood et al. (1966) grouped all living species of Clupeiformes in Clupeoidei except Denticeps clupeoides. On the basis of gill arch morphology, G. J. Nelson (1967, 1970b) delimited four lineages of Clupeoidei: Chirocentridae (wolf herrings), Clupeidae (shads and sardines), Engraulidae (anchovies), and Pristigasteridae (longfin herrings). Analyses of morphological characters and molecular phylogenetic studies consistently support the monophyly of Clupeoidei (L. Grande 1985; Di Dario 2004; Lavoué et al. 2007, 2013; Li and Ortí 2007; Lavoué, Miya, Kawaguchi, et al. 2008; de Pinna and Di Dario 2010; Bloom and Lovejoy 2014; Lavoué, Konstantinidis, et al. 2014; Bloom and Egan 2018; Egan et al. 2018; Milec et al. 2022; Q. Wang et al. 2022); however, the traditional delimitation of Clupeidae is not resolved as monophyletic (e.g., Lavoué, Konstantinidis, et al. 2014; Egan et al. 2018; Vernygora 2020). The lack of clupeid monophyly has prompted the recognition of the lineages Alosidae (shads), Dorosomatidae (gizzard shads), Dussumieriidae (round herrings), Ehiravidae (ehiravines), and Spratelloididae (small round herrings) (Bloom and Egan 2018; Vernygora 2020; Q. Wang et al. 2022). Traditionally, Dussumieriidae included Dussumieria and Etrumeus (Whitehead 1985; J. S. Nelson et al. 2016:170), but phylogenomic analysis resolves Dussumieria and Chirocentrus as sister lineages and Etrumeus as the sister lineage of all sampled species of Clupeidae (Q. Wang et al. 2022).

  • Molecular phylogenies inferred from combinations of Sanger-sequenced mitochondrial and nuclear genes and phylogenomic analysis of 1,165 exons resolve Spratelloididae as the sister lineage of all other Clupeoidei (Bloom and Egan 2018; Egan et al. 2018; Q. Wang et al. 2022). Morphological characters seem to support Chirocentridae and Engraulidae as sister groups (Di Dario 2009; Malabarba and Di Dario 2017; Vernygora 2020, figs. 6, 7), but molecular phylogenies place Chirocentrus as the sister lineage to Pristigasteridae (Bloom and Egan 2018; Egan et al. 2018; Vernygora 2020, figs. 6–9). A morphological phylogenetic analysis of 175 characters sampled from 101 clupeiform species resulted in unresolved relationships with poor node support (Vernygora 2020, figs. 6, 7).

  • Composition. Clupeoidei currently contains 447 living species (Fricke et al. 2023) classified in Alosidae, Chirocentrus, Clupeidae, Dorosomatidae, Dussumieria, Ehiravidae, Engraulidae, Pristigasteridae, and Spratelloididae (Q. Wang et al. 2022). Fossil clupeoids include the pan-clupeids †Italoclupea and †Lecceclupea (Taverne 2007a, 2011a), the pan-dussumieriids †Nardoclupea and †Portoselvaggioclupea (Taverne 2002, 2007b), and the pan-alosids †Eoalosa and †Pugliaclupea (Taverne 2004; Marramà and Carnevale 2018). Details regarding the ages and locations of the fossil taxa are given in Appendix 1. Over the past 10 years 43 new living species of Clupeoidei have been described (Fricke et al. 2023), comprising 9.6% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Clupeoidei include (1) first uroneural and first preural fused (L. Grande 1985; Zaragüeta-Bagils 2004), (2) relative size of first ural centrum reduced (L. Grande 1985; Zaragüeta-Bagils 2004), (3) absence of lateral line scales (L. Grande 1985; Vernygora 2020), (4) parhypural and first ural centrum separated (L. Grande 1985; Zaragüeta-Bagils 2004), (5) absence of a complete series of ventral scutes between isthmus and anus (Zaragüeta-Bagils 2004), (6) ventral limb of hyomandibula and quadrate separated by metapterygoid (Di Dario 2009; Vernygora 2020), (7) single row of gill rakers on first through third arches (de Pinna and Di Dario 2010), (8) close proximity of dorsal gill arch elements to the midline (de Pinna and Di Dario 2010), (9) second and third infrapharyngobranchials produced anteriorly as a narrow long process (de Pinna and Di Dario 2010), and (10) presence of notch on third hypural (Vernygora 2020).

  • Synonyms. There are no synonyms of Clupeoidei.

  • Comments. Clupeoidei is among the most economically important lineages of fishes (FAO 2020). The generation of phylogenomic datasets that include hundreds of clupeoid species is a major priority for future teleost phylogenetics. This priority goes beyond the inherent interest in resolving this portion of the tree of life and is justified by the clade's economic importance and growing conservation concerns (FAO 2020; Birge et al. 2021).

  • Fossil taxa phylogenetically nested within crown subclades of Clupeoidei include †Knightia eocaena in Clupeidae, †Chasmoclupea aegyptica and †Trollichthys bolcensis in Spratelloididae, and †Eoengraulis fasoloi in Engraulidae (Vernygora 2020). The earliest fossil lineage of Clupeoidei is †Audenaerdia casieri with an uncertain phylogenetic resolution with Clupeidae or Alosidae from the Santonian (85.7–83.2 Ma) in the Cretaceous (Taverne 1997a, 1997b). The earliest Clupeoidei fossil lineages with a more confident phylogenetic resolution include the pan-clupeids †Italoclupea and †Lecceclupea (Taverne 2007a, 2011a), the pan-dussumieriids †Nardoclupea and †Portoselvaggioclupea (Taverne 2007b), and the pan-alosid †Pugliaclupea (Taverne 2004) from the Campanian-Maastrichtian (83.26–66.0 Ma) in the Cretaceous. Bayesian relaxed molecular clock analyses of Clupeoidei result in an average posterior crown age estimate of 91.4 million years ago, with the credible interval ranging between 76.1 and 107.3 million years ago (Q. Wang et al. 2022).

  • img-z47-6_03.gif

    Alepocephaliformes N. B. Marshall 1962:265
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Alepocephalus rostratus Risso 1820, Alepocephalus bairdii Goode and Bean 1879, Bathylaco nigricans Goode and Bean 1896, and Platytroctes apus Günther 1878b. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek λεπἴς (l̍εpIs), meaning the scale of a fish, with the prefix “a” meaning without scales and ϰεϕαλή (kεf̍αːlә) meaning the head of a human or other animal. The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 902.

  • Reference phylogeny. A phylogeny of Alepocephaliformes inferred from DNA sequences of complete mitochondrial genomes (Poulsen et al. 2009, fig. 3). Although Alepocephalus rostratus is not included in the reference phylogeny, it clusters with Xenodermichthys copei as the only sampled species of Alepocephaliformes in a DNA barcoding study (Landi et al. 2014, fig. S1). Phylogenetic relationships of Alepocephaliformes are shown in Figure 8.

  • Phylogenetics. The phylogenetic placement of Alepocephaliformes within Teleostei has shifted substantially over the past century, previously being grouped with Clupeiformes (Gregory and Conrad 1936), a delimitation of Salmoniformes that includes Salmonidae, Argentiniformes, Galaxiidae, Osmeriformes, Stomiiformes, and Esocidae (Greenwood et al. 1966; Markle 1976), and Osmeriformes (Gosline 1969). Greenwood and Rosen (1971) hypothesized Alepocephaliformes and Argentiniformes are sister lineages on the basis of a modified posterior pharyngobranchial structure they named the crumenal organ, which was the basis for the resolution of this clade in subsequent morphological studies (Begle 1992; G. D. Johnson and Patterson 1996). However, Ahlstrom et al. (1984) rejected the hypothesized common ancestry of Alepocephaliformes and Argentiniformes because alepocephaliform species hatch from larger eggs and exhibit direct development; additionally, the two lineages share no unique ontogenetic characters. Molecular phylogenetic analyses consistently resolve Alepocephaliformes in a clade with Clupeiformes and Ostariophysi (Ishiguro et al. 2003; Lavoué, Miya, Poulsen, et al. 2008; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Straube et al. 2018), prompting the classification of these three clades in Otocephala.

  • Two morphological analyses of relationships within Alepocephaliformes result in very different phylogenetic trees, with all alepocephaliforms classified as Alepocephalidae (slickheads) in Begle (1992) and the resolution of Alepocephalidae and Platytroctidae (tubeshoulders) in G. D. Johnson and Patterson (1996). Molecular phylogenetic analyses of DNA sequences from whole mitochondrial genomes or combinations of mitochondrial and nuclear genes resolve Alepocephalidae and Platytroctidae each as monophyletic groups (Lavoué, Miya, Poulsen, et al. 2008; Poulsen et al. 2009; Betancur-R et al. 2017; Rabosky et al. 2018; Poulsen 2019), but one molecular analysis resulted in Platytroctidae nested within Alepocephalidae (Betancur-R et al. 2017).

  • Composition. There are currently 142 living species of Alepocephaliformes classified in Alepocephalidae and Platytroctidae (Fricke et al. 2023). Over the past 10 years two new living species of Alepocephaliformes have been described, accounting for 1.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. The apomorphies of Alepocephaliformes are uncertain because of the reductive nature of morphological characters in the lineages and the fact that all morphological phylogenetic analyses assumed a relationship with Argentiniformes (Begle 1992; G. D. Johnson and Patterson 1996). Morphological apomorphies for Alepocephaliformes include (1) separation of parietals by supraoccipital (Greenwood and Rosen 1971; G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (2) absence of posttemporal fossa (Gosline 1969; G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (3) presence of branchiostegal cartilages (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (4) reduction of dorsal portion of opercle (Begle 1992; G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (5) forward extension of ossified epipleural series to third vertebra (Patterson and Johnson 1995; G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (6) absence of urodermal (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), and (7) presence of single postcleithrum (Markle 1976; G. D. Johnson and Patterson 1996; Wiley and Johnson 2010).

  • Synonyms. Alepocephaloidei (Bleeker 1859:xxx; Wiley and Johnson 2010:141), Alepocephaloidea (Greenwood and Rosen 1971:39–40; Begle 1992:351; G. D. Johnson and Patterson 1996:312), and Alepocephali (Betancur-R et al. 2017:15) are ambiguous synonyms of Alepocephaliformes. Alepocephaloidei and Bathylaconoidei (Greenwood et al. 1966:394) are approximate synonyms of Alepocephaliformes.

  • Comments. Marshall (1962:265) applied the name Alepocephaliformes to the lineage comprising Alepocephalidae and Searsiidae, a synonym of Platytroctidae (Parr 1951; Van der Laan et al. 2014:58). Long considered a subclade of Argentiniformes (Greenwood and Rosen 1971; Begle 1992; G. D. Johnson and Patterson 1996), Alepocephaliformes is now placed with Clupeiformes and Ostariophysi in Otocephala (Near, Eytan, et al. 2012; Arratia 2018; Straube et al. 2018). The name Alepocephaliformes was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The fossil record of Alepocephaliformes is relatively poor when compared with other lineages of Otocephala. The earliest skeletal fossils of Alepocephaliformes date to the Rupelian (33.9–28.1 Ma) in the Oligocene and the earliest otoliths are from the Ypresian (56.0–47.8 Ma) in the Eocene (Přikryl and Carnevale 2019). A maximum likelihood relaxed molecular clock analysis of Alepocephaliformes resulted in a crown age estimate of 38.8 million years ago (Rabosky et al. 2018).

  • img-z49-3_03.gif

    Ostariophysi M. Sagemehl 1885:22
    (as Ostariophysen) [Lundberg 2020]

  • Definition. Defined as a minimum-crown-clade in Lundberg (2020a) as: “The crown clade originating in the most recent common ancestor of Gonorynchus (originally Cyprinus) gonorynchus (Linnaeus 1766), Cyprinus carpio Linnaeus 1758 (Cypriniformes), Charax (originally Salmo) gibbosus (Linnaeus 1758) (Characiformes), Gymnotus carpio Linnaeus 1758 (Gymnotiformes; Gymnotoidei on the reference phylogeny), and Silurus glanis Linnaeus 1758 (Siluriformes; Siluroidei on the reference phylogeny).”

  • Etymology. From the ancient Greek ὀστἀριoν (ho͡Ʊst̍α͡ɹɹi͡әn), meaning little bone, and ϕῦσα (f̍uːsә), meaning bladder.

  • Registration number. 196.

  • Reference phylogeny. Fink and Fink (1981, fig. 1) was designated as the reference phylogeny by Lundberg (2020a). Phylogenetic relationships of the living and fossil lineages of Ostariophysi are presented in Figure 8. The placement of the pan-otophysan fossil lineages †Chanoides, †Clupavus, †Lusitanichthys, †Nardonoides, and †Santanichthys are on the basis of inferences from morphology (Patterson 1984a; Taverne 1995; Diogo et al. 2008; Diogo 2009; Malabarba and Malabarba 2010; Mayrinck 2011).

  • Phylogenetics. The monophyly of Ostariophysi as a lineage that includes Gonorynchiformes and Otophysi was first inferred from the morphology of the caudal skeleton and cervical vertebrae (Rosen and Greenwood 1970), a conclusion not universally accepted at the time (T. R. Roberts 1973). Subsequent summaries and phylogenetic analyses of morphological characters consistently resolve Ostariophysi as monophyletic (Fink and Fink 1981, 1996; Patterson 1984a, 1984b, 1994; Arratia 1999, 2000a, 2008, 2010b; Diogo et al. 2008). Early molecular phylogenetic analyses of whole mitochondrial genomes resolved Gonorynchiformes and Clupeiformes as sister lineages (Ishiguro et al. 2003; Saitoh et al. 2003; Peng et al. 2006). However, the monophyly of Ostariophysi is supported in all subsequent molecular phylogenetic studies that include analyses of whole mitochondrial genomes (Lavoué et al. 2005, 2007, 2010; Jondeung et al. 2007; Lavoué, Miya, Poulsen, et al. 2008; Poulsen et al. 2009; Nakatani et al. 2011; Davis et al. 2013), collections of Sanger-sequenced mitochondrial or nuclear genes (Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; W.-J. Chen et al. 2013), and phylogenomic datasets (Arcila et al. 2017; Chakrabarty et al. 2017; Dai et al. 2018; Hughes et al. 2018; Straube et al. 2018; Mu et al. 2022).

  • Composition. Ostariophysi currently contains 11,682 species (Fricke et al. 2023) classified in Gonorynchiformes and Otophysi. Fossil ostariophysans include the pan-otophysans †Chanoides, †Clupavus, †Lusitanichthys, †Nardonoides, and †Santanichthys (Patterson 1984a; Fink and Fink 1996; Cavin 1999; Filleul and Maisey 2004; Diogo et al. 2008; Malabarba and Malabarba 2010; Mayrinck 2011; Mayrinck et al. 2015b). Details regarding the ages and locations of the fossil taxa are listed in Appendix 1. Over the past 10 years there have been 1,686 new living species of Ostariophysi described (Fricke et al. 2023), comprising 14.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Ostariophysi include (1) sacculi and lagena situated posteriorly and along the midline (Rosen and Greenwood 1970; Fink and Fink 1981, 1996; Wiley and Johnson 2010; Lundberg 2020a), (2) swim bladder divided into small anterior and large posterior chamber (Rosen and Greenwood 1970; Fink and Fink 1981, 1996; Wiley and Johnson 2010), (3) anterior chamber of swim bladder covered with silvery peritoneal tunic (Rosen and Greenwood 1970; Fink and Fink 1981, 1996; Lundberg 2020a), (4) peritoneal tunic covering anterior chamber of swim bladder attached to two most anterior pleural ribs (Rosen and Greenwood 1970; Fink and Fink 1981, 1996; Wiley and Johnson 2010; Lundberg 2020a), (5) absence of basisphenoid (Fink and Fink 1981, 1996; Wiley and Johnson 2010; Lundberg 2020a), (6) absence of dermopalatine (Fink and Fink 1981, 1996; Wiley and Johnson 2010; Lundberg 2020a), (7) absence of supramaxillae (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (8) dorsal mesentery suspending swim bladder thickened anterodorsally (Fink and Fink 1981, 1996; Wiley and Johnson 2010; Lundberg 2020a), (9) absence of supraneural or accessory neural arch anterior to first vertebra (Fink and Fink 1981, 1996; Lundberg 2020a), (10) presence of expanded anterior neural arches that form roof over neural canal (Fink and Fink 1981, 1996; Wiley and Johnson 2010; Lundberg 2020a), (11) absence of neural arch anterior to arch of first vertebral centrum (Fink and Fink 1981, 1996; Wiley and Johnson 2010; Lundberg 2020a), (12) haemal spines anterior to second preural centrum fused to centrum (Fink and Fink 1981, 1996), (13) presence of Schreckstoff pheromone, an alarm substance produced by epidermal cells, stimulating a fright reaction in conspecifics (Pfeiffer 1977; Fink and Fink 1981, 1996; Wiley and Johnson 2010; Lundberg 2020a), and (14) absence of supraneural 1 and its cartilaginous precursor (Hoffmann and Britz 2006; Wiley and Johnson 2010).

  • Synonyms. There are no synonyms of Ostariophysi.

  • Comments. Sagemehl (1885:22) applied the name Ostariophysen to a group consisting of Cypriniformes, Gymnotiformes, Siluriformes, Characiformes, and Cithariniformes, which are classified here as Otophysi. This more exclusive definition of Ostariophysi was maintained for nearly a century (Boulenger 1904b:573–596; Goodrich 1909:371; Regan 1911d, 1911e; Jordan 1923:134–153; Greenwood et al. 1966; Gosline 1971:120–124). Citing several shared morphological traits, Rosen and Greenwood (1970) expanded Ostariophysi to include Gonorynchiformes and placed Cypriniformes, Gymnotiformes, Siluriformes, and Characiformes (sensu lato) in Otophysi. The earliest fossil Ostariophysi is the pan-otophysan †Santanichthys diasii from the Aptian-Albian (121.4–100.5 Ma) in the Cretaceous of Brazil. Bayesian relaxed molecular clock analyses of Ostariophysi result in an average posterior crown age estimate of 160.6 million years ago, with the credible interval ranging between 154.2 and 169.6 million years ago (Hughes et al. 2018).

  • img-z50-5_03.gif

    Gonorynchiformes P. H. Greenwood,
    D. E. Rosen, S. H. Weitzman, and G. S. Myers
    1966:374 [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Gonorynchus gonorynchus (Linnaeus 1766), Gonorynchus greyi (Richardson 1845), Chanos chanos [Fabricius in Niebuhr (ex Forsskål) 1775], and Kneria paucisquamata Poll and Stewart 1975. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek γωνία (ɡˈo͡Ʊniә), meaning angle, and ῤυγχoς (ɹˈuːɡko͡Ʊz), meaning snout or beak. The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 903.

  • Reference phylogeny. A time-calibrated phylogeny inferred from morphological characters and nine Sanger-sequenced nuclear genes (Near, Dornburg, and Friedman 2014, fig. 4). Although Gonorynchus gonorynchus is not included in the reference phylogeny, it resolves in a clade with four other species of Gonorynchus, including G. greyi, in a phylogenetic analysis of morphological characters (T. Grande 1999, fig. 10). Phylogenetic relationships among living and fossil lineages of Gonorynchiformes are shown in Figure 8. The placement of fossil lineages in the phylogeny is on the basis of analyses of morphological characters (Gayet 1993; T. Grande 1994, 1996; T. Grande and Poyato-Ariza 1999; Poyato-Ariza et al. 2010; Near, Dornburg, and Friedman 2014; Ribeiro, Poyato-Ariza, et al. 2018).

  • Phylogenetics. The first studies of relationships within Gonorynchiformes differed as to the earliest divergences in the clade. Greenwood et al. (1966) hypothesized that Gonorynchus was the likely sister lineage to all other gonorynchiforms, but Rosen and Greenwood (1970) argued that Chanos is the least morphologically specialized lineage of Gonorynchiformes and presented a classification reflecting this hypothesis.

  • Phylogenetic analyses using morphological characters have frequently included fossil lineages and all resolve Chanos as the living sister lineage of all other Gonorynchiformes (Patterson 1984b; Blum 1991; Gayet 1993; T. Grande 1994, 1996; T. Grande and Poyato-Ariza 1995, 1999; Poyato-Ariza 1996b; G. D. Johnson and Patterson 1997; Poyato-Ariza et al. 2010; Amaral and Brito 2012; Amaral et al. 2013; Ribeiro, Poyato-Ariza, et al. 2018). A diversity of molecular phylogenetic analyses of mtDNA, Sanger-sequenced nuclear genes, and phylogenomic datasets resolve Gonorynchus as the sister lineage of all other Gonorynchiformes (Lavoué et al. 2005; Lavoué, Miya, Moritz, et al. 2012; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Davis et al. 2013; Chakrabarty et al. 2017; Straube et al. 2018). Morphological and molecular phylogenetic analyses consistently support Phractolaemus ansorgii and all other species of Kneriidae as sister taxa (T. Grande 1994; G. D. Johnson and Patterson 1997; T. Grande and Poyato-Ariza 1999; Lavoué et al. 2005; Lavoué, Miya, Moritz, et al. 2012; Davis et al. 2013; Near, Dornburg, and Friedman 2014). Relationships among the five living species of Gonorynchus were resolved in a phylogenetic analysis of 12 morphological characters (T. Grande 1999).

  • Composition. Gonorynchiformes currently contains 40 living species (Fricke et al. 2023; Kalumba et al. 2023) that include Chanos chanos and species classified in Gonorynchus and Kneriidae (T. Grande and Poyato-Ariza 1999). Fossil lineages of Gonorynchiformes include the pan-chanids †Aethalionopsis, †Dastilbe, †Gordichthys, †Parachanos, †Rubiesichthys, and †Tharrhias (Poyato-Ariza 1994, 1996a; Fara et al. 2010). Fossil pan-gonorynchid lineages include †Charitopsis, †Charitosomus, †Hakeliosomus, †Judeichthys, †Notogoneus, and †Ramallichthys (L. Grande and Grande 1999; Fara et al. 2010). Details of the ages and locations of the fossil taxa are given in Appendix 1. One new living species of Gonorynchiformes has been described over the past 10 years, comprising 2.6% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Gonorynchiformes include (1) bone and cartilage of interorbital septum reduced where orbitosphenoid absent (Fink and Fink 1981, 1996; Patterson 1984b; T. Grande and Poyato-Ariza 1995; Poyato-Ariza et al. 2010; Wiley and Johnson 2010), (2) parietals reduced in size to canal-bearing ossicles (Rosen and Greenwood 1970; Fink and Fink 1981; Patterson 1984b; Poyato-Ariza et al. 2010; Wiley and Johnson 2010), (3) middle region of suspensorium, bounded by articular condyle for quadrate and hyomandibular, longer relative to height of suspensorium and opercular series (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (4) premaxilla thin and flat (Fink and Fink 1981, 1996; Patterson 1984b), (5) presence of bilateral pouches in branchial chamber located posterior to fourth epibranchial (Greenwood et al. 1966; Fink and Fink 1981, 1996; Wiley and Johnson 2010), (6) absence of teeth on fifth ceratobranchial (Fink and Fink 1981, 1996; Patterson 1984b; T. Grande and Poyato-Ariza 1999; Poyato-Ariza et al. 2010; Wiley and Johnson 2010), (7) anterior neural arch large, forming tight joint with exoccipital or exoccipital and supraoccipital (Fink and Fink 1981, 1996; Patterson 1984b; T. Grande and Poyato-Ariza 1999; Poyato-Ariza et al. 2010; Wiley and Johnson 2010), (8) presence of epicentral bones, also referred to as cephalic ribs (Patterson and Johnson 1995; Fink and Fink 1996; T. Grande and Poyato-Ariza 1999; Wiley and Johnson 2010), (9) absence of Baudelot's ligament (Patterson and Johnson 1995; Fink and Fink 1996; Wiley and Johnson 2010), (10) presence of exceptionally long esophagus (Fink and Fink 1996; Wiley and Johnson 2010), (11) pterosphenoids either slightly reduced, not articulating anteroventrally but in close proximity anterodorsally or greatly reduced and well-separated both anteroventrally and anterodorsally (T. Grande and Poyato-Ariza 1999; Poyato-Ariza et al. 2010), (12) parietals partially or completely separated by supraoccipital (T. Grande and Poyato-Ariza 1999; Poyato-Ariza et al. 2010), (13) ascending process of premaxilla absent (T. Grande and Poyato-Ariza 1999; Poyato-Ariza et al. 2010), (14) maximum height of dentary at midpoint or at anterior region close to symphysis (T. Grande and Poyato-Ariza 1999), (15) fewer than five infraorbitals (T. Grande and Poyato-Ariza 1999), (16) anterior neural arches slightly in contact with adjacent arches or exhibit overlapping lateral contact with adjacent arches (T. Grande and Poyato-Ariza 1999), (17) rib on third vertebral centrum wider and shorter than posterior ribs (T. Grande and Poyato-Ariza 1999; Poyato-Ariza et al. 2010), and (18) premaxilla, maxilla, and dentary without teeth (Poyato-Ariza et al. 2010).

  • Synonyms. Gonorhynchoidei (Gosline 1960:357, 1971:113–114), Anotophysi (Rosen and Greenwood 1970:23), and Anotophysa (Betancur-R et al. 2017:15) are ambiguous synonyms of Gonorynchiformes.

  • Comments. Gosline (1960, 1971:113–114) was the first investigator to delimit a group containing Gonorynchus, Chanos chanos, Cromeria, Kneria, and Phractolaemus ansorgii, which he named Gonorhynchoidei (sic). Greenwood et al. (1966) included the kneriid Grasseichthys gabonensis and named the group Gonorynchiformes. The name Gonorynchiformes was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest fossil Gonorynchiformes is the pan-chanid †Rubiesichthys gregalis from the Berriasian and Valanginian (143.1–132.6 Ma) in the Cretaceous of Spain (Poyato-Ariza 1996a). Bayesian relaxed molecular clock analyses of Gonorynchiformes using fossil tip-dating result in an average posterior crown age estimate of 219.8 million years ago, with the credible interval ranging between 201.7 and 240.0 million years ago (Near, Dornburg, and Friedman 2014).

  • img-z52-6_03.gif

    Otophysi W. Garstang 1931:253, 256
    [Lundberg 2020]

  • Definition. Defined as a minimum-crown-clade in Lundberg (2020a) as: “The crown clade originating in the most recent common ancestor of Cyprinus carpio Linnaeus 1758 (Cypriniformes), Charax (originally Salmo) gibbosus (Linnaeus 1758) (Characiformes), Gymnotus carpio Linnaeus 1758 (Gymnotiformes; Gymnotoidei on the reference phylogeny), and Silurus glanis Linnaeus 1758 (Siluriformes; Siluroidei on the reference phylogeny).”

  • Etymology. From the ancient Greek γτός (hˈo͡Ʊtˈo͡Ʊz), meaning belonging to the ear, and ϕῦσα (fˈuːsә), meaning bladder.

  • Registration number. 197.

  • Reference phylogeny. Fink and Fink (1981, fig. 1) was designated as the primary reference phylogeny by Lundberg (2020a). Phylogenetic relationships of the major lineages of Otophysi are presented in Figure 8. The placements of the pan-siluriform †Andinichthyidae and the pan-citharinid †Eocitharinus in the phylogeny are based on inferences from morphology (Arratia and Gayet 1995; Gayet and Meunier 2003; Murray 2003; Guinot and Cavin 2018).

  • Phylogenetics. Phylogenetic analyses of morphological characters resolve Otophysi as monophyletic and place Cypriniformes as the sister group of a clade containing Characiformes (sensu lato), Siluriformes, and Gymnotiformes (Fink and Fink 1981, 1996; Arratia 1992; Diogo et al. 2008). The monophyly of Otophysi is consistently supported by molecular phylogenetic studies, including analyses of whole mitochondrial genomes (Lavoué et al. 2005; Jondeung et al. 2007; Poulsen et al. 2009; Nakatani et al. 2011), collections of Sanger-sequenced mitochondrial or nuclear genes (Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; W.-J. Chen et al. 2013), and phylogenomic datasets (Arcila et al. 2017; Chakrabarty et al. 2017; Dai et al. 2018; Hughes et al. 2018; Straube et al. 2018; Faircloth et al. 2020). Within Otophysi, morphological and molecular phylogenies are incongruent with regards to the relationships of Characiformes (sensu lato), Siluriformes, and Gymnotiformes. Specifically, the traditional delimitation of Characiformes that includes Cithariniformes is not resolved as monophyletic relative to Siluriformes or Gymnotiformes in phylogenetic studies ranging from the early single locus analyses in the mid-1990s to phylogenomic analyses in the early 21st century (Ortí and Meyer 1996, 1997; Nakatani et al. 2011; Betancur-R, Broughton, et al. 2013; W.-J. Chen et al. 2013; Chakrabarty et al. 2017; Dai et al. 2018; Hughes et al. 2018, fig. S2; Faircloth et al. 2020; Simion et al. 2020; Melo, Sidlauskas, et al. 2022; Yang et al. 2023).

  • Composition. Otophysi currently contains more than 11,640 species (Fricke et al. 2023) classified in Characiformes, Cithariniformes, Cypriniformes, Gymnotiformes, and Siluriformes. Fossil otophysans include the Pan-Siluriformes lineage †Andinichthyidae (Gayet 1988a, 1990; Arratia and Gayet 1995; Gayet and Meunier 1998, 2003; Bogan et al. 2018) and the pan-cithariniform †Eocitharinus macrognathus (Murray 2003). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years there have been 1,683 new living species of Otophysi described (Fricke et al. 2023), comprising 14.5% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Otophysi include (1) axe-shaped endochondral portion of metapterygoid (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (2) first or first and second anterior supraneurals with ventral expansion that forms synchondral joint with neural arches (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (3) scaphium and claustrum of Weberian apparatus present (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (4) reduction of second neural arch that is modified into intercalarium (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (5) centra of anterior vertebrae shortened (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (6) fusion of first two parapophyses and centra (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (7) presence of the tripus, a bone that is an element of the Weberian apparatus and is likely a modified pleural rib (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (8) presence of the os suspensorium (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (9) pelvic bone bifurcated (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (10) presence of compound terminal vertebrae (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (11) hypural 2 fused with compound centrum (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (12) the sinus impar of inner ear present (Fink and Fink 1981, 1996; Wiley and Johnson 2010), (13) loss of supradorsal 2 and all supradorsals posterior to vertebra 4 (Hoffmann and Britz 2006; Wiley and Johnson 2010), and (14) fusion of supradorsals 3 and 4 with supraneural 2 and 3 cartilages to form neural complex (Hoffmann and Britz 2006; Wiley and Johnson 2010).

  • Synonyms. Ostariophysen (Sagemehl 1885:22), Ostariophysi (Boulenger 1904b:573–596; Goodrich 1909:371; Regan 1911d:13–15, 1911e:554; Jordan 1923:134–153; Greenwood et al. 1966:380–382, 395–396; Gosline 1971:120–124), and Plectospondyli (Cope 1871a:454; Jordan 1923:134–153) are approximate synonyms of Otophysi. Cypriniformes is an ambiguous synonym of Otophysi (Bertin and Arambourg 1958:2285–2287; McAllister 1968:67–78).

  • Comments. Garstang (1931) delimited a more inclusive Otophysi that in addition to Siluriformes and Characiformes included Osteoglossiformes, Elopiformes, and Clupeiformes. Sagemehl (1885) applied the name Ostariophysen to a group now delimited as Otophysi. Rosen and Greenwood (1970) expanded Ostariophysi to include Gonorynchiformes, and placed Cypriniformes, Gymnotiformes, Siluriformes, and Characiformes (sensu lato) in Otophysi. The name Otophysi was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • Bayesian relaxed molecular clock analyses of Otophysi result in an average posterior crown age estimate of 146.9 million years ago, with the credible interval ranging between 137.9 and 156.5 million years ago (Hughes et al. 2018).

  • img-z54-4_03.gif

    Cypriniformes E. S. Goodrich 1909:371
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Catostomus catostomus (Forster 1773), Gyrinocheilus pustulosus Vaillant 1902, Cobitis taenia Linnaeus 1758, Cyprinus carpio Linnaeus 1758, and Paedocypris progenetica Kottelat et al. 2006. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek κυπρῖνoς (kuːpɹˈ̍iːno͡Ʊz), frequently applied to the European Carp, Cyprinus carpio (D. W. Thompson 1947:135–136). The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 904.

  • Reference phylogeny. A phylogeny of 1,703 species of Cypriniformes inferred from a supermatrix of 27 nuclear and mitochondrial genes (Rabosky et al. 2018; J. Chang et al. 2019). The phylogeny is available on the Dryad data repository (Rabosky et al. 2019). Although the reference phylogeny does not include Paedocypris progenetica, the species resolves within the cyprinoid clade Danionidae (danionins) in phylogenetic analysis of mtDNA, combined mtDNA and nuclear gene sequences, and morphological characters (Rüber et al. 2007; Fang et al. 2009; K. L. Tang et al. 2010; Britz, Conway, et al. 2014). Phylogenetic analysis of DNA sequences of nuclear genes resolves Paedocypris as the sister lineage of Cyprinoidei or Cypriniformes (Mayden and Chen 2010; Stout et al. 2016; Malmstrøm et al. 2018). Phylogenetic relationships of living and fossil lineages of Cypriniformes are shown in Figure 9. The resolution of †Jianghanichthys in the phylogeny is based on analysis of morphological characters (J. Liu et al. 2015).

  • Phylogenetics. Greenwood et al. (1966) argued for monophyly of Cypriniformes on the basis of morphological characters from the pharyngeals, skull, oral jaws, vertebrae, and Weberian apparatus. X.-W. Wu et al. (1981) mapped morphological character changes onto a phylogeny that included Cyprinoidei and the cobitoid subclade Balitoridae (hillstream loaches) as sister lineages, a relationship that is not supported in any subsequent study of cypriniform phylogeny. Analysis of discretely coded morphological characters consistently resolves Cyprinoidei as the sister lineage of a clade containing Gyrinocheilus (algae eaters), Catostomidae (suckers), and Cobitoidei (Siebert 1987; Conway and Mayden 2007; Conway 2011). Inferred relationships differ among morphological analyses, with Gyrinocheilus and Catostomidae as successive sister lineages to Cobitoidei (Siebert 1987; Conway and Mayden 2007) or Gyrinocheilus and Catostomidae resolved as a clade that is the sister lineage of Cobitoidei (Conway 2011; Mabee et al. 2011; Britz, Conway, et al. 2014). Morphological phylogenetic analyses that include the Eocene-aged †Jianghanichthys result in a set of 116 most parsimonious trees. The strict consensus of these trees resolves the most recent common ancestor of Cypriniformes as a polytomy subtending Gyrinocheilus, Catostomidae, Cobitoidei, Cyprinoidei, and †Jianghanichthys (J. Liu et al. 2015).

  • The monophyly of Cypriniformes is supported in a range of molecular phylogenetic studies that include analyses of whole mitochondrial genomes (Saitoh et al. 2006, 2011; Jondeung et al. 2007; He, Gu, et al. 2008; Poulsen et al. 2009; Nakatani et al. 2011), trees inferred from collections of Sanger-sequenced mitochondrial or nuclear genes (Mayden et al. 2008, 2009; Mayden and Chen 2010; Betancur-R et al. 2017; Hirt et al. 2017; Luo et al. 2023), and analysis of phylogenomic datasets (Stout et al. 2016; Hughes et al. 2018). Molecular phylogenetic analyses uniformly resolve Catostomidae, Cobitoidei, Cyprinoidei, and Gyrinocheilus as monophyletic (Šlechtová et al. 2007; W.-J. Chen et al. 2009; Mayden and Chen 2010; Stout et al. 2016; Hirt et al. 2017; Tao et al. 2019); however, molecular analyses result in five different hypotheses of relationships among these four lineages (Saitoh et al. 2006; Šlechtová et al. 2007; W.-J. Chen et al. 2008; Mayden et al. 2008; Bohlen and Šlechtová 2009; Mayden and Chen 2010; Stout et al. 2016; Hirt et al. 2017; Tao et al. 2019).

  • Composition. Cypriniformes currently contains 4,827 living species (Fricke et al. 2023) classified in Catostomidae, Cobitoidei, Cyprinoidei, and Gyrinocheilus (Mayden and Chen 2010; Conway 2011; Tan and Armbruster 2018). †Jianghanichthys is the only fossil taxon of Cypriniformes that is not a lineage of Catostomidae, Cobitoidei, or Cyprinoidei (J. Liu et al. 2015). The age and location of †Jianghanichthys is presented in Appendix 1. Over the past 10 years 661 new living species of Cypriniformes have been described (Fricke et al. 2023), comprising 13.7% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Cypriniformes include (1) kinethmoid present (Fink and Fink 1981, 1996; Conway et al. 2010; Wiley and Johnson 2010; Conway 2011), (2) preethmoid present (Fink and Fink 1981, 1996; Conway et al. 2010; Wiley and Johnson 2010; Conway 2011), (3) dorsomedial autopalatine process present (Fink and Fink 1981, 1996; Conway et al. 2010; Wiley and Johnson 2010), (4) autopalatime-endopterygoid articulation present (Fink and Fink 1981, 1996; Conway et al. 2010; Wiley and Johnson 2010; Conway 2011), (5) loss of ectopterygoid-autopalatine overlap (Fink and Fink 1981, 1996; Conway et al. 2010; Wiley and Johnson 2010), (6) premaxilla extends furthest dorsally adjacent to midline (Fink and Fink 1981, 1996; Conway et al. 2010; Wiley and Johnson 2010; Conway 2011), (7) presence of ankylosed teeth on ceratobranchial 5 (Fink and Fink 1981, 1996; Conway et al. 2010; Wiley and Johnson 2010; Conway 2011), (8) lateral process of second vertebral centrum is elongate and projects into somatic musculature (Fink and Fink 1981, 1996; Conway et al. 2010; Wiley and Johnson 2010; Conway 2011), (9) absence of pharyngobranchial uncinate processes (Siebert 1987; Conway 2011), (10) three branchiostegal rays (Conway et al. 2010; Conway 2011; Mabee et al. 2011), and (11) teeth on ceratobranchial 5 arranged in a single row (Mabee et al. 2011).

  • Synonyms. Eventognathi (T. N. Gill 1861a:8–9; Gregory 1907:477–478; Jordan 1923:139–145), Cyprinidae (Boulenger 1904b:581–586; Goodrich 1909:375–376), Cyprinoidei (Berg 1940:444–446; Greenwood et al. 1966:384–386, 396; X.-W. Wu et al. 1981:572), Cyprinoidea (McAllister 1968:70–71), Cyprinoidae (Gosline 1971:121), and Cypriniphysae (Betancur-R et al. 2017) are ambiguous synonyms of Cypriniformes.

  • Comments. The taxa delimited here as Cypriniformes were grouped together in several pre-phylogenetic classifications (T. N. Gill 1893; Boulenger 1904b:581–586; Gregory 1907:477–478; Goodrich 1909:375–376; Regan 1911d; Jordan 1923:139–145; Berg 1940). The name Cypriniformes was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade. Despite the strong support for cypriniform monophyly, relationships among the constituent lineages are not well-resolved, and there are disparate hypotheses on the phylogenetic placement of Paedocypris (Britz and Conway 2011; Britz, Conway, et al. 2014; Tan and Armbruster 2018).

  • The earliest fossil Cypriniformes is †Jianghanichthys hubeiensis from the early Eocene (56.0–47.8 Ma) of China (J. Liu et al. 2015). Bayesian relaxed molecular clock analyses of Cypriniformes result in an average posterior crown age estimate of 97.2 million years ago, with the credible interval ranging between 84.9 and 115.3 million years ago (Hughes et al. 2018).

  • img-z57-1_03.gif

    FIGURE 9.

    Phylogenetic relationships of the major living lineages and fossil taxa of Cypriniformes, Cobitoidei, and Cyprinoidei. Filled circles identify the common ancestor of clades, with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z55-1_03.jpg

    Cobitoidei L. J. F. J. Fitzinger 1832:332
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Botia almorhae Gray 1831 and Cobitis taenia Linnaeus 1758. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek κωβπτις (cobitis), which is an adjective of the gudgeon Gobio gobio (Linnaeus 1758), translating to “like a gudgeon” (D. W. Thompson 1947:139; Kottelat 2012).

  • Registration number. 905.

  • Reference phylogeny. A phylogeny of 1,703 species of Cypriniformes inferred from a supermatrix of 27 nuclear and mitochondrial genes (Rabosky et al. 2018; J. Chang et al. 2019). The phylogeny is available on the Dryad data repository (Rabosky et al. 2019). Phylogenetic relationships of the major lineages of Cobitoidei are presented in Figure 9.

  • Phylogenetics. Analysis of morphological characters results in the resolution of a clade containing Catostomidae, Cobitoidei, and Gyrinocheilus (Siebert 1987; Conway and Mayden 2007; Conway 2011; Britz, Conway, et al. 2014), which had been called Cobitidoidea (Siebert 1987) and Cobitoidea (Sawada 1982; Conway et al. 2010; Simons and Gidmark 2010; J. S. Nelson et al. 2016:186). Morphological and molecular phylogenetic analyses consistently resolve Cobitoidei as monophyletic (Siebert 1987; Saitoh et al. 2006; Q. Tang et al. 2006; Šlechtová et al. 2007; Mayden et al. 2008, 2009; Bohlen and Šlechtová 2009; W.-J. Chen et al. 2009; Mayden and Chen 2010; Conway 2011; Britz, Conway, et al. 2014; Stout et al. 2016; Rabosky et al. 2018; Luo et al. 2023), but some molecular analyses resolve Cobitoidea as paraphyletic (W.-J. Chen et al. 2009; Stout et al. 2016).

  • Phylogenetic inferences of relationships within Cobitoidei are broadly congruent between morphological and molecular studies (e.g., Saitoh et al. 2006; Šlechtová et al. 2007; W.-J. Chen et al. 2009; Conway 2011; Mabee et al. 2011; Tao et al. 2019) with Botiidae (bottid loaches) placed as the sister lineage of all other Cobitoidei and resolution of a clade containing Cobitidae (loaches), Balitoridae (hillstream loaches), and Nemacheilidae (stone loaches) (Q. Tang et al. 2006; Mayden et al. 2008; Bohlen and Šlechtová 2009; Mayden and Chen 2010; S. Liu et al. 2012; Stout et al. 2016; Luo et al. 2023). Molecular phylogenies place Barbucca (fire-eyed loaches) and Serpenticobitis (serpent loaches) in Balitoridae (Šlechtová et al. 2007; Bohlen and Šlechtová 2009; Rabosky et al. 2018), Ellopostoma (enigmatic loaches) as the sister lineage of Nemacheilidae, Balitoridae, or a clade containing Balitoridae and Nemacheilidae (Bohlen and Šlechtová 2009; W.-J. Chen et al. 2009; Rabosky et al. 2018; Luo et al. 2023), and Vaillantella (longfin loaches) as the sister lineage of the cobitoid clade that contains Cobitidae, Ellopostoma, Balitoridae, and Nemacheilidae (Q. Tang et al. 2006; Šlechtová et al. 2007; Bohlen and Šlechtová 2009; W.-J. Chen et al. 2009; S. Liu et al. 2012; Stout et al. 2016; Rabosky et al. 2018).

  • Composition. There are currently 1,361 species of Cobitoidei (Fricke et al. 2023) classified in Balitoridae, Botiidae, Cobitidae, Ellopostoma, Nemacheilidae, and Vaillantella (Šlechtová et al. 2007; Bohlen and Šlechtová 2009; W.-J. Chen et al. 2009; Tan and Armbruster 2018). Over the past 10 years there have been 266 new living species of Cobitoidei described (Fricke et al. 2023), comprising 19.5% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological synapomorphies for Cobitoidei include (1) presence of transversus ventralis V process on ceratobranchial 5 (Siebert 1987; Conway 2011), (2) presence of second preethmoid (Conway 2011), (3) anteriormost edge of orbitosphenoid contacts ethmoid complex (Conway 2011), (4) presence of preautopalatine (Conway 2011), (5) presence of cleithral-occipital ligament (Conway 2011), and (6) third and fourth lateral line ossifications much larger than other lateral line ossifications (Conway 2011).

  • Synonyms. Cobitidoidea (Siebert 1987:43) and Cobitoidea (Sawada 1982:212) are approximate synonyms of Cobitoidei.

  • Comments. Cobitoidei was the name applied to the paraphyletic group that included all non-cyprinoid cypriniforms to the exclusion of Catostomidae (Kottelat 2012). Given the uncertainty in the phylogenetic relationships among major lineages of Cypriniformes (Figure 9), we apply the group name Cobitoidei to all non-cyprinoid cypriniforms to the exclusion of Gyrinocheilus and Catostomidae.

  • Cobitoidei are frequently classified with Catostomidae and Gyrinocheilus (Siebert 1987; Conway et al. 2010; J. S. Nelson et al. 2016:186) or with Gyrinocheilus to the exclusion of Catostomidae (Kottelat 2012). Some classifications of ray-finned fishes include Barbuccidae, Gastromyzontidae, and Serpenticobitidae as taxonomic families of Cobitoidei (Kottelat 2012; J. S. Nelson et al. 2016:191–193; Betancur-R et al. 2017; Tan and Armbruster 2018). Along with many other researchers, we include Barbucca, Gastromyzontinae, and Serpenticobitis in Balitoridae (Q. Tang et al. 2006; Šlechtová et al. 2007; Bohlen and Šlechtová 2009; W.-J. Chen et al. 2009; S. Liu et al. 2012; Z. S. Randall and Page 2015; Tao et al. 2019). This inclusive delimitation of Balitoridae is both consistent with phylogenetic relationships and reduces the number of redundant group names in the classification of Cobitoidei, as both Barbuccidae and Serpenticobitidae contain a single genus and ranking these clades as equivalent to Gastromyzontidae and Balitoridae conveys no information about their phylogenetic relationships.

  • The cobitoid fossil record is sparse and limited to Asia and Europe (Conway et al. 2010). The earliest fossil cobitoids are †Cobitis longipectoralis from the late early Miocene (18 Ma) and †C. nanningensis from the early-middle Oligocene in China (G.-J. Chen et al. 2010, 2015). Relaxed molecular clock analyses estimate the age of Cobitoidei to be between 50 and 78 million years ago (Hughes et al. 2018).

  • img-z58-6_03.gif

    Cyprinoidei L. J. F. J. Fitzinger 1832:332
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Cyprinus carpio Linnaeus 1758, Danio rerio (Hamilton 1822), Leuciscus leuciscus (Cuvier 1816), and Paedocypris progenetica Kottelat et al. 2006. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek κυπρῖνoς (kuːpɹ̍iːno͡Ʊz), frequently applied to the European Carp, Cyprinus carpio (D. W. Thompson 1947:135–136).

  • Registration number. 906.

  • Reference phylogeny. A phylogeny of 1,703 species of Cypriniformes inferred from a supermatrix of 27 nuclear and mitochondrial genes (Rabosky et al. 2018; J. Chang et al. 2019). The phylogeny is available on the Dryad data repository (Rabosky et al. 2019). Although the reference phylogeny does not include Paedocypris progenetica, the species resolves within the cyprinoid clade Danionidae (danionins) in phylogenetic analysis of mtDNA, combined mtDNA and nuclear gene sequences, and morphological characters (Rüber et al. 2007; Fang et al. 2009; K. L. Tang et al. 2010; Britz, Conway, et al. 2014). Phylogenetic analysis of DNA sequences of nuclear genes resolves Paedocypris as the sister lineage of Cyprinoidei or Cypriniformes (Mayden and Chen 2010; Stout et al. 2016; Malmstrøm et al. 2018). Phylogenetic relationships among the major clades of Cyprinoidei are presented in Figure 9.

  • Phylogenetics. Inference of the phylogenetics of Cyprinoidei is challenged by the high diversity of species in the clade; incongruent relationships of miniature species classified in Paedocypris, Sundadanio, and Fangfangia (Mayden and Chen 2010; Britz et al. 2011; Britz, Conway, et al. 2014); and phylogenetic resolution of the southeast Asian Tanichthys (cardinal minnows) and the European Tinca tinca (Tench) (Conway et al. 2010; Simons and Gidmark 2010; Tan and Armbruster 2018). Despite the remaining problems in the phylogeny of Cyprinoidei, incremental resolution of their relationships over the past 30 years has led to the elevation of 11 taxonomic families that were all classified as Cyprinidae (carps) for over 100 years (T. N. Gill 1872, 1893; Hensel 1970; W.-J. Chen and Mayden 2009; Tan and Armbruster 2018).

  • Phylogenetic analysis of Cyprinoidei using morphological characters resolves Cyprinidae as the sister lineage of all other cyprinoid lineages and Danionidae (dianions) as the sister lineage of a clade containing Acheilognathidae (bitterlings), Gobionidae (gudgeons), Leuciscidae (true minnows), and Xenocyprididae (Cavender and Coburn 1992; Conway 2011). In addition, Tinca tinca has uncertain resolution and Psilorhynchus (torrent minnows) is the sister lineage of all other Cyprinoidei (Cavender and Coburn 1992; Conway 2011). Analysis of a dataset that expands the character matrix from Conway (2011) resolves the cyprinoid miniature lineages Paedocypris and Sundadanio in a clade with Danionella (Britz, Conway, et al. 2014).

  • There are many molecular phylogenetic studies of Cyprinoidei that collectively include all known major lineages. The types of molecular data include whole mtDNA genomes (Saitoh et al. 2006; He, Gu, et al. 2008; Mayden et al. 2008; Chen et al. 2023; Hao et al. 2023), individual mtDNA or nuclear genes (e.g., Cunha et al. 2002; X. Z. Wang et al. 2007; He, Mayden, et al. 2008), combinations of mtDNA and nuclear genes (e.g., W.-J. Chen and Mayden 2009; Mayden and Chen 2010; Tao et al. 2019), and phylogenomic datasets (Stout et al. 2016; Hughes et al. 2018). Molecular phylogenies consistently nest Psilorhynchus within Cyprinoidei as the sister lineage of Cyprinidae (Šlechtová et al. 2007; He, Gu, et al. 2008; W.-J. Chen and Mayden 2009; Mayden and Chen 2010; K. L. Tang et al. 2013; Hirt et al. 2017; Rabosky et al. 2018; Tao et al. 2019), resolve a clade containing Acheilognathidae, Gobionidae, Leptobarbus, Leuciscidae, Sundadanio, Tanichthys, Tinca tinca, and Xenocyprididae (W.-J. Chen and Mayden 2009; Fang et al. 2009; Mayden and Chen 2010; K. L. Tang et al. 2013; Stout et al. 2016; Hirt et al. 2017; Rabosky et al. 2018), and a lineage that includes Acheilognathidae, Gobionidae, Leuciscidae, Tanichthys, Tinca tinca, and Xenocyprididae (Saitoh et al. 2006; Rüber et al. 2007; X. Z. Wang et al. 2007, 2012; W.-J. Chen et al. 2008; He, Mayden, et al. 2008; Mayden et al. 2008, 2009; W.-J. Chen and Mayden 2009; Fang et al. 2009; Mayden and Chen 2010; K. L. Tang et al. 2013; Tao et al. 2013, 2019; Stout et al. 2016; Hirt et al. 2017; Hughes et al. 2018; Rabosky et al. 2018; F. Chen et al. 2023; Hao et al. 2023). The relationships of Tinca tinca and Tanichthys vary among studies, with some resolving these two lineages as closely related (Fang et al. 2009; Mayden and Chen 2010; K. L. Tang et al. 2013) and sharing phylogenetic affinities with Leuciscidae (X. Z. Wang et al. 2007, 2012; Stout et al. 2016; Rabosky et al. 2018) or Xenocyprididae (Rüber et al. 2007; Tao et al. 2013; Hirt et al. 2017; Tao et al. 2019).

  • Phylogenetic analysis of mtDNA, combined mtDNA and nuclear gene sequences, and morphological characters resolve Paedocypris within the cyprinoid subclade Danionidae (Rüber et al. 2007; Fang et al. 2009; K. L. Tang et al. 2010, 2013; Britz, Conway, et al. 2014); however, analysis of nuclear gene datasets places this taxon as the sister lineage of Cyprinoidei or Cypriniformes (Mayden and Chen 2010; Stout et al. 2016; Malmstrøm et al. 2018). Gene trees inferred from each of the six loci examined by Mayden and Chen (2010) exhibit six different phylogenetic resolutions of Paedocypris: nested in Cyprinidae, the sister lineage of Cyprinidae, the sister lineage of Cypriniformes, the sister lineage of Catostomidae, the sister lineage of Gyrinocheilus, and nested within Danionidae (Britz, Conway, et al. 2014). It is possible that the disparate phylogenetic placements of Paedocypris among molecular datasets are the result of long branch attraction related to the dramatically reduced size of its genome (Britz, Conway, et al. 2014; Malmstrøm et al. 2018).

  • Composition. There are currently more than 3,375 living species of Cyprinoidei (Fricke et al. 2023) that includes Tinca tinca and species in Acheilognathidae, Cyprinidae, Danionidae, Gobionidae, Leptobarbus (sultan barbs), Leuciscidae, Paedocypris, Psilorhynchus, Sundadanio, Tanichthys, Tinca, and Xenocyprididae. Over the past 10 years 395 new living species of Cyprinoidei have been described (Fricke et al. 2023), comprising 11.7% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Cyprinoidei include (1) absence of uncinate process on epibranchials 1 and 2 (Siebert 1987; Cavender and Coburn 1992; Conway et al. 2010), (2) pharyngobranchial 1 absent (Siebert 1987; Cavender and Coburn 1992; Conway et al. 2010), (3) pharyngobranchial 3 overlaps pharyngobranchial 2 (Siebert 1987; Cavender and Coburn 1992; Conway et al. 2010), (4) presence of well-developed subtemporal fossae (Siebert 1987; Cavender and Coburn 1992; Conway et al. 2010), (5) anterior opening of trigeminal-facial chamber positioned between prootic and pterosphenoid, (6) loss of contact between infraorbital 5 and supraorbital (Cavender and Coburn 1992; Conway et al. 2010), and (7) presence of opercular canal (Cavender and Coburn 1992; Conway et al. 2010).

  • Synonyms. Cyprinidae (T. N. Gill 1872:18, 1893:132; Siebert 1987:43; Howes 1991b:8–17; Cavender and Coburn 1992:296–300; J. S. Nelson 2006:139–143; Simons and Gidmark 2010:419–425) and Cyprinoidea (Greenwood et al. 1966:396; Conway 2011, fig. 43; J. S. Nelson et al. 2016:181) are ambiguous synonyms of Cyprinoidei.

  • Comments. The lineages of Cyprinoidei were long classified as Cyprinidae and highlighted as one of the most species-rich taxonomic families of vertebrates (T. N. Gill 1872; Jordan 1923; Greenwood et al. 1966; Howes 1991b; J. S. Nelson et al. 2016:181); however, the disassembly of Cyprinidae sensu lato into 12 taxonomic families was the result of a desire to preserve the Linnaean taxonomic family rank of the monogeneric Psilorhynchidae (Hora 1925) that is phylogenetically nested within Cyprinoidei (W.-J. Chen and Mayden 2009; Tan and Armbruster 2018). The name Cyprinoidei was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade. While it is appropriate to view the changing classification of Cyprinoidei as an outcome of greater resolution of their phylogenetic relationships, the uncertainty as to the phylogeny of the miniature lineages Paedocypris, Sundadanio, and Fangfangia (Britz, Conway, et al. 2014) remains one of the most important issues in vertebrate phylogeny.

  • The earliest fossil taxon of Cyprinoidei is †Palaeogobio zhongyuanensis, classified in Gobionidae from the early middle Eocene (approximately 47 Ma) of China (Zhou 1990; M.-M. Chang and Chen 2008). An interesting set of cyprinoid fossils from the Sangkarewang Formation in Sumatra, Indonesia is classified in Cyprinidae and Danionidae; however, the age of the formation is only tentatively assigned to the middle Eocene (Murray 2019, 2020). Bayesian relaxed molecular clock analyses of the crown age of Cyprinoidei result in a credible interval ranging between 67 and 98 million years ago (Hirt et al. 2017).

  • img-z60-7_03.gif

    Gymnotiformes C. T. Regan 1911d:23
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Gymnotus carapo Linnaeus 1758, Gymnotus pantherinus (Steindachner 1908), Apteronotus albifrons (Linnaeus 1766), and Sternopygus macrurus (Bloch and Schneider 1801). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek γυµνός (d͡Ʒ̍Imno͡Ʊz), meaning naked, and νῶτoν (n̍a͡Ʊtәn), meaning back. The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 907.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 966 ultraconserved element (UCE) loci (Alda et al. 2019). Although Gymnotus carapo is not included in the reference phylogeny, it resolves in a clade with five other species of Gymnotus, including G. pantherinus, in a phylogenetic analysis of Sanger-sequenced mitochondrial and nuclear genes (Tagliacollo et al. 2016, figs. 2–4). Phylogenetic relationships among the major lineages of Gymnotiformes are presented in Figure 8.

  • Phylogenetics. There is substantial morphological evidence supporting the monophyly of Gymnotiformes, which is consistently corroborated in molecular phylogenetic analyses (e.g., Fink and Fink 1981, 1996; Nakatani et al. 2011; W.-J. Chen et al. 2013; Arcila et al. 2017). Phylogenetic relationships among the five major lineages of Gymnotiformes differ among analyses of morphological characters (Triques 1993; Gayet et al. 1994; Albert 2001), short Sanger-sequenced fragments of mtDNA genes (Alves-Gomes et al. 1995), combined analyses of morphology and DNA sequences of mtDNA and nuclear genes (Albert and Crampton 2005; Tagliacollo et al. 2016), and phylogenomic datasets (Arcila et al. 2017; Alda et al. 2019). The phylogenies differ as to the resolution of the sister lineage of all other Gymnotiformes: analyses of morphology and combined molecular and morphological datasets place Gymnotidae (nakedback knifefishes) as the sister lineage of all other Gymnotiformes (Albert 2001; Albert and Crampton 2005; Tagliacollo et al. 2016) and phylogenies inferred from alternative morphological datasets and phylogenomic datasets composed of exons and UCEs resolve Apteronotidae (ghost knifefishes) as the sister to all other Gymnotiformes (Triques 1993; Gayet et al. 1994; Arcila et al. 2017; Alda et al. 2019). Coalescent-based species tree analysis of UCE loci results in a phylogenetic tree in which lineages that produce a pulse-type electrical signal [Gymnotidae, Hypopomidae (bluntnose knifefishes), and Rhamphichthyidae (sand knifefishes)] are a monophyletic group that is the sister group of a clade comprising lineages that produce a wave-type electrical signal [Apteronotidae and Sternopygidae (glass knifefishes)] (Alda et al. 2019).

  • Morphological studies imply that Gymnotiformes and Siluriformes share a common ancestor relative to other major clades of Otophysi (Fink and Fink 1981, 1996); however, no unconstrained phylogenetic analysis of molecular data supports this relationship (Dimmick and Larson 1996; Ortí and Meyer 1996; Nakatani et al. 2011; Betancur-R, Broughton, et al. 2013; W.-J. Chen et al. 2013; Arcila et al. 2017; Chakrabarty et al. 2017; Hughes et al. 2018; Melo, Sidlauskas, et al. 2022). The phylogeny of Otophysi inferred from molecular data suggests that the passive electroreception and its associated specialized neural anatomy, cytology, and physiology in Gymnotiformes and Siluriformes has multiple evolutionary origins or multiple losses within the clade (Fink and Fink 1996; Albert et al. 1998).

  • Composition. There are currently 272 living species of Gymnotiformes (Fricke et al. 2023) classified in Apteronotidae, Gymnotidae, Hypopomidae, Rhamphichthyidae, and Sternopygidae (Ferraris et al. 2017). Over the past 10 years 68 new living species of Gymnotiformes have been described (Fricke et al. 2023), comprising 25% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Gymnotiformes include (1) absence of palatine ossification and palatine cartilage with flexure permitting mobility (Fink and Fink 1981; Albert 2001), (2) mesopterygoid with vertical strut that usually articulates with orbitosphenoid (Fink and Fink 1981), (3) claustrum of Weberian apparatus absent as a separate ossified element (Fink and Fink 1981; Albert 2001), (4) anterior and posterior parts of Baudelot's ligament attach to cleithrum (Fink and Fink 1981), (5) pelvic girdle and pelvic fin absent (Fink and Fink 1981; Albert 2001), (6) dorsal fin absent (Fink and Fink 1981; Albert 2001), (7) presence of elongate anal fin (Fink and Fink 1981; Albert 2001), (8) anal fin rays articulate directly with proximal radials and distal radials are reduced (Fink and Fink 1981; Albert 2001), (9) caudal skeleton reduced to single element and caudal fin reduced or absent (Mago-Leccia and Zaret 1978; Fink and Fink 1981; Albert 2001), (10) anus placed ventral or anterior to pectoral fin origin (Fink and Fink 1981; Albert 2001), (11) absence of maxillary teeth (Albert 2001), (12) articular surface of maxilla on stalk (Albert 2001), (13) levator posterior muscle not differentiated, (14) lateral margins of parasphenoid not extending to a horizontal with trigeminal foramen (Albert 2001), (15) dorsal telencephalic area with large dorsalis centralis and small dorsalis medialis (Albert 2001), (16) eye in adults covered by epidermis (Albert 2001), (17) Schreckstoff club cells and fright response absent (Albert 2001), (18) ampullary organs organized into rosettes (Albert 2001), (19) ectopterygoid absent (Albert 2001), (20) metapterygoid triangular in shape (Albert 2001), (21) sixth epibranchial with elongate ascending process, and (22) presence of electric organs composed of rows of modified elongate myofibrils (Albert 2001).

  • Synonyms. Gymnonoti (T. N. Gill 1872:18; Jordan 1923:138), Gymnotidae (Boulenger 1904b:579–581), Gymnotoidei (Goodrich 1909:376–377; Berg 1940:443–444; Greenwood et al. 1966:383–384; Fink and Fink 1981:303), Gymnotoidea (McAllister 1968:69; Rosen and Greenwood 1970:23), and Gymnotoidae (Gosline 1971:121) are ambiguous synonyms of Gymnotiformes.

  • Comments. The group name Gymnotiformes has long been applied to the clade as defined above (Regan 1911d; Mago-Leccia 1978; J. S. Nelson 1984:154–156; Fink and Fink 1996; Betancur-R et al. 2017) and was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The fossil record of Gymnotiformes is limited to a handful of fragmentary fossils, including †Humboldtichthys kirschbaumi from the Miocene of Bolivia (Gayet et al. 1994; Gayet and Meunier 2000; Albert and Fink 2007). A morphological phylogenetic analysis places the holotype specimen of †H. kirschbaumi within Sternopygidae (Albert and Fink 2007). Bayesian relaxed molecular clock analyses of Gymnotiformes result in an average posterior crown age estimate of 62.4 million years ago, with the credible interval ranging between 46.0 and 81.6 million years ago (Hughes et al. 2018).

  • img-z62-5_03.gif

    Cithariniformes J. M. Mirande 2017:342
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Citharinus citharus (Geoffroy St. Hilaire 1809) and Distichodus mossambicusPeters1852.Thisisaminimum-crown-clade definition.

  • Etymology. From the ancient Greek κυθνρα (kIθ̍α͡ә), meaning harp or lute. The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 908.

  • Reference phylogeny. A phylogeny inferred from a maximum likelihood analysis of DNA sequences from two mtDNA genes and two nuclear genes (Arroyave et al. 2013, fig. 4). Phylogenetic relationships of Cithariniformes are shown in Figure 8.

  • Phylogenetics. Analyses of morphological and molecular characters consistently support the monophyly of Cithariniformes (Vari 1979; Ortí and Meyer 1997; Buckup 1998; Calcagnotto et al. 2005; Arroyave and Stiassny 2011; Arroyave et al. 2013; Arcila et al. 2017, 2018; Lavoué et al. 2017; Rabosky et al. 2018; Betancur-R. et al. 2019; Burns and Sidlauskas 2019; Melo, Sidlauskas, et al. 2022).

  • Composition. There are currently 117 species of Cithariniformes (Fricke et al. 2023) classified in Citharinidae (citharinids) and Distichodontidae (distichodontids). Over the past 10 years nine new living species of Cithariniformes have been described (Fricke et al. 2023), comprising 7.7% of the living species diversity in the clade.

  • Diagnostic apomorphies. Vari (1979) listed 14 morphological synapomorphies that support monophyly of Cithariniformes; however, eight of these character states are either ancestral within Otophysi or are secondarily derived in some lineages of Characiformes (Fink and Fink 1981). Morphological character states consistent with the monophyly of Cithariniformes include (1) second and third vertebrae with ventral elaborations and ventral expansion of os suspensorium (Vari 1979), (2) bicuspidate teeth (Vari 1979), (3) postcleithra 2 and 3 fused (Vari 1979), (4) hypurals 1 and 2 fused (Vari 1979), (5) absences of lateral wings on supraethmoid (Vari 1979), and (6) large and ventrally ovate third posttemporal fossa bordered by epioccipital and exoccipital (Vari 1979).

  • Synonyms. Citharinidae (Regan 1911d:21–22; J. S. Nelson 1994:142–143) and Citharinoidei (Buckup 1993:138; J. S. Nelson et al. 2016:194–195; Betancur-R et al. 2017:17) are ambiguous synonyms of Cithariniformes.

  • Comments. Cithariniformes was a group name applied to the clade containing Citharinidae and Distichodontidae (Mirande 2017, table 3), but long classified as a lineage of Characiformes (Vari 1979; Fink and Fink 1981, 1996; Buckup 1998; Betancur-R et al. 2017). Molecular phylogenetic analyses consistently resolve Characiformes, Siluriformes, and Gymnotiformes as a clade to the exclusion of Cithariniformes (Nakatani et al. 2011; Betancur-R, Broughton, et al. 2013; W.-J. Chen et al. 2013; Chakrabarty et al. 2017; Hughes et al. 2018, fig. S2; Faircloth et al. 2020; Melo, Sidlauskas, et al. 2022; Yang et al. 2023). It is unknown whether the seven morphological apomorphies identified by Fink and Fink (1981) supporting the hypothesis that Cithariniformes and Characiformes share common ancestry are present in a wider range of characiform taxa; their study did not include species of Acestrorhynchus (needlejaws), Gasteropelecidae (freshwater hatchetfishes), Iguanodectidae (tetras), Serrasalmidae (pacus), or Triportheidae (elongate hatchetfishes). The monophyly of both Cithariniformes and Characiformes is validated in phylogenetic analyses of morphological data matrices that use an explicit optimality criterion (Buckup 1998; de Pinna et al. 2018). However, the relationships of these two lineages relative to Siluriformes and Gymnotiformes have not been investigated using morphological phylogenetic analyses that seek a tree or set of trees with an optimal distribution of character state changes.

  • The earliest Cithariniformes fossil is a tooth identified as a species of Distichodus from the Lower Nawata formation at Lothagam, Kenya dated to approximately 7.5 million years ago (McDougall and Feibel 1999; K. M. Stewart 2001, 2003). Bayesian relaxed molecular clock analyses of Cithariniformes result in an average posterior crown age estimate of 119.7 million years ago, with the credible interval ranging between 93.2 and 149.3 million years ago (Melo, Sidlauskas, et al. 2022).

  • img-z63-6_03.gif

    Siluriformes O. P. Hay 1929:25
    [Lundberg 2020]

  • Definition. Defined as a minimum-crown-clade in Lundberg (2020d) as: “The crown clade originating in the most recent common ancestor of Loricaria cataphracta Linnaeus 1758 (Loricarioidei), Diplomystes (originally Silurus) chilensis (Molina 1782) (Diplomystidae), and Silurus glanis Linnaeus 1758 (Siluroidei).”

  • Etymology. From the ancient Greek σίλoυρoς (sIl̍Ʊ͡ o͡Ʊz), which is the name applied to several species of catfishes in Europe and Egypt, including the Wels Catfish, Silurus glanis (D. W. Thompson 1947:233–237). The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 199.

  • Reference phylogeny. Sullivan et al. (2006, figs. 1, 2) was designated as the primary reference phylogeny by Lundberg (2020d). Phylogeny of the living and fossil lineages of Siluriformes is presented in Figure 10. The placements of the fossil lineages †Bachmannia and †Hypsidoris are on the basis of inferences from morphology (L. Grande 1987; L. Grande and de Pinna 1998; Azpelicueta and Cione 2011).

  • Phylogenetics. There are several reviews on the phylogenetics of Siluriformes prior to the application of molecular data (de Pinna 1998; Diogo 2003, 2004; Teugels 2003). The monophyly of Siluriformes is supported in analyses of morphological characters (Fink and Fink 1981, 1996; Mo 1991; Arratia 1992; de Pinna 1993; Diogo 2004) and in all molecular phylogenetic studies, which includes analyses of whole mitochondrial genomes (Jondeung et al. 2007; Poulsen et al. 2009; Nakatani et al. 2011; Schedel et al. 2022; Duong et al. 2023), collections of Sanger-sequenced mitochondrial or nuclear genes (Betancur-R, Broughton, et al. 2013; W.-J. Chen et al. 2013), and phylogenomic datasets (Arcila et al. 2017; Chakrabarty et al. 2017; Hughes et al. 2018).

  • The first explicit phylogenetic studies of relationships within Siluriformes were morphological analyses aimed at determining the relationships among the Loricarioidei (Howes 1983; Schaefer 1990) and the relationships of the Eocene fossil taxon †Hypsidoris (L. Grande 1987; L. Grande and de Pinna 1998). More inclusive morphological phylogenetic studies aimed at including representatives of all the taxonomic families of Siluriformes place Diplomystidae (velvet catfishes) as the sister lineage of all other catfishes (Mo 1991; de Pinna 1993, 1998); support the monophyly of Loricarioidei; and do not resolve Siluroidei as monophyletic (Mo 1991; de Pinna 1993, 1998). A subsequent morphological analysis roots the phylogeny on Diplomystidae and resolves both Loricarioidei and Siluroidei as monophyletic (Diogo 2004, fig. 3.124); however, this study was critiqued on issues involving character state polarity and homology (Schaefer 2006). The Eocene fossil taxa †Bachmannia and †Hypsidoris are resolved as the sister lineages of Diplomystidae and Siluroidei, respectively (L. Grande 1987; L. Grande and de Pinna 1998; Azpelicueta and Cione 2011).

  • Molecular phylogenetic analyses of Siluriformes consistently resolve Loricarioidei as the sister lineage of a clade containing Diplomystidae and Siluroidei (Sullivan et al. 2006; Lundberg et al. 2007; Nakatani et al. 2011; W.-J. Chen et al. 2013; Kappas et al. 2016; Arcila et al. 2017; Rivera-Rivera and Montoya-Burgos 2017, 2018; Schedel et al. 2022). In molecular phylogenetic studies, the monophyly of Loricarioidei and Siluroidei are strongly supported (e.g., Sullivan et al. 2006; Nakatani et al. 2011; Arcila et al. 2017; Schedel et al. 2022). Molecular evolutionary rate heterogeneity among lineages is proposed as a mechanism for the incongruence between morphological and molecular phylogenies with regard to the placement of Loricarioidei in contrast to Diplomystidae as the sister lineage of all other Siluriformes (Rivera-Rivera and Montoya-Burgos 2018).

  • Composition. Siluriformes currently contains 4,188 species (Fricke et al. 2023) classified in Loricarioidei, Diplomystidae, and Siluroidei. Siluriformes includes the pan-diplomystid †Bachmannia and the pan-siluroid †Hypsidoris (L. Grande 1987; L. Grande and de Pinna 1998; Gayet and Meunier 2003; Azpelicueta and Cione 2011). Details of the ages and locations of the siluriform fossil taxa are presented in Appendix 1. Over the past 10 years 628 new living species of Siluriformes have been described (Fricke et al. 2023), comprising 15% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Siluriformes include (1) parietal bones absent (Fink and Fink 1981, 1996; Arratia 1992, 2003a; Wiley and Johnson 2010; Lundberg 2020d), (2) autopalatine bone separate from suspensorium (Fink and Fink 1981, 1996; Wiley and Johnson 2010; Lundberg 2020d), (3) ectopterygoid and endopterygoid reduced and not articulating with metapterygoid, quadrate, and hyomandibular (Fink and Fink 1981, 1996; Arratia 1992; Wiley and Johnson 2010; Lundberg 2020d), (4) metapterygoid anterodorsal to quadrate (Fink and Fink 1981, 1996; Arratia 1992; Wiley and Johnson 2010; Lundberg 2020d), (5) symplectic and posterior process of quadrate absent (Fink and Fink 1981, 1996; Arratia 2003a; Wiley and Johnson 2010; Lundberg 2020d), (6) preopercle and interopercle shortened (Fink and Fink 1981, 1996; Wiley and Johnson 2010; Lundberg 2020d), (7) subopercles absent (Fink and Fink 1981, 1996; Arratia 1992, 2003a; Wiley and Johnson 2010; Lundberg 2020d), (8) complex centrum formed by fusion of centra 2, 3, and 4 (Fink and Fink 1981, 1996; Arratia 1992; Wiley and Johnson 2010; Lundberg 2020d), (9) third and fourth neural arches fused to each other and to complex centrum (Fink and Fink 1981, 1996; Arratia 1992; Wiley and Johnson 2010; Lundberg 2020d), (10) parapophysis of second vertebral centrum absent (Fink and Fink 1981, 1996; Arratia 1992; Wiley and Johnson 2010; Lundberg 2020d), (11) transformator process of Weberian apparatus tripus separated posteriorly by width of the complex centrum (Fink and Fink 1981, 1996; Arratia 1992; Wiley and Johnson 2010), (12) parapophysis of fourth vertebral centrum expanded and articulating with posttemporal supracleithrum (Fink and Fink 1981, 1996; Arratia 1992; Wiley and Johnson 2010; Lundberg 2020d), (13) parapophysis of fourth vertebral centrum fused to complex centrum (Fink and Fink 1981, 1996; Arratia 1992; Wiley and Johnson 2010), (14) os suspensorium of Weberian apparatus consisting of only an anterior horizontal process (Fink and Fink 1981, 1996; Arratia 1992; Wiley and Johnson 2010), (15) suspensorium of pectoral girdle consisting of single ossified element including the supracleithrum, an ossified Baudelot's ligament, and posttemporal (Fink and Fink 1981, 1996; Arratia 1992; Wiley and Johnson 2010; Lundberg 2020d), (16) Baudelot's ligament distally bifurcated (Fink and Fink 1981, 1996), (17) dorsal fin with two tightly bound anterior spines (Lundberg et al. 2007; Lundberg 2020d), (18) skin naked with no bony-ridge scales that are present in most other lineages of Teleostei (Fink and Fink 1981, 1996; Lundberg 2020d), (19) palatoquadrate with separate pars autopalatine (Arratia 1992), (20) posterior palatoquadrate fused with symplectic cartilage (Arratia 1992), (21) articulation of autopalatine and vomer at midpoint of autopalatine (Arratia 1992), (22) articulation of autopalatine and lateral ethmoid at mid length of autopalatine (Arratia 1992), (23) entopterygoid not the main support for the eye (Arratia 1992), (24) retroarticular and anguloarticular fused (Arratia 1992), (25) Meckel cartilage with coronoid process (Arratia 1992), (26) upper pharyngeal tooth plate with retractor muscles (de Pinna 1993), (27) first pharyngobranchial lies parallel to first epibranchial (de Pinna 1993), (28) second pharyngobranchial elongated and rod-like (de Pinna 1993), (29) first basibranchial absent (de Pinna 1993), (30) intermuscular epineural and epipleural bones absent (Arratia 2003b; Lundberg 2020d), (31) maxillary bears a fleshy barbel (Lundberg 2020d), (32) basihyal absent (Lundberg 2020d), (33) postcleithra absent (Lundberg 2020d), and (34) pectoral fin with single spine with rotating and locking joint that articulates with cleithrum (Lundberg 2020d).

  • Synonyms. Siluridae (Swainson 1838:325–360; Günther 1864a:1–2), Siluri (Bleeker 1858:13–43), Nematognathi (T. N. Gill 1861a:11; Eigenmann and Eigenmann 1890:5; Jordan 1923:145–153), Siluroidei (Goodrich 1909:377–384; Bertin and Arambourg 1958:2302–2304; McAllister 1968:71–78), and Siluroidea (Regan 1911e) are ambiguous synonyms of Siluriformes.

  • Comments. The lineages delimited here as Siluriformes were grouped together in several pre-Darwinian and pre-cladistic classifications of teleosts (Bleeker 1858; Günther 1864a; Boulenger 1904b), with a degree of sophistication exemplified by placing Diplomystidae apart from all other groups of Siluriformes based on the presence of a toothed maxillary (Goodrich 1909:380; Regan 1911e). The monophyly of Siluriformes is consistently supported, from the first phylogenetic treatments of fishes to recent molecular analyses (Greenwood et al. 1966; Fink and Fink 1981; Nakatani et al. 2011; Arcila et al. 2017; Hughes et al. 2018). Remaining problems in the phylogenetics of Siluriformes include the incongruence among morphological and molecular studies regarding Diplomystidae or Loricarioidei as the sister lineage of all other Siluriformes (Mo 1991; de Pinna 1993; Arcila et al. 2017; Rivera-Rivera and Montoya-Burgos 2018) and the slight morphological support for the monophyly of Siluroidei (Diogo 2004; Lundberg et al. 2014).

  • The earliest fossils of Siluriformes that are not Loricarioidei or Siluroidei are Campanian (83.2–72.2 Ma) and Maastrichtian (72.2–66.0 Ma) pectoral spines identified as Diplomystidae from Argentina and Bolivia (Cione 1987; Arratia and Cione 1996; Gayet and Meunier 1998). Bayesian relaxed molecular clock analyses of Siluriformes result in an average posterior crown age estimate of 121.4 million years ago, with the credible interval ranging between 111.3 and 131.7 million years ago (Hughes et al. 2018).

  • img-z67-1_03.gif

    FIGURE 10.

    Phylogenetic relationships of the major living lineages and fossil taxa of Siluriformes, Loricarioidei, and Siluroidei. Filled circles identify the common ancestor of clades, with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z64-1_03.jpg

    Loricarioidei P. Bleeker 1858:37
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Nematogenys inermis (Guichenot 1848), Loricaria cataphracta Linnaeus 1758, Loricaria simillima Regan 1904a, and Trichomycterus guianense (Eigenmann 1909). This is a minimum-crown-clade definition.

  • Etymology. From the Latin lorica, a coat of chain mail armor, in reference to the bony plates on the body of many species in this clade.

  • Registration number. 909.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of two nuclear genes (Sullivan et al. 2006, fig. 1). Although Loricaria cataphracta is not included in the reference phylogeny, it resolves with other species of Loricaria in molecular phylogenetic analyses (Covain et al. 2016, fig. 7; Moreira et al. 2017, fig. 3). Phylogenetic relationships among the major lineages of Loricarioidei are presented in Figure 10.

  • Phylogenetics. Morphological and molecular phylogenetic analyses consistently support the monophyly of Loricarioidei (Howes 1983; Schaefer 1990; Mo 1991; de Pinna 1993, 1998; Diogo 2004; Sullivan et al. 2006; Lundberg et al. 2007; Covain et al. 2016; Arcila et al. 2017; Moreira et al. 2017; Schedel et al. 2022). Within Loricarioidei, morphological and molecular analyses resolve two primary clades: Trichomycteridae (pencil catfishes) and Nematogenys inermis (Mountain Catfish) form a monophyletic group that is the sister lineage of a clade containing Callichthyidae (callichthyid armored catfishes), Astroblepus (climbing catfishes), and Loricariidae (sucker-mouth armored catfishes) (Mo 1991; de Pinna 1993, 1998; Arcila et al. 2017).

  • Composition. Loricarioidei currently contains 1,773 species (Ferraris 2007; Fricke et al. 2023) that includes Nematogenys inermis and species classified in Astroblepus, Callichthyidae, Loricariidae, Nematogenys, Scoloplax (spiny dwarf catfishes), and Trichomycteridae (Sullivan et al. 2006). There have been 391 new living species of Loricarioidei described over the past 10 years (Fricke et al. 2023), comprising 22.1% of the living species diversity of the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Loricarioidei include (1) odontodes present (Baskin 1973; Howes 1983; Schaefer and Lauder 1986; Schaefer 1990; de Pinna 1998), (2) encapsulated swim bladder present (Howes 1983; Schaefer and Lauder 1986; Schaefer 1990), (3) median processes of exoccipitals do not meet at midline (de Pinna 1998), (4) absence of anterior cartilages on arms of basipeterygium (de Pinna 1998), (5) bifid cusps on oral jaw teeth (de Pinna 1998), and (6) autopalatine compressed dorsoventrally with dorsal process that forms surface of articulation with neurocranium (Diogo 2004).

  • Synonyms. Loricarioidea (Schaefer and Lauder 1986, fig. 1; de Pinna 1998:292–294, fig. 6) is an ambiguous synonym of Loricarioidei.

  • Comments. On the basis of the results of morphological and molecular phylogenetic studies (e.g., Schaefer 1990; Sullivan et al. 2006), the group name Loricarioidei was applied to the clade containing Loricariidae, Astroblepus, Scoloplax, Callichthyidae, Trichomycteridae, and Nematogenys (Sullivan et al. 2006).

  • The earliest phylogenetic analyses within Siluriformes aimed to resolve relationships among lineages of Loricarioidei (Howes 1983; Schaefer 1990). The results of morphological and molecular phylogenetic analyses of relationships within Loricarioidei are broadly congruent (e.g., de Pinna 1993; Arcila et al. 2017).

  • The earliest fossil Loricarioidei is the callichthyid †Corydoras revelatus from the Late Paleocene (58.5 Ma) of Argentina (Marshall et al. 1997; Lundberg et al. 1998; Reis 1998). Relaxed molecular clock analyses estimate the crown age of Loricarioidei at approximately 90 million years ago (Rabosky et al. 2018).

  • img-z68-1_03.gif

    Siluroidei P. Bleeker 1858:34
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Silurus glanis Linnaeus 1758, Cetopsis coecutiens (Lichtenstein 1819), and Pimelodus maculatus Lacepède 1803, but not Loricaria simillima Regan 1904a or Diplomystes nahuelbutaensis Arratia 1987b. This is a minimum-crown-clade definition with external specifiers.

  • Etymology. From the ancient Greek σίλoυρoς (sIl̍Ʊɹo͡Ʊz), which is the name applied to several species of catfishes in Europe and Egypt, including the Wels Catfish, Silurus glanis (D. W. Thompson 1947:233–237).

  • Registration number. 910.

  • Reference phylogeny. A phylogeny of 752 species of Siluroidei inferred from a supermatrix of 27 nuclear and mitochondrial genes (Rabosky et al. 2018; J. Chang et al. 2019). The phylogeny is available on the Dryad data repository (Rabosky et al. 2019). Phylogenetic relationships among the major lineages of Siluroidei are presented in Figure 10.

  • Phylogenetics. The first phylogenetic studies to resolve Siluroidei as monophyletic include analyses of 440 morphological characters and DNA sequences of two nuclear genes (Diogo 2004, fig. 3.124; Sullivan et al. 2006). Molecular phylogenetic analyses consistently support the monophyly of Siluroidei; however, relationships among deeper nodes are typically unresolved and poorly supported (Sullivan et al. 2006; Lundberg et al. 2007; Kappas et al. 2016; Arcila et al. 2017; Rivera-Rivera and Montoya-Burgos 2018; K. Zhang, Liu, et al. 2021; Schedel et al. 2022; Duong et al. 2023). Despite the lack of strong resolution along the backbone of the siluroid phylogeny, there are several well-supported conclusions from morphological and molecular phylogenetic analyses. For example, Cetopsidae (whale catfishes) are either deeply branching or specifically placed as the sister lineage of all other Siluroidei in many phylogenetic studies (Mo 1991; de Pinna 1993; Diogo 2004; Lundberg et al. 2007; Q. Wang et al. 2015; Arcila et al. 2017; Rivera-Rivera and Montoya-Burgos 2018; K. Zhang, Liu, et al. 2021; Schedel et al. 2022; Duong et al. 2023). Phylogenies inferred from both morphological and molecular datasets result in the polyphyly of the traditional delimitation of Schilbeidae (butter catfishes) distributed in freshwater habitats of Africa and Asia (Mo 1991; de Pinna 1993; Hardman 2005; Sullivan et al. 2006; Schedel et al. 2022), with Asian lineages subsequently classified in Ailiidae (Asian schilbeids) (J. Wang et al. 2016; X. Li and Zhou 2018). Phylogenetic analyses consistently support the common ancestry of several groupings of siluroid lineages: Aspredinidae (banjo catfishes), Auchenipteridae (driftwood catfishes), and Doradidae (thorny catfishes) (Sullivan et al. 2006, 2008; Nakatani et al. 2011; Q. Wang et al. 2015; Arcila et al. 2017; Rabosky et al. 2018; Rivera-Rivera and Montoya-Burgos 2018; Cui et al. 2020; K. Zhang, Liu,et al. 2021; Schedel et al. 2022; Duong et al. 2023); Clariidae (airbreathing catfishes) and Heteropneustes (airsac catfishes) (Mo 1991; Diogo 2004; Hardman 2005; Sullivan et al. 2006; Nakatani et al. 2011; Q. Wang et al. 2015; J. Wang et al. 2016; Rabosky et al. 2018; Cui et al. 2020; K. Zhang, Liu, et al. 2021; Schedel et al. 2022; Duong et al. 2023); the Madagascar endemic Anchariidae (Malagasy catfishes) and the marine Ariidae (sea catfishes) (de Pinna 1993; Sullivan et al. 2006; J. Wang et al. 2016); and the east Asian Cranoglanis (armorhead catfishes) and the North American Ictaluridae (bullhead catfishes) (Diogo 2004; Hardman 2005; Sullivan et al. 2006; Nakatani et al. 2011; Q. Wang et al. 2015; Kappas et al. 2016; J. Wang et al. 2016; Rabosky et al. 2018; Cui et al. 2020; Schedel et al. 2022; Duong et al. 2023).

  • An important result from the earliest inclusive molecular phylogenetic analyses of Siluroidei was the resolution of two inclusive clades: Big Asia [Ailiidae, Akysidae (stream catfishes), Amblycipitidae (torrent catfishes), Bagridae (bagrid catfishes), Horabagridae (sun catfishes), and Sisoridae (sisorid catfishes)] and Big Africa [Amphiliidae (loach catfishes), Claroteidae (claroteids), Lacantunia enigmatica (Chiapas Catfish), Malapteruridae (electric catfishes), Mochokidae (squeakers), and Schilbeidae], which highlighted freshwater habitats in Asia and Africa as important areas of siluroid diversification (Sullivan et al. 2006; Lundberg et al. 2007). Molecular phylogenetic analyses resolve the enigmatic Conorhynchos conirostris (Anteater Catfish), which is currently not classified in a Linnean-ranked taxonomic family (Eschmeyer and Fricke 2023), in a clade with other South American freshwater lineages that includes Heptapteridae (threebarbeled catfishes), Pimelodidae (long-whiskered catfishes), and Pseudopimelodidae (bumblebee catfishes) (Sullivan et al. 2006, 2013; G. S. C. Silva et al. 2021). An analysis of a supermatrix of 27 nuclear and mitochondrial genes places this South American siluroid lineage in the Big Africa clade (Rabosky et al. 2018; J. Chang et al. 2019).

  • Morphological and molecular studies provide insight into the phylogenetic relationships of the enigmatic South African lineage Austroglanis (rock catfishes) and two species discovered and described in the early 21st century, Lacantunia enigmatica and Kryptoglanis shajii (Subterranean Catfish) (Skelton et al. 1984; Rodiles-Hernández et al. 2005; Vincent and Thomas 2011; Britz, Kakkassery, et al. 2014). Austroglanis was initially classified in Bagridae (Skelton et al. 1984), but morphological and molecular analyses place this lineage in a clade that contains Cranoglanis, Ictaluridae, and Anchariidae or as the sister lineage of Pangasiidae (Diogo 2004; Rabosky et al. 2018; Schedel et al. 2022). Lacantunia enigmatica was discovered in the Rio Usumacinta basin in Chiapas, Mexico and Kryptoglanis shajii was discovered from subterranean waters in Kerala, India (Rodiles-Hernández et al. 2005; Vincent and Thomas 2011). Both species were each classified in monotypic taxonomic families, Lacantuniidae and Kryptoglanidae (Rodiles-Hernández et al. 2005; Britz, Kakkassery, et al. 2014). Molecular analyses resolve L. enigmatica and the African freshwater Claroteidae as sister lineages (Lundberg et al. 2007; Rabosky et al. 2018). Morphological characters suggest K. shajii is closely related to Siluridae (Lundberg et al. 2014).

  • Composition. There are currently 2,408 living species of Siluroidei (Ferraris 2007; Fricke et al. 2023) that includes Conorhynchos conirostris, Kryptoglanis shajii, Lacantunia enigmatica, Rita (ritas), and species classified in Ailiidae, Akysidae, Amblycipitidae, Amphiliidae, Anchariidae, Ariidae, Aspredinidae, Auchenipteridae, Auchenoglanididae (auchenoglanids), Austroglanis, Bagridae, Cetopsidae, Chaca (squarehead catfishes), Clariidae, Claroteidae, Cranoglanis, Doradidae, Heptapteridae, Heteropneustes, Horabagridae, Ictaluridae, Malapteruridae, Mochokidae, Pangasiidae (shark catfishes), Phreatobius (underground catfishes), Pimelodidae, Plotosidae (eeltail catfishes), Pseudopimelodidae, Schilbeidae, Siluridae, and Sisoridae. Over the past 10 years 236 new living species of Siluroidei have been described (Fricke et al. 2023), comprising 9.8% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Siluroidei include (1) protractor hyoideus differentiated into pars dorsalis, ventralis, and lateralis (Diogo 2004), (2) articulatory surface of autopalatine for neurocranium directed mesially (Diogo 2004), (3) coronomeckelian bone reduced (Diogo 2004), (4) barbels located on anterior rim of posterior nostril (Lundberg et al. 2014), and (5) parasphenoid positioned along anterior margin of trigeminofacial foramen (Lundberg et al. 2014).

  • Synonyms. There are no synonyms of Siluroidei.

  • Comments. Siluriformes is a clade that was long recognized as a taxonomic group and its composition was unchanged in post-Darwinian and phylogenetic classifications, but Siluroidei is a subclade discovered as a result of phylogenetic analyses in the first 10 years of the 21st century (Diogo 2004; Sullivan et al. 2006). Work remains in resolving the phylogenetic relationships among the lineages of Siluroidei, with initial phylogenomic analyses showing considerable potential (Arcila et al. 2017).

  • The earliest fossils of Siluroidei are Campanian (83.2–72.2 Ma) and Maastrichtian (72.2–66.0 Ma) pectoral spines and fragments of skull bones of Ariidae in Argentina (Cione 1987; Arratia and Cione 1996; Gayet and Meunier 1998). Relaxed molecular clock analyses estimate the crown age of Siluroidei between 100 and 105 million years ago (Lundberg et al. 2007).

  • img-z70-2_03.gif

    Characiformes C. T. Regan 1911d:15
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Crenuchus spilurus Günther 1863a, Alestes inferus Stiassny, Schelly, and Mamonekene 2009, Charax gibbosus (Linnaeus 1758), and Charax metae Eigenmann 1922, but not Citharinus congicus Boulenger 1897. This is a minimum-crown-clade definition with an external specifier.

  • Etymology. From the ancient Greek χάραξ (k̍α͡ɹɹæks) as a name for species of Sparidae that exhibit teeth on the oral jaws (D. W. Thompson 1947:284–285). The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 911.

  • Reference phylogeny. A phylogeny of 293 species of Characiformes inferred from DNA sequences of 1,288 ultraconserved element (UCE) loci (Melo, Sidlauskas, et al. 2022, fig. 1). Although Charax gibbosus is not included in the reference phylogeny, it resolves in a clade with three other species of Charax in a phylogenetic analysis of morphological characters (Mattox and Toledo-Piza 2012, fig. 41). See Figure 11 for the phylogenetic relationships among the major lineages of Characiformes.

  • Phylogenetics. The first phylogenetic studies of Characiformes utilized morphological characters to investigate relationships among the subclades Curimatidae (toothless characiforms), Prochilodontidae (flannelmouth characiforms), Anostomidae (toothed headstanders) (Vari 1983), Ctenoluciidae (pike characids), Lebiasinidae (pencilfishes), Hepsetus (African pikes), and Erythrinidae (trahiras) (Vari 1995). Phylogenetic analyses of molecular and morphological matrices consistently support the monophyly of Characiformes relative to Cithariniformes and other otophysans (Ortí 1997; Ortí and Meyer 1997; Buckup 1998; Calcagnotto et al. 2005; Hubert et al. 2005a, 2005b; Mirande 2009; Oliveira et al. 2011; Arcila et al. 2017, 2018; de Pinna et al. 2018; Betancur-R. et al. 2019; Burns and Sidlauskas 2019; Melo, Sidlauskas, et al. 2022). There is extensive incongruence among phylogenetic analyses of Characiformes; the analysis of multiple morphological datasets results in different phylogenies (Buckup 1998; Mirande 2009), different trees are inferred from different molecular datasets (e.g., Ortí and Meyer 1997; Oliveira et al. 2011), and there are substantial differences between phylogenies inferred from morphological and molecular datasets (e.g., Vari 1995; Mirande 2009; Betancur-R. et al. 2019; Melo, Sidlauskas, et al. 2022). Two sets of relationships that are congruent between phylogenies inferred from morphological and molecular datasets are the resolution of the phenotypically unique Tarumania walkerae (Muck Fish) as the sister lineage of all other species of Erythrinidae (Arcila et al. 2018; de Pinna et al. 2018; Melo, de Pinna, et al. 2022) and the monophyly of Anostomoidea that contains Anostomidae, Chilodontidae (headstanders), Curimatidae, and Prochilodontidae (Vari 1983; Buckup 1998; Oliveira et al. 2011; Dillman et al. 2016; Arcila et al. 2017, 2018; Melo et al. 2018; Betancur-R. et al. 2019; Burns and Sidlauskas 2019; Melo, Sidlauskas, et al. 2022).

  • Molecular phylogenetic studies with dense taxon sampling using either collections of Sanger-sequenced mtDNA and nuclear genes or phylogenomic datasets exhibit adequate congruence to highlight several consistent results. Crenuchidae (South American darters) is the sister lineage of all other Characiformes (Oliveira et al. 2011; Arcila et al. 2017, 2018; Betancur-R. et al. 2019; Burns and Sidlauskas 2019; Melo, de Pinna, et al. 2022; Melo, Sidlauskas, et al. 2022). The lineage Chalceus (toucanfishes) was traditionally classified in Alestidae (African tetras) (Zanata and Vari 2005; Mirande 2009, 2010), but is resolved as the sister lineage of a clade containing Acestrorhynchidae (needlejaws), Bryconidae (South American trouts), Characidae (tetras), Gasteropelecidae (freshwater hatchetfishes), Iguanodectidae (tetras), and Triportheidae (elongate hatchetfishes) (Arroyave and Stiassny 2011; Oliveira et al. 2011; Arcila et al. 2017, 2018; Betancur-R. et al. 2019; Burns and Sidlauskas 2019; Melo, Sidlauskas, et al. 2022). The predatory lineages Ctenoluciidae and Hepsetidae are sister lineages in a morphological phylogeny (Buckup 1998); however, molecular phylogenies resolve a monophyletic group containing both African characiform lineages Hepsetidae and Alestidae (Oliveira et al. 2011; Arcila et al. 2017, 2018; Betancur-R. et al. 2019; Melo, de Pinna, et al. 2022; Melo, Sidlauskas, et al. 2022). The traditional delimitation of the species-rich Characidae (Lima et al. 2003; Mirande 2009, 2010) is not monophyletic in molecular phylogenies, prompting the elevation of Acestrorhynchidae, Bryconidae, Iguanodectidae, and Triportheidae; Characidae was restricted to species lacking a supraorbital (Lucena and Menezes 1998; Mirande 2009; Oliveira et al. 2011).

  • Composition. Characiformes currently contains 2,238 species (Fricke et al. 2023) classified in Acestrorhynchidae, Alestidae, Anostomidae, Bryconidae, Chalceus, Characidae, Chilodontidae, Crenuchidae, Ctenoluciidae, Curimatidae, Cynodontidae (dogtooth characins), Erythrinidae, Gasteropelecidae, Hemiodontidae, Hepsetus, Iguanodectidae, Lebiasinidae, Parodontidae (scrapetooths), Prochilodontidae, Serrasalmidae (pacus), and Triportheidae. Over the past 10 years 317 new living species of Characiformes have been described (Fricke et al. 2023), comprising 14.2% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological synapomorphies of Characiformes include (1) fourth neural arch fused to vertebra (Fink and Fink 1981; Buckup 1998), (2) synchondral joint between third and fourth neural arches reduced or absent (Fink and Fink 1981; Buckup 1998), (3) pelvic girdle slightly emarginate anteriorly (Fink and Fink 1981), (4) medial portion of joint between mesethmoid and vomer either flat or covered by midsagittal osseus or cartilaginous crest (Buckup 1998), and (5) A1 and A2 muscles of adductor mandibulae completely separated at their origins (Datovo and Castro 2012).

  • Synonyms. Heterognathi (T. N. Gill 1893:131; Jordan 1923:134–138), Characoidei (Greenwood et al. 1966:383–384), and Characoidea (McAllister 1968:69; Rosen and Greenwood 1970:23; T. R. Roberts 1973:377) are approximate synonyms of Characiformes. Characoidei (Buckup 1998, table 3; Betancur-R et al. 2017:17) is an ambiguous synonym of Characiformes.

  • Comments. As delimited here, Characiformes resolves as a monophyletic group in morphological and molecular phylogenetic analyses (e.g., Buckup 1998; Melo, Sidlauskas, et al. 2022). Characiformes and Cithariniformes were classified together in a more inclusive Characiformes from the mid-19th century to the present day (Günther 1864a; Betancur-R et al. 2017). Molecular phylogenetic analyses consistently fail to resolve Characiformes and Cithariniformes as a monophyletic group relative to other clades of Otophysi. The phylogenies of Characiformes inferred from phylogenomic datasets are not only resolving relationships among the most inclusive lineages in the clade (Arcila et al. 2017; Betancur-R. et al. 2019; Melo, Sidlauskas, et al. 2022), but also illuminating the effect of Gondwanan fragmentation on the distribution of characiforms in South America and Africa (Melo, Sidlauskas, et al. 2022). South American characiforms are paraphyletic relative to the clade containing the African Alestidae and Hepsetidae. The relaxed molecular clock age estimate for the divergence of African characiforms is consistent with the timing of the separation of South America and Africa (Melo, Sidlauskas, et al. 2022), validating Characiformes as an iconic example of continental-drift-driven vicariance in the diversification of freshwater lineages (Lundberg 1993; Ortí and Meyer 1997).

  • The earliest fossil Characiformes are from the Maastrichtian (72.2–66.0 Ma) in Bolivia and include intermediate teeth and skeletal fragments identified as species of Acestrorhynchidae, Characidae, and Serrasalmidae (Gayet et al. 2001, 2003). Isolated teeth from the Cenomanian (100.5–93.9 Ma) of Morocco are often cited as the earliest characiform fossils (Dutheil 1999; Malabarba and Malabarba 2010), but these teeth may be attributed to pan-lepisosteiforms (Cavin 2017:105). Bayesian relaxed molecular clock analyses of Characiformes result in an average posterior crown age estimate of 129.4 million years ago, with the credible interval ranging between 110.0 and 148.7 million years ago (Melo, Sidlauskas, et al. 2022).

  • img-z73-3_03.gif

    FIGURE 11.

    Phylogenetic relationships of the major living lineages and fossil taxa of Characiformes. Filled circles identify the common ancestor of clades, with formal names defined in the clade accounts.

    img-z71-1_03.jpg

    Euteleostei P. H. Greenwood, G. S. Myers,
    D. E. Rosen, and S. H. Weitzman 1967:227
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Lepidogalaxias salamandroides Mees 1961, Salmo salar Linnaeus 1758, and Perca fluviatilis Linnaeus 1758, but not Clupea harengus Linnaeus 1758. This is a minimum-crown-clade definition with an external specifier.

  • Etymology. From the ancient Greek εὖ (ˌiːj̍uː), meaning good or well; τέλειoς (t̍εlƗˌo͡Ʊz), meaning perfect or complete; and ὀστέoν (̍αːstIәn), meaning bone.

  • Registration number. 912.

  • Reference phylogeny. A phylogeny inferred from a phylogenomic dataset composed of DNA sequences from more than 1,100 exons (Hughes et al. 2018, fig. S2). Phylogenetic relationships among the major living and fossil lineages of Euteleostei are presented in Figure 7. The placement of the pan-argentiniform †Surlykus; the pan-salmoniforms †Barcarenichthys, †Kermichthys, †Pyrenichthys, and †Stompooria; the pan-stomiat †Nybelinoides; and the pan-osmeriform †Spaniodon are on the basis of inferences from morphology (Taverne 1982, 1992; Gayet and Lepicard 1985; Gayet 1988b; Anderson 1998; Fielitz 2002; Taverne and Filleul 2003; Guinot and Cavin 2018; Schrøder and Carnevale 2023).

  • Phylogenetics. Euteleostei is resolved as monophyletic in molecular phylogenetic studies that range from analysis of whole mtDNA genomes (J. Li, Xia, et al. 2010; Campbell, López, et al. 2013) to DNA sequences from multiple nuclear and mtDNA genes (Burridge et al. 2012; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Davis et al. 2016), and phylogenomic datasets (Campbell, Alfaro, et al. 2017; Hughes et al. 2018; Straube et al. 2018; Rosas Puchuri 2021). A phylogenetic analysis of 42 morphological characters resolves the otocephalan lineage Alepocephaliformes nested within Euteleostei as the sister lineage of Argentiniformes, nests Stomiiformes in Neoteleostei, and places Esocidae (pikes and mudminnows) as the sister lineage of Neoteleostei (G. D. Johnson and Patterson 1996). A supertree analysis that utilized phylogenies resulting from morphological and molecular studies as input trees resolved Salmoniformes as the sister lineage of a clade named Zoroteleostei that includes all other euteleosts (M. V. H. Wilson and Williams 2010). The phylogenies of Euteleostei presented in G. D. Johnson and Patterson (1996) and Wilson and Williams (2010) are incongruent with trees inferred from molecular phylogenetic analyses (e.g., J. Li, Xia, et al. 2010; Near, Eytan, et al. 2012; Hughes et al. 2018).

  • One of the most remarkable results from molecular phylogenetic analyses of fishes is the resolution of the unique and enigmatic freshwater Lepidogalaxias salamandroides (Salamanderfish) as the sister lineage of all other Euteleostei (J. Li, Xia, et al. 2010; McDowall and Burridge 2011; Burridge et al. 2012; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Campbell, López, et al. 2013; Davis et al. 2016; W. L. Smith et al. 2016; Campbell, Alfaro, et al. 2017; Hughes et al. 2018; Straube et al. 2018; Rosas Puchuri 2021; Mu et al. 2022). Within euteleosts, molecular studies consistently resolve three sets of sister lineages: Salmonidae (salmons and trouts) and Esocidae (includes Umbridae); Stomiiformes and Osmeriformes; and a lineage containing Galaxiidae (galaxiids) and Neoteleostei (Burridge et al. 2012; Near, Eytan, et al. 2012; Davis et al. 2016; Straube et al. 2018; Rosas Puchuri 2021). The phylogenetic relationships of Argentiniformes remain unresolved, with molecular studies resulting in four different hypotheses: as the sister lineage of the clade containing Salmonidae and Esocidae (C. H. Li et al. 2008; J. Li, Xia, et al. 2010; Near, Eytan, et al. 2012; Campbell, López, et al. 2013; Hughes et al. 2018; Straube et al. 2018; Rosas Puchuri 2021); the sister lineage of a clade containing Galaxiidae, Salmonidae, and Esocidae (Betancur-R, Broughton, et al. 2013); the sister lineage of a clade containing Stomiiformes, Osmeriformes, and Galaxiidae (Burridge et al. 2012); or as the sister lineage of a clade containing Stomiiformes, Osmeriformes, Galaxiidae, and Neoteleostei (Campbell, Alfaro, et al. 2017; Rosas Puchuri 2021).

  • Composition. Euteleostei currently consists of more than 21,405 species (Fricke et al. 2023) that include Lepidogalaxias salamandroides and species classified in Salmoniformes, Stomiatii, Argentiniformes, Galaxiidae, and Neoteleostei. Fossil taxa include the pan-argentiniform †Surlykus, the pan-stomiat †Nybelinoides (Taverne 1982), and the pan-salmoniforms †Kermichthys (Taverne 1992), †Barcarenichthys (Gayet 1988b, 1989), †Stompooria (Anderson 1998), and †Pyrenichthys (Gayet and Lepicard 1985). Details of the locations and ages of the fossil taxa are presented in Appendix 1. Over the past 10 years 1,789 new living species of Euteleostei have been described (Fricke et al. 2023), comprising 8.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Euteleostei include (1) presence of stegural, a membranous outgrowth of uroneural 1 (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (2) caudal median cartilages present (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), and (3) unique supraneural shape (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010). The first two of these proposed synapomorphies for Euteleostei are also present in Alepocephaliformes, which is nested in Otocephala and distantly related to Euteleostei (Figure 2).

  • Synonyms. Protacanthopterygii (Greenwood et al. 1966:366–387, 394–396; Wiley and Johnson 2010:141–143; Betancur-R et al. 2017:18), Zoroteleostei (M. V. H. Wilson and Williams 2010:404; J. S. Nelson et al. 2016:251), and Osmeromorpha (J. S. Nelson et al. 2016:252) are partial synonyms of Euteleostei. Euteleosteomorpha (Wiley and Johnson 2010:140; Betancur-R et al. 2017:18) is an ambiguous synonym of Euteleostei.

  • Comments. Along with Osteoglossomorpha, Elopomorpha, and Otocephala, Euteleostei is one of the four major clades of Teleostei (Dornburg and Near 2021). On its initial delimitation, Euteleostei included Ostariophysi (Greenwood et al. 1966, 1967), which was accepted in subsequent studies and classifications (Rosen 1973, 1974; Travers 1981; W. L. Fink and Weitzman 1982; Lauder and Liem 1983; W. L. Fink 1984a; J. S. Nelson 1984:117–119, 1994:124–125; Sanford 1990; Begle 1992). On the basis of the morphology of the teleost skull occipital region, Rosen (1985) suggested that ostariophysans, esocoids, and argentinoids are not euteleosts. Within Euteleostei, the presence of acellular bone was proposed as a synapomorphy for a clade containing Esocidae, Osmeriformes, and Neoteleostei (L. R. Parenti 1986). With the consistent resolution of Otocephala as a clade that includes Ostariophysi and Clupeiformes in molecular and morphological phylogenetic analyses (e.g., Lê et al. 1993; Lecointre and Nelson 1996; Arratia 1997; Near, Eytan, et al. 2012; Straube et al. 2018), classifications no longer include Ostariophysi in Euteleostei and thus closely match the composition of the clade to which we are applying that name (J. S. Nelson 2006:189; Wiley and Johnson 2010; J. S. Nelson et al. 2016:241; Betancur-R et al. 2017; Dornburg and Near 2021).

  • The earliest fossil Euteleostei is the pan-stomiat †Nybelinoides brevis from the Barremian and Aptian (126.5–113.2 Ma) of Belgium (Appendix 1; Taverne 1982; Guinot and Cavin 2018). Bayesian relaxed molecular clock analyses of Euteleostei result in an average posterior crown age estimate of 210.5 million years ago, with the credible interval ranging between 196.4 and 223.8 million years ago (Hughes et al. 2018).

  • img-z75-3_03.gif

    Argentiniformes G. D. Johnson and C. Patterson 1996:315
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Argentina sphyraena Linnaeus 1758 and Microstoma microstoma (Risso 1810). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek ἀργύρεoς (̍αː͡ɹɡjjƱɹɹɪo͡Ʊz), meaning silvery. The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 913.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 1,133 exons (Rosas Puchuri 2021, fig. 3.1). The phylogenetic relationships of the major lineages of Argentiniformes are presented in Figure 7.

  • Phylogenetics. Over the past century Argentiniformes was classified with combinations of Salmonidae (salmons and trouts), Alepocephaliformes, Galaxiidae (galaxiids), Osmeriformes, Stomiiformes, Esocidae (pikes and mudminnows), and Myctophiformes (Gosline 1960; Greenwood et al. 1966; G. J. Nelson 1970a). Greenwood and Rosen (1971) hypothesized that Argentiniformes and Alepocephaliformes are sister lineages as evidenced by the presence of a modified posterior pharyngobranchial structure they named the crumenal organ, which was the basis for the resolution of this clade in subsequent morphological studies (Begle 1992; G. D. Johnson and Patterson 1996). Molecular phylogenetic analyses consistently resolve Argentiniformes and Alepocephaliformes as distantly related: Alepocephaliformes is related to Clupeiformes and Ostariophysi in Otocephala and Argentiniformes is phylogenetically nested in Euteleostei (Figure 2; Ishiguro et al. 2003; Lavoué, Miya, Poulsen, et al. 2008; J. Li, Xia, et al. 2010; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Davis et al. 2016; Campbell, Alfaro, et al. 2017; Hughes et al. 2018; Straube et al. 2018; Rosas Puchuri 2021).

  • Morphological and molecular phylogenetic analyses consistently support the monophyly of Argentiniformes (Begle 1992; Patterson and Johnson 1995; Ishiguro et al. 2003; J. Li, Xia, et al. 2010; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Straube et al. 2018; Schrøder and Carnevale 2023). Phylogenetic analysis of morphological characters resolves Argentinidae (argentines) as the sister lineage of all other Argentiniformes, with Bathylagidae (deep-sea smelts) and Opisthoproctidae (barreleyes) as sister taxa (Rosen 1974) or Bathylagidae and Microstomatidae (pencilsmelts) as sister taxa (Patterson and Johnson 1995). Molecular phylogenetic analyses resolve the four major lineages of Argentiniformes into two sets of sister lineages: one clade containing Argentinidae and Opisthoproctidae and the other including Bathylagidae and Microstomatidae (J. Li, Xia, et al. 2010; Rosas Puchuri 2021).

  • Composition. There are currently 100 living species of Argentiniformes (Fricke et al. 2023) classified in Argentinidae, Bathylagidae, Microstomatidae, and Opisthoproctidae. Over the past 10 years there have been eight new living species of Argentiniformes described (Fricke et al. 2023), comprising 8.0% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Argentiniformes include (1) metapterygoid reduced in size (Begle 1992; G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (2) endopterygoid teeth absent (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (3) parietal carrying commissural sensory canal (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (4) premaxilla without teeth (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (5) maxilla without teeth (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (6) supramaxillae absent (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (7) basibranchials 1-3 without teeth (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (8) epibranchial 4 with distinct levator process (G. D. Johnson and Patterson 1996), (9) pharyngobranchial 2 without teeth (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (10) pharyngobranchial 3 without teeth (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), and (11) supraneurals develop in “pattern 2” (G. D. Johnson and Patterson 1996).

  • Synonyms. Argentinoidei (Greenwood et al. 1966:394; Wiley and Johnson 2010:141) and Argentinoidea (Greenwood and Rosen 1971:39; J. S. Nelson 1984:160–162, 1994:179–181; Begle 1992:351; Johnson and Patterson 1996:309) are ambiguous synonyms of Argentiniformes.

  • Comments. Subsequent to the resolution of Alepocephaliformes within Otocephala (e.g., Ishiguro et al. 2003), classifications of Actinopterygii consistently use the group name Argentiniformes for the clade containing Argentinidae, Bathylagidae, Microstomatidae, and Opisthoproctidae (Davis et al. 2016; J. S. Nelson et al. 2016:252–254; Betancur-R et al. 2017; Dornburg and Near 2021).

  • The earliest fossils of Argentiniformes are otoliths from the Maastrichtian (72.2–66.0 Ma) of Maryland, USA, identified as Argentinidae and †Argentina voigti from Bavaria, Germany (Nolf and Stringer 1996; Schwarzhans 2010; Schwarzhans and Jagt 2021; Stringer and Schwarzhans 2021). The earliest skeletal argentiniform fossil is †Glossanodon musceli from the Rupelian (33.9–27.3 Ma) of the Czech Republic and Poland (Paucă 1929; Gregorová 2011; Přikryl et al. 2016). Relaxed molecular clock analyses estimate the crown age of Argentiniformes between 34.5 and 76.5 million years ago (Near, Eytan, et al. 2012).

  • img-z76-7_03.gif

    Salmoniformes P. H. Greenwood, D. E. Rosen,
    S. H. Weitzman, and G. S. Myers 1966:394

  • Definition. The least inclusive crown clade that contains Salmo salar Linnaeus 1758 and Esox lucius Linnaeus 1758. This is a minimum-crown-clade definition, but the clade is not defined using the PhyloCode.

  • Etymology. Salmo is Latin for Salmo trutta, dating to Pliny (N.H. 9.68) in the first century CE (Andrews 1955). The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Reference phylogeny. A phylogeny inferred from a phylogenomic dataset consisting of DNA sequences from more than 1,100 exons (Hughes et al. 2018, fig. S2). Phylogenetic relationships of the major lineages of Salmoniformes are presented in Figure 7. The placements of the fossil taxa †Oldmanesox and †Estesesox in the phylogeny follow inferences from morphology (Brinkman et al. 2014).

  • Phylogenetics. Morphological studies result in a disparate set of phylogenetic relationships for Salmonidae (trouts and salmons) and Esocidae (pikes and mudminnows). Salmonidae are resolved as the sister lineage of Osmeriformes (Rosen 1974; G. D. Johnson and Patterson 1996); Galaxiidae (Rosen 1974); Neoteleostei (Lauder and Liem 1983; W. L. Fink 1984a); a clade containing Osmeriformes, Argentiniformes, and Alepocephaliformes (Sanford 1990); or unresolved among euteleosts (W. L. Fink and Weitzman 1982; Begle 1991, 1992). Morphological studies place Esocidae as the sister lineage of a clade containing Argentiniformes, Galaxiidae, Salmonidae, and Osmeriformes (Rosen 1974); a clade containing Salmonidae and Osmeriformes (Rosen 1974); the sister lineage of all other Euteleostei (W. L. Fink and Weitzman 1982; W. L. Fink 1984a; Sanford 1990; Begle 1991, 1992); or as the sister lineage of Neoteleostei (G. D. Johnson and Patterson 1996). One set of morphological studies resolves Salmonidae and Esocidae as sister lineages (R. R. G. Williams 1987; M. V. H. Wilson and Williams 2010), a result that is congruent with molecular phylogenetic analyses (Ishiguro et al. 2003; Lopez et al. 2004; Osinov and Lebedev 2004; C. H. Li et al. 2008; Davis 2010; J. Li, Xia, et al. 2010; Burridge et al. 2012; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Campbell, López, et al. 2013; Faircloth et al. 2013; Davis et al. 2014, 2016; Campbell, Alfaro, et al. 2017; Hughes et al. 2018; Straube et al. 2018; Musilova et al. 2019; Harvey et al. 2021; Rosas Puchuri 2021; Mu et al. 2022).

  • Composition. Salmoniformes includes 275 species classified in Salmonidae and Esocidae (Fricke et al. 2023). Fossil taxa include the Cretaceous pan-esocids †Estesesox from the Campanian and Maastrichtian (83.2–66.0 Ma) of Montana, USA, and †Oldmanesox from the Campanian (83.2–72.2 Ma) of Alberta, Canada (M. V. H. Wilson et al. 1992). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years 30 new living species of Salmoniformes have been described, comprising 10.9% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Salmoniformes include (1) hyomandibular with unique process that extends towards the symplectic and metapterygoid (R. R. G. Williams 1987; M. V. H. Wilson and Williams 2010), (2) absence of distinct ligament connecting adductor mandibulae and maxilla-mandibular ligament (R. R. G. Williams 1987; M. V. H. Wilson and Williams 2010), and (3) absence of caudal scutes, median bony plates that form anterior to procurrent caudal fin rays [caudal scutes are also absent in Alepocephaliformes] (G. D. Johnson and Patterson 1996).

  • Synonyms. Protacanthopterygii (M. V. H. Wilson and Williams 2010:404; J. S. Nelson et al. 2016:243–251) is an ambiguous synonym of Salmoniformes.

  • Comments. The earliest fossil Salmoniformes are the western North American pan-esocids †Oldmanesox from the Campanian (83.6–72.1 Ma) and †Estesesox from the Campanian and Maastrichtian (72.1–66.0 Ma) (M. V. H. Wilson et al. 1992; Brinkman et al. 2014). Bayesian relaxed molecular clock analyses of Salmoniformes result in an average posterior crown age estimate of 82.8 million years ago, with the credible interval ranging between 76.8 and 88.3 million years ago (Hughes et al. 2018).

  • img-z77-7_03.gif

    Esocidae C. S. Rafinesque 1815:89

  • Definition. The least inclusive crown clade that contains Esox lucius Linnaeus 1758 and Umbra krameri Walbaum 1792. This is a minimum-crown-clade definition, but the clade is not defined using the PhyloCode.

  • Etymology. Isox is Latin, possibly Celtic or Basque in origin, which was the name for Salmo salar dating to Pliny (N.H. 9.44) in the first century CE (D. W. Thompson 1947:95; Andrews 1955).

  • Reference phylogeny. A phylogeny inferred from analysis of DNA sequences of 53 ultraconserved element (UCE) loci (Campbell, Alfaro, et al. 2017, fig. 1). Phylogenetic relationships of living and fossil lineages of Esocidae are presented in Figure 7. Placements of the fossil taxa †Boltyshia, †Palaeoesox, and †Proumbra in the phylogeny are on the basis of inferences from morphology (J. Gaudant 2012).

  • Phylogenetics. Relationships inferred from morphology place Esox as the sister lineage to a clade previously classified as Umbridae that contains Umbra, Dallia, and Novumbra (Cavender 1969; G. J. Nelson 1972; M. V. H. Wilson and Veilleux 1982); however, a study of meristic and morphometric traits noted the lack of morphological evidence for the monophyly of Umbridae (Reist 1987). To date, there is no phylogenetic investigation of Esocidae that employs explicit analysis of coded morphological character states. Molecular phylogenetic analyses consistently resolve Umbridae as paraphyletic, with Umbra placed as the sister lineage of all other Esocidae (Lopez et al. 2000, 2004; Burridge et al. 2012; Near, Eytan, et al. 2012; Campbell, López, et al. 2013; Campbell, Alfaro, et al. 2017; Marić et al. 2017; Pan et al. 2021). Several molecular phylogenetic studies are aimed at resolving relationships among species of Esox and providing a basis for species discovery and delimitation in the clade (T. Grande et al. 2004; Denys et al. 2014, 2018).

  • Composition. There are currently 13 living species of Esocidae (T. Grande et al. 2004; Lucentini et al. 2011; Denys et al. 2014; Kuehne and Olden 2014; Fricke et al. 2023). The species Dallia admirabilis Chereshnev and D. delicatissima are synonyms of Dallia pectoralis Bean (Campbell and Lopéz 2014; Dyldin et al. 2020). Fossil taxa of Esocidae include †Novumbra oregonensis from the Rupelian (33.90–27.82 Ma) in Oregon (Cavender 1969; Woodburne 2004), several species of Esox (M. V. H. Wilson 1980; L. Grande 1999), and the pan-umbrines †Boltyshia, †Palaeoesox, and †Proumbra (J. Gaudant 2012). Details of the ages and locations of the fossil taxa are given in Appendix 1. Over the past 10 years one new species of Esocidae has been described (Fricke et al. 2023), comprising 7.7% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Esocidae include (1) ethmoidal and antorbital canals present as pitlines (G. J. Nelson 1972; Rosen 1974), (2) presence of mandibulopreopercular, subnasal, and opercular pitlines (G. J. Nelson 1972; Rosen 1974), (3) presence of paired elongate proethmoids (Rosen 1974; G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (4) basibranchial tooth plate in two parts (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (5) second pharyngobranchial conical in shape with tip enclosed in bone (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), (6) single upper pharyngeal tooth plate composed of upper fourth upper pharyngeal (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010), and (7) presence of a single postcleithrum (G. D. Johnson and Patterson 1996; Wiley and Johnson 2010).

  • Synonyms. Esociformes (J. S. Nelson 1994:176–178, 2006:204–206; G. D. Johnson and Patterson 1996:316; López et al. 2004, fig. 2; J. S. Nelson et al. 2016:248–251; Betancur-R et al. 2017:18; Pan et al. 2021, fig. 1), Esocoidei (Berg 1940:429; Gosline 1960:358; Greenwood et al. 1966:394; G. J. Nelson 1972:32; J. S. Nelson 1984:157–159; Wiley and Johnson 2010:142), and Esocoidea (Rosen 1974:311) are ambiguous synonyms of Esocidae. Umbridae (Greenwood et al. 1966:394; G. J. Nelson 1972:32–33; Rosen 1974:311; J. S. Nelson 1984:158_159, 1994:177–178) and Esocinae (López et al. 2000:429) are partial synonyms of Esocidae.

  • Comments. As a consequence of molecular phylogenetic analyses (e.g., López et al. 2000, 2004; Campbell, Alfaro, et al. 2017), the classification of esociform fishes was modified by the inclusion of Dallia and Novumbra into Esocidae with Esox and limiting Umbridae to Umbra (J. S. Nelson 2006:205–206; J. S. Nelson et al. 2016:251). This change makes Umbra and Umbridae redundant group names. Historically, Esocidae and Umbridae were classified as Esocoidei (e.g., Wiley and Johnson 2010) or Esociformes (e.g., Betancur-R et al. 2017); however, these group names are redundant with Esocidae as delimited here. Esocidae is a valid family-group name under the International Code of Zoological Nomenclature (Van der Laan et al. 2014:64).

  • The earliest fossil Esocidae is †Esox tiemani from Tiffanian (60.2–56.8 Ma) North American Land Mammal Age dated rocks in Alberta, Canada (M. V. H. Wilson 1980; L. Grande 1999; Speijer et al. 2020, fig. 28.12) or †Boltyshia brevicauda from the Thanetian (59.2–56.0 Ma) in Ukraine (Cavagnetto and Gaudant 2000; J. Gaudant 2012). Bayesian relaxed molecular clock analyses of Esocidae result in an average posterior age estimate of 88.6 million years ago, with the credible interval ranging between 85.1 and 95.6 million years ago (Campbell, López, et al. 2013).

  • img-z79-2_03.gif

    Stomiatii R. Betancur-R, R. E. Broughton, E. O. Wiley, K. Carpenter, J. A. López, C. Li, N. I. Holcroft, D. Arcila, M. Sanciangco, J. C. Cureton II, F. Zhang, T. Buser, M. A. Campbell, J. A. Ballesteros, A. Roa-Varón, S. Willis, W. C. Borden, T. Rowley, P. C. Reneau, D. J. Hough, G. Lu, T. Grande, G. Arratia, and G. Ortí 2013: app. 2 [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Osmerus mordax (Mitchill 1814) and Stomias boa (Risso 1810). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek στγµᾰ (st̍o͡Ʊmә), meaning mouth.

  • Registration number. 920.

  • Reference phylogeny. A phylogeny inferred from nine Sanger-sequenced nuclear genes (Near, Eytan, et al. 2012, fig. S1). Phylogenetic relationships of the living and fossil lineages of Stomiatii are presented in Figure 7. The placements of the pan-stomiiform †Paravinciguerria and the pan-osmeriform †Spaniodon are on the basis of inferences from morphology (Taverne and Filleul 2003; Carnevale and Rindone 2011).

  • Phylogenetics. A pre-phylogenetic morphological study proposed that Stomiiformes and Osmeridae exhibit a “relatively close relationship” (Weitzman 1967:523). Subsequent morphological studies resulted in varied and incongruent phylogenetic hypotheses among major lineages of Euteleostei and did not resolve Stomiatii as monophyletic (e.g., W. L. Fink 1984a; Rosen 1985; G. D. Johnson and Patterson 1996; Wilson and Williams 2010). Molecular phylogenetic analyses of Euteleostei consistently resolve Stomiatii as a monophyletic lineage that includes Osmeriformes and Stomiiformes (Davis 2010; J. Li, Xia, et al. 2010; Burridge et al. 2012; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Campbell, López, et al. 2013; Davis et al. 2014, 2016; W. L. Smith et al. 2016; Campbell, Alfaro, et al. 2017; Malmstrøm et al. 2017; Hughes et al. 2018; Straube et al. 2018; Musilova et al. 2019; Rosas Puchuri 2021; Mu et al. 2022).

  • Composition. There are currently 500 living species of Stomiatii (Fricke et al. 2023) classified in Osmeriformes and Stomiiformes. Fossil lineages of Stomiatii include the pan-osmeriform †Spaniodon and the pan-stomiiform †Paravinciguerria (Appendix 1; Taverne and Filleul 2003; Carnevale and Rindone 2011). Over the past 10 years 31 new living species of Stomiatii have been described (Fricke et al. 2023), comprising 6.2% of the living species diversity in the clade.

  • Diagnostic apomorphies. There are no known morphological synapomorphies for Stomiatii (Betancur-R et al. 2017; Straube et al. 2018).

  • Synonyms. Stomiati (Betancur-R et al. 2017:18) is a variant spelling of Stomiatii.

  • Comments. Stomiatii is a group name applied to the clade containing Osmeriformes and Stomiiformes (Betancur-R, Broughton, et al. 2013).

  • The earliest fossil Stomiatii is the pan-stomiiform †Paravinciguerria praecursor from the Cenomanian (100.5–93.9 Ma) of Morocco and Sicily (Appendix 1; Khalloufi et al. 2010; Carnevale and Rindone 2011). Bayesian relaxed molecular clock analyses of Stomiatii result in an average posterior crown age estimate of 115.4 million years ago, with the credible interval ranging between 81.1 and 147.6 million years ago (Hughes et al. 2018).

  • img-z79-24_03.gif

    Stomiiformes W. L. Fink and S. H. Weitzman 1982:32
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Stomias boa (Risso 1810), Gonostoma denudatum Rafinesque 1810b, Sternoptyx diaphana Hermann 1781, and Vinciguerria nimbaria (Jordan and Williams in Jordan and Starks 1895). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek στὀµᾰ (st̍o͡Ʊmә), meaning mouth. The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 922.

  • Reference phylogeny. A phylogeny of 99 species of Stomiiformes inferred from a supermatrix of 27 nuclear and mitochondrial genes (Rabosky et al. 2018; J. Chang et al. 2019). The phylogeny is available on the Dryad data repository (Rabosky et al. 2019). Phylogenetic relationships among the major lineages of Stomiiformes are presented in Figure 7.

  • Phylogenetics. Morphological phylogenetic studies consistently support the monophyly of Stomiiformes (Rosen 1973; Weitzman 1974; W. L. Fink and Weitzman 1982; W. L. Fink 1984b; Harold and Weitzman 1996; Harold 1998). Aside from phylogenies with limited taxon sampling (Harold and Weitzman 1996; Harold 1998), there is no morphological phylogenetic analysis of Stomiiformes that includes a comprehensive taxon sampling of the major lineages in the clade. Morphological phylogenetic analyses are aimed at stomiiform subclades: Sternoptychidae (marine hatchetfishes) (Harold 1993, 1994; Harold and Weitzman 1996), Gonostomatidae (bristlemouths) (Harold and Weitzman 1996; Harold 1998), and Stomiidae (barbeled dragonfishes) (W. L. Fink 1984b; W. L. Fink 1985). Weitzman (1974:338) introduced Phosichthyidae (lightfishes) to contain Pollichthys mauli (Stareye Lightfish), Phosichthys argenteus (Silver Lightfish), and Vinciguerria, Yarrella, Polymetme, Ichthyococcus, and Woodsia. There are no morphological apomorphies identified for Phosichthyidae and the group is resolved as paraphyletic in morphological studies (W. L. Fink 1984b; Harold and Weitzman 1996).

  • Molecular phylogenetic analyses resolve Stomiiformes as monophyletic (Miya et al. 2003; J. Li, Xia, et al. 2010; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Davis et al. 2014, 2016; W. L. Smith et al. 2016; Rosas Puchuri 2021). A small number of molecular phylogenetic analyses include a sampling of the major lineages of Stomiiformes (Davis et al. 2014; Kenaley et al. 2014; Rabosky et al. 2018; Rosas Puchuri 2021). In analyses of DNA sequence data, the phosichthyid Vinciguerria is resolved as the sister lineage of all other Stomiiformes (Kenaley et al. 2014; Betancur-R et al. 2017; Rosas Puchuri 2021); however, analysis of translated amino acid sequences from more than 1,000 exons places Vinciguerria as nested well within Stomiiformes (Rosas Puchuri 2021). In molecular phylogenies, Stomiidae, Gonostomatidae, and Sternoptychidae are each resolved as monophyletic, but Phosichthyidae is deeply paraphyletic (Davis et al. 2014; Kenaley et al. 2014; Rabosky et al. 2018; Rosas Puchuri 2021). The former lineages of Phosichthyidae that include Pollichthys mauli, Polymetme, Yarrella, Vinciguerria, and Ichthyococcus do not resolve with other lineages that are delimited in named Linnaean taxonomic families (Figure 7; Rabosky et al. 2018; Rosas Puchuri 2021), and there are no available family-group names to accommodate any of these genera (Van der Laan et al. 2014). We delimit Phosichthyidae to include Phosichthys argenteus and species of Woodsia.

  • Composition. There are currently 458 living species of Stomiiformes (Fricke et al. 2023), including Pollichthys mauli and species classified in Ichthyococcus, Polymetme, Vinciguerria, Yarrella, Gonostomatidae, Phosichthyidae, Sternoptychidae, and Stomiidae. Over the past 10 years there have been 31 new living species of Stomiiformes described (Fricke et al. 2023), comprising 6.8% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Stomiiformes include (1) single broad termination of the second epibranchial that articulates with the second and third pharyngobranchials (Rosen 1973; W. L. Fink and Weitzman 1982; Wiley and Johnson 2010), (2) unique structure of the photophores (W. L. Fink and Weitzman 1982; Wiley and Johnson 2010), (3) type 3 tooth attachment (W. L. Fink and Weitzman 1982; Wiley and Johnson 2010), (4) medial section of adductor mandibulae divided into two sections, dorsal section inserting directly onto the maxilla and ventral portion inserting on primordial ligament (W. L. Fink and Weitzman 1982; Wiley and Johnson 2010), (5) unique crossing pattern of ethmoid-premaxillary ligament (W. L. Fink and Weitzman 1982; Wiley and Johnson 2010), (6) greatly enlarged posterior branchiostegal rays (W. L. Fink and Weitzman 1982; Wiley and Johnson 2010), (7) some branchiostegal rays articulating with ventral hypohyals (W. L. Fink and Weitzman 1982; Wiley and Johnson 2010), (8) rete mirabile located at posterior of swim bladder (W. L. Fink and Weitzman 1982; Wiley and Johnson 2010), (9) part of obliquus dorsalis 4 attached to fourth pharyngobranchial (Springer and Johnson 2004; Wiley and Johnson 2010), and (10) adductor 5 attaches to fourth epibranchial (Springer and Johnson 2004; Wiley and Johnson 2010).

  • Synonyms. Stomiatoidei (Jordan 1923:126–127; Gregory and Conrad 1936:25–27; Gosline 1960:358; Greenwood et al. 1966:372–373, 394; McAllister 1968:48–52; Weitzman 1974:338), Stomiatoidea (Beebe and Crane 1939:69), Stenopterygii (Rosen 1973:509), Stomiatiformes (Rosen 1973:509; Wiley and Johnson 2010:144; Betancur-R et al. 2017:144), and Stomiatia (Wiley and Johnson 2010:144) are ambiguous synonyms of Stomiiformes.

  • Comments. Prior to the development of phylogenetic systematics, lineages of Stomiiformes were consistently recognized as a natural group in taxonomic classifications (Regan 1923a; Gregory and Conrad 1936, fig. 3; Beebe and Crane 1939; Gosline 1960). The group name Stomiiformes has been applied to this clade since the early 1980s (W. L. Fink and Weitzman 1982; J. S. Nelson 1984:172–177, 1994:196–201, 2006:207–212; Near, Eytan, et al. 2012; Davis et al. 2016; J. S. Nelson et al. 2016:259–264; Dornburg and Near 2021), which is why it is selected as the clade name over its synonyms. While consistently resolved as monophyletic in phylogenetic analyses of Teleostei (e.g., Davis et al. 2014; Rabosky et al. 2018), relationships within Stomiiformes are not consistent among molecular analyses and there is no morphological phylogenetic study that includes a robust sampling of the major lineages in the clade. The lack of a robust understanding of the phylogenetic relationships within Stomiiformes is reflected by the deep paraphyly of Phosichthyidae, which prevents the establishment of a ranked Linnaean classification where the taxonomic families reflect monophyletic groups. The lineages not currently placed in Linnaean families are listed with generic names in the classification outlined in Appendix 2 and in the constituent lineages section below.

  • The earliest fossil Stomiiformes is †Eosternoptyx discoidalis, a species of Sternoptychidae from the Bartonian-aged (41.2–37.7 Ma) deposit in the Pabdeh Formation, Iran (Afsari et al. 2014). Relaxed molecular clock analyses estimate the crown age of Stomiiformes to be between 63 and 120 million years ago (Kenaley et al. 2014).

  • img-z81-6_03.gif

    Osmeriformes D. P. Begle 1991:46
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Osmerus eperlanus (Linnaeus 1758), Osmerus mordax (Mitchill 1814), and Retropinna semoni (Weber 1895). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek σσµᾰ (h̍o͡Ʊsme͡I), meaning odor. The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 925.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of nine nuclear genes (Near, Eytan, et al. 2012, fig. S1). Although Osmerus eperlanus is not included in the reference phylogeny, it resolves in a clade with other species of Osmerus in a molecular phylogenetic analysis (Ilves and Taylor 2009, fig. 3). Phylogenetic relationships of living and fossil lineages of Osmeriformes are presented in Figure 7. Placement of the fossil taxon †Speirsaenigma in the phylogeny is on the basis of a phylogenetic analysis of morphological characters (M. V. H. Wilson and Williams 1991).

  • Phylogenetics. Among the multiple morphological studies of relationships among lineages of Euteleostei (McDowall 1969, 1984; G. J. Nelson 1970a; W. L. Fink and Weitzman 1982; W. L. Fink 1984a; Sanford 1990; Begle 1991; G. D. Johnson and Patterson 1996), only Rosen (1974:311) proposed a grouping of Osmeridae (smelts), Plecoglossus altivelis (Ayu), Salangidae (noodlefishes), and Retropinnidae (southern smelts) that is consistent with the current delimitation of Osmeriformes. Molecular phylogenetic analyses consistently resolve Osmeriformes as monophyletic (Waters et al. 2002; López et al. 2004; J. Li, Xia, et al. 2010; Near, Eytan, et al. 2012; Rabosky et al. 2018; Straube et al. 2018). Phylogenies inferred from morphology nest Plecoglossus and Salangidae within Osmeridae (Howes and Sanford 1987; G. D. Johnson and Patterson 1996); however, molecular studies resolve Osmeridae as monophyletic and the sister lineage of a clade containing Plecoglossus and Salangidae (Ilves and Taylor 2009; J. Li, Xia, et al. 2010; Burridge et al. 2012; Near, Eytan, et al. 2012; Rosas Puchuri 2021).

  • Composition. There are currently 42 living species of Osmeriformes (Fricke et al. 2023) that includes Plecoglossus altivelis and species classified in Osmeridae, Salangidae, and Retropinnidae. Fossil taxa of Osmeriformes include †Speirsaenigma lindoei from the Thanetian (59.2–56.0 Ma) in Alberta, Canada (Appendix 1; M. V. H. Wilson and Williams 1991; Lofgren et al. 2004). Over the past 10 years no new living species of Osmeriformes have been described.

  • Diagnostic apomorphies. In an effort to identify morphological apomorphies consistent with the monophyly of Osmeriformes, we used maximum parsimony as executed in Mesquite v. 3.70 (Maddison and Maddison 2021) to map 112 morphological character state changes reported in G. D. Johnson and Patterson (1996, app. 1) onto a phylogeny of Euteleostei that matches the tree in Figure 7. Relationships within Osmeridae and Retropinnidae matched those inferred in molecular phylogenetic analyses (Waters et al. 2002; Ilves and Taylor 2009). There is one character state change identified in the mapping exercise that appears as an unambiguous apomorphy; however, several other character state changes exhibit a pattern that is compelling for the hypothesis of osmeriform monophyly. The six characters include (1) pelvic girdle with ventral condyle (McDowall 1969, 1984; G. D. Johnson and Patterson 1996), (2) vomer without shaft [species of Salangidae have a vomer with a shaft and Plecoglossus lacks a vomer] (G. D. Johnson and Patterson 1996), (3) fifth epibranchial fused with fourth epibranchial at both ends [species of Plecoglossus, Prototroctes, and Stokellia have fifth epibranchial that is free or fused with fourth epibranchial only at its lower end] (G. D. Johnson and Patterson 1996), (4) epineural bones or ligaments originate on the centrum of several anterior vertebrate [the epineural bones or ligaments originate on the neural arch in species of Prototroctes and Retropinna] (G. D. Johnson and Patterson 1996), (5) cleithrum with narrow columnar process [cleithrum in species of Salangidae lacks a process] (McDowall 1969; G. D. Johnson and Patterson 1996), and (6) presence of an enlarged first pectoral radial that partially covers the scapula [the first pectoral radial is unmodified in species of Mallotus and Salangidae] (G. D. Johnson and Patterson 1996).

  • Synonyms. Osmeroidea (Rosen 1974:311) is an ambiguous synonym of Osmeriformes. Osmeroidei (G. D. Johnson and Patterson 1996:307; Wiley and Johnson 2010:142) is a partial synonym of Osmeriformes.

  • Comments. When first applied as a group name, Osmeriformes was delimited as a polyphyletic group that included Plecoglossus, Osmeridae, Salangidae, Retropinnidae, Argentiniformes, Alepocephaliformes, Lepidogalaxias, and Galaxiidae (Begle 1991; J. S. Nelson 1994:178–189); the paraphyletic group containing Plecoglossus, Osmeridae, Salangidae, Retropinnidae, Lepidogalaxias, and Galaxiidae (J. S. Nelson 2006:194–199); and the monophyletic group as delimited here (Davis et al. 2016; J. S. Nelson et al. 2016:256–259; Betancur-R et al. 2017; Rosas Puchuri 2021). The name Osmeriformes was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest fossil Osmeriformes is the pan-plecoglossid †Speirsaenigma lindoei from the Thanetian (59.2–56.0 Ma) in Alberta, Canada (M. V. H. Wilson and Williams 1991; Lofgren et al. 2004). Relaxed molecular clock analyses estimate the age of Osmeriformes to be between 80.0 and 125.7 million years ago (Near, Eytan, et al. 2012).

  • img-z83-3_03.gif

    Neoteleostei G. J. Nelson 1969a:534
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Ateleopus japonicus Bleeker 1853b (Ateleopodidae), Alepisaurus ferox Lowe 1833 (Aulopiformes), Scopelengys tristis Alcock 1890 (Myctophiformes), and Micropterus salmoides (Lacépède 1802) (Centrarchiformes), but not Osmerus mordax (Mitchill 1814) (Osmeriformes). This is a minimum-crown-clade definition with an external specifier.

  • Etymology. From the ancient Greek νέoϛ (n̍iːo͡Ʊz), meaning new; τέλειoς (t̍εlƗᵻo͡Ʊz), meaning perfect or complete; and ὀστέoν (̍αːstIәn), meaning bone.

  • Registration number. 926.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of nine concatenated Sanger-sequenced nuclear genes (Near, Eytan, et al. 2012, fig. S1). See Figures 2 and 12 for the phylogeny of lineages comprising Neoteleostei.

  • Phylogenetics. When initially delimited, Neoteleostei was represented in phylogenetic trees of the major lineages of vertebrates as a clade including Atherinoidei, Myctophiformes, Paracanthopterygii (sensu lato), and Acanthopterygii (G. J. Nelson 1969a). Neoteleostei was expanded to include Stomiiformes on the basis of three morphological synapomorphies (Rosen 1973); however, two of these traits were subsequently rejected on the basis of homology and phylogenetic incongruence (W. L. Fink and Weitzman 1982). The monophyly of a Neoteleostei that includes Stomiiformes was widely accepted in reviews of actinopterygian phylogeny and classification (Lauder and Liem 1983; Stiassny 1986; G. D. Johnson 1992; J. S. Nelson 1994, 2006; A. C. Gill and Mooi 2002; Stiassny et al. 2004; Wiley and Johnson 2010).

  • Molecular phylogenetic studies resolve Neoteleostei as a monophyletic group to the exclusion of Stomiiformes (Davis 2010; J. Li, Xia, et al. 2010; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Davis et al. 2014, 2016; W. L. Smith et al. 2016; Malmstrøm et al. 2017; Hughes et al. 2018; Musilova et al. 2019; Mu et al. 2022). Within Neoteleostei molecular phylogenies resolve either Ateleopodidae (jellynose fishes) (e.g., Near, Eytan, et al. 2012), Aulopiformes (e.g., Hughes et al. 2018), or a clade containing Ateleopodidae and Aulopiformes (Mu et al. 2022) as the sister lineage of all other Neoteleostei. Morphology of the dorsal gill arch musculature suggests that Ateleopodidae forms a clade with Aulopiformes (Springer and Johnson 2004; Wiley and Johnson 2010). The uncertainty in the phylogenetic relationships of Ateleopodidae is a challenge to the delimitation of Eurypterygii, which is a hypothesized clade that includes Aulopiformes and Ctenosquamata (G. D. Johnson 1992). There is no morphological phylogenetic analysis of discretely coded morphological character state changes aimed at resolving relationships among the lineages of Neoteleostei.

  • Composition. Currently there are more than 20,460 living species of Neoteleostei (Fricke et al. 2023) classified in Ateleopodidae, Aulopiformes, and Ctenosquamata. Over the past 10 years there have been 1,707 new living species of Neoteleostei described (Fricke et al. 2023), comprising 8.3% of the living species diversity in the clade.

  • Diagnostic apomorphies. The morphological apomorphies proposed for Neoteleostei considered a delimitation of the clade that includes Stomiiformes. The apomorphies include (1) presence of retractor dorsalis (Rosen 1973; G. D. Johnson 1992; Olney et al. 1993; Wiley and Johnson 2010), (2) third internal levator inserts on fifth upper pharyngeal tooth plate (G. D. Johnson 1992; Olney et al. 1993; Wiley and Johnson 2010), (3) type 4 tooth attachment (G. D. Johnson 1992; Olney et al. 1993; Wiley and Johnson 2010), (4) transversus dorsalis attaches to second epibranchial (Springer and Johnson 2004; Wiley and Johnson 2010), and (5) presence of transversus epibranchialis 2 (Springer and Johnson 2004; Wiley and Johnson 2010).

  • Synonyms. There are no synonyms of Neoteleostei.

  • Comments. The monophyly of Neoteleostei is one of the key discoveries in the early efforts of applying phylogenetic systematics to the relationships of ray-finned fishes (G. J. Nelson 1969a; Rosen 1973). There have been slight modifications to the original delimitation of Neoteleostei, but the integrity of the clade remains largely intact (Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013). Neoteleostei is consistently applied as the group name for the clade as delimited here (Near, Eytan, et al. 2012; Davis et al. 2016; J. S. Nelson et al. 2016:264; Betancur-R et al. 2017; Hughes et al. 2018). Neoteleostei contains more than 58% of the living species diversity of ray-finned fishes and composes the dominant group of vertebrates occupying marine habitats.

  • The earliest fossil Neoteleostei is the aulopiform †Atolvorator longipectoralis from the Barremian (126.5–121.4 Ma) in the Cretaceous of Brazil (Gallo and Coelho 2008; Newbrey and Konishi 2015). Bayesian relaxed molecular clock analyses of Neoteleostei result in an average posterior crown age estimate of 161.0 million years ago, with the credible interval ranging between 152.2 and 172.0 million years ago (Hughes et al. 2018).

  • img-z85-7_03.gif

    FIGURE 12.

    Phylogenetic relationships of the major living lineages and fossil taxa of Neoteleostei, Aulopiformes, Ctenosquamata, Myctophiformes, Acanthomorpha, and Lampriformes. Filled circles identify the common ancestor of clades with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z84-1_03.jpg

    Aulopiformes D. E. Rosen 1973:509
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Aulopus filamentosus (Bloch 1792a) and Alepisaurus ferox Lowe 1833. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek αὐλωπίαϛ (ͻːl̍o͡Ʊpi͡әz) of unknown origin, a name applied to species of Scombridae by ancient Mediterranean authors (D. W. Thompson 1947:20–21). The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 928.

  • Reference phylogeny. A phylogeny inferred from a combined dataset of mitochondrial and nuclear gene DNA sequences and 138 morphological characters (Davis 2010, fig. 7). Phylogenetic relationships of living and fossil lineages of Aulopiformes are presented in Figure 12. The fossil taxa are placed in the phylogeny on the basis of analyses of morphological characters (Fielitz 2004; Davis and Fielitz 2010; Marramà and Carnevale 2017; Beckett, Giles, et al. 2018).

  • Phylogenetics. In pre-phylogenetic classifications, Aulopiformes and Myctophiformes were grouped together in Iniomi or a more inclusive Myctophiformes (e.g., Regan 1911a; Greenwood et al. 1966; Gosline 1971). Aulopiformes was first delimited as a monophyletic group in one of the earliest efforts to resolve the phylogenetic relationships of Euteleostei (Rosen 1973). Several phylogenetic analyses using morphological characters inferred paraphyly of Aulopiformes (R. K. Johnson 1982; Rosen 1985; Hartel and Stiassny 1986); however, subsequent morphological and molecular studies resolved the lineage as monophyletic (G. D. Johnson 1992; Baldwin and Johnson 1996; Sato and Nakabo 2002; Fielitz 2004; Davis 2010; Fielitz and González-Rodríguez 2010). The most comprehensive phylogeny of Aulopiformes is one resulting from analysis of combined morphological and molecular characters (Davis 2010). Incongruence in the phylogenies inferred from the combined morphological and molecular dataset and those based solely on morphological characters involve the relationships of Alepisaurus (lancetfishes and daggertooths), Bathysauropsis (black lizardfishes), Chlorophthalmidae (greeneyes), Evermannellidae (sabertooth fishes), Ipnopidae (deep-sea tripod fishes), Notosudidae (waryfishes), Paralepididae (barracudinas), and Sudis (Baldwin and Johnson 1996; Sato and Nakabo 2002; Davis 2010). Analysis of morphological characters resolves the phylogenetic relationships of fossil lineages of Aulopiformes that include †Argillichthys, †Apateodus, †Cimolichthys, †Enchodontoidei, †Holosteus, †Labrophagus, and †Pavlovichthys (Fielitz 2004; Davis and Fielitz 2010; Fielitz and González-Rodríguez 2010; H. M. A. Silva and Gallo 2011; Cavin et al. 2012; Marramà and Carnevale 2017; Beckett, Giles, et al. 2018; Díaz-Cruz et al. 2020, 2021).

  • Composition. There are 298 living species of Aulopiformes (Fricke et al. 2023), including Alepisaurus, Bathysauroides gigas (Pale Deepsea Lizardfish) and species classified in Bathysaurus, Bathysauropsis, Gigantura (telescopefishes), Paraulopus (cucumberfishes), Pseudotrichonotus (sand-diving lizardfishes), Sudis, Aulopidae (flagfins), Chlorophthalmidae, Evermannellidae, Ipnopidae, Notosudidae, Paralepididae, Scopelarchidae (pearleyes), and Synodontidae (lizardfishes). Fossil taxa of Aulopiformes include †Apateodus, †Cimolichthys, †Argillichthys, †Labrophagus, †Enchodontoidei, †Holosteus, and †Pavlovichthys (Fielitz and González-Rodríguez 2010; Marramà and Carnevale 2017; Beckett, Giles, et al. 2018; Díaz-Cruz et al. 2020, 2021). Details of the ages and locations of the fossil taxa are given in Appendix 1. Over the past 10 years 25 new species of Aulopiformes have been described (Fricke et al. 2023), comprising 8.4% of the species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies of Aulopiformes include (1) presence of elongate uncinate process on second epibranchial (Rosen 1973; Sato and Nakabo 2002; Davis 2010; Beckett, Giles, et al. 2018), (2) cartilaginous condyle on dorsal surface of third pharyngobranchial does not articulate with second epibranchial (G. D. Johnson 1992; Baldwin and Johnson 1996; Sato and Nakabo 2002; Davis 2010; Wiley and Johnson 2010), (3) fourth epibranchial with enlarged proximal end capped with a large band of cartilage and uncinate process at middle portion (Sato and Nakabo 2002; Davis 2010; Beckett, Giles, et al. 2018), (4) presence of fifth epibranchial (Baldwin and Johnson 1996; Sato and Nakabo 2002; Davis 2010; Beckett, Giles, et al. 2018), (5) ventral portion of palatine not expanded laterally (Sato and Nakabo 2002; Davis 2010), (6) posterior placement of the palatine cartilaginous facet for articulation with lateral ethmoid (Sato and Nakabo 2002; Davis 2010), (7) epipleurals extend anteriorly to at least second vertebrae (Patterson and Johnson 1995; Baldwin and Johnson 1996; Sato and Nakabo 2002; Davis 2010; Wiley and Johnson 2010), (8) one or more epipleurals displaced dorsally into horizontal septum (Patterson and Johnson 1995; Baldwin and Johnson 1996; Davis 2010; Wiley and Johnson 2010), (9) some ribs ossify in membrane bone (Baldwin and Johnson 1996; Davis 2010), (10) proximal portion of principal caudal-fin rays with modified segment (Baldwin and Johnson 1996; Sato and Nakabo 2002; Davis 2010), (11) medial process of pelvic girdle joined with cartilage (Baldwin and Johnson 1996; Sato and Nakabo 2002; Davis 2010; Wiley and Johnson 2010), (12) presence of two adductor profundus elements, (13) absence of swim bladder (R. K. Johnson 1982; Baldwin and Johnson 1996; Sato and Nakabo 2002; Davis 2010; Wiley and Johnson 2010), and (14) presence of head spines on larvae (Baldwin and Johnson 1996; Sato and Nakabo 2002; Davis 2010).

  • Synonyms. Scopeliformes (Gosline 1961:10–11), Cyclosquamata (Rosen 1973:509; Betancur-R et al. 2017:19), and Aulopa (Wiley and Johnson 2010:144) are ambiguous synonyms of Aulopiformes. Iniomi is a partial synonym of Aulopiformes (Gosline et al. 1966:1-2).

  • Comments. Aulopiformes has been consistently used as the group name for the clade outlined in the definition (Rosen 1973; W. L. Fink 1984a; Davis 2010; Wiley and Johnson 2010; Davis et al. 2016; J. S. Nelson et al. 2016:266–276) and is chosen as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • Eight of the 15 taxonomic families of Aulopiformes contain either a single species or a single genus (Davis 2010). Future efforts aimed at reducing group names in the phylogenetic-based classification of Aulopiformes could place Harpadon, Pseudotrichonotus, Saurida, Synodus, and Trachinocephalus into Aulopidae; Bathysaurus and Bathysauroides gigas into Giganturidae; and Bathysauropsis into Ipnopoidae.

  • The earliest fossil Aulopiformes is †Atolvorator longipectoralis from the Barremian (129.4–121.4 Ma) in Brazil (Gallo and Coelho 2008; Newbrey and Konishi 2015). The phylogenetic affinities of †Atolvorator within Aulopiformes are unresolved (Gallo and Coelho 2008). Bayesian relaxed molecular clock analyses of Aulopiformes result in an average posterior crown age estimate of 140 million years ago, with the credible interval ranging between 127 and 156 million years ago (Davis and Fielitz 2010).

  • img-z87-4_03.gif

    Ctenosquamata D. E. Rosen 1973
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Scopelengys tristis Alcock 1890 and Micropterus salmoides (Lacépède 1802). This is a minimum-crown-clade definition.

  • Etymology. Derived from the ancient Greek κτείς (t̍iːnIs), meaning comb, and the Latin squama, meaning scale.

  • Registration number. 929.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of nine concatenated Sanger-sequenced nuclear genes (Near, Eytan, et al. 2012, figs. 1, S1). Phylogenetic relationships of the major lineages of Ctenosquamata are presented in Figures 2 and 12. The placement of the fossil taxa †Ctenothrissiformes, †Sardinoides, and †Neocassandra are on the basis of analysis and inferences from morphological characters (M. Gaudant 1978b, 1979; Prokofiev 2006a; Dietze 2009; Davesne et al. 2016).

  • Phylogenetics. In a groundbreaking study of euteleost phylogeny on the basis of osteology and musculature of the jaws, pharyngobranchials, and caudal skeleton, Rosen (1973) introduced the group name Ctenosquamata for the clade containing Myctophiformes and Acanthomorpha. Rosen (1973) argued that Myctophiformes, comprising Myctophidae and Neoscopelidae, was more closely related to Acanthomorpha, in contrast to traditional classifications that grouped Aulopiformes with Myctophiformes (e.g., Regan 1911a; Jordan 1923:153–156; Berg 1940:437–438; Gosline et al. 1966; R. K. Johnson 1982). In a study of occipital anatomy, Rosen (1985) later rejected the monophyly of Ctenosquamata, proposing a phylogenyinwhichMyctophidae(lanternfishes) and Acanthomorpha share a common ancestry to the exclusion of Neoscopelidae (blackchins). Johnson (1992) convincingly pointed out problems in the interpretation of character variation in Rosen (1985) and reviewed evidence for the monophyly of Ctenosquamata.

  • The monophyly of Ctenosquamata is supported in phylogenetic analyses of discretely coded morphological characters (Stiassny 1996; Wiley et al. 1998; Dietze 2009). Manual cladistic solutions representing ctenosquamate monophyly (Lauder and Liem 1983; Stiassny 1986) and other summary phylogenies of ray-finned fishes and teleosts depict monophyly of Ctenosquamata (W. L. Fink and Weitzman 1982; Rosen 1982; W. L. Fink 1984a; G. J. Nelson 1989; G. D. Johnson 1992; C. D. Roberts 1993; Yamaguchi 2000; A. C. Gill and Mooi 2002; Springer and Johnson 2004). Phylogenetic analysis of morphological characters resolves the Late Cretaceous †Ctenothrissiformes and Acanthomorpha as sister lineages (Davesne et al. 2016; Cantalice et al. 2021), a result consistent with a pre-cladistic study (Patterson 1964). Studies by M. Gaudant (1978b, 1979) hypothesized that †Ctenothrissiformes are stem lineage ctenosquamates.

  • Molecular phylogenetic analyses consistently resolve Ctenosquamata as monophyletic (Wiley et al. 1998; Alfaro, Santini, et al. 2009; Davis 2010; Betancur-R, Broughton, et al. 2013; T. Grande et al. 2013; Poulsen et al. 2013; W.-J. Chen, Santini, et al. 2014; Davis et al. 2014, 2016; Denton 2014; Malmstrøm et al. 2016; W. L. Smith et al. 2016; Mirande 2017; Hughes et al. 2018; Martin et al. 2018; Mu et al. 2022; J.-F. Wang et al. 2023). In a maximum parsimony analysis of complete mtDNA genomic sequences, the ateleopid lineages Ateleopus and Ijimaia are nested in Ctenosquamata (Miya et al. 2001, 2003), but subsequent analyses using model based phylogenetic analysis of complete mtDNA genomic sequences result in ctenosquamate monophyly (Poulsen et al. 2013; J.-F. Wang et al. 2023).

  • Composition. Ctenosquamata includes more than 21,150 living species (Fricke et al. 2023) classified in Acanthomorpha and Myctophiformes. Fossil taxa of Ctenosquamata include the pan-acanthomorph †Ctenothrissiformes (Patterson 1964; M. Gaudant 1978b; Davesne et al. 2016) and the pan-myctophiforms †Sardinoides monasteri and †Neocassandra mica (Prokofiev 2006a; Dietze 2009). Details on the ages and locations of the fossil taxa are given in Appendix 1. Over the past 10 years there have been 1,681 new living species of Ctenosquamata described (Fricke et al. 2023), comprising 8.3% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Ctenosquamata include (1) absences of fifth upper pharyngeal tooth plates and associated third levatores interni (G. D. Johnson 1992; Olney et al. 1993; Stiassny 1996; Wiley and Johnson 2010), (2) two or fewer branchiostegal rays on posterior ceratohyal (McAllister 1968; Stiassny 1996; Wiley and Johnson 2010), (3) loss of craniotemporalis musculature (Stiassny 1986, 1996; Wiley and Johnson 2010), (4) absence of supraorbital bones (Stiassny 1996; Wiley and Johnson 2010), and (5) presence of single and medially fused neural arch on first vertebral centrum (Stiassny 1996; Wiley and Johnson 2010).

  • Synonyms. There are no synonyms of Ctenosquamata.

  • Comments. The earliest fossil Ctenosquamata includes several pan-lampriforms, pan-acanthopterygians, pan-holocentrids, pan-trachichthyiforms, pan-percomorphs, and †Ctenothrissa signifer from the Cenomanian (100.5–93.2 Ma) in the Cretaceous of Lebanon (Bannikov and Bacchia 2005; Davesne et al. 2016). Bayesian relaxed molecular clock analyses of Ctenosquamata result in an average posterior crown age estimate of 149.7 million years ago, with the credible interval ranging between 141.8 and 159.1 million years ago (Hughes et al. 2018).

  • img-z88-8_03.gif

    Myctophiformes C. T. Regan 1911a:121
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Neoscopelus macrolepidotus J. Y. Johnson 1863 and Myctophum punctatum Rafinesque 1810b. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek µυκτήρ (mˈuːktɚ), meaning nose, and όφίς (ˈo͡Ʊfiz), meaning snake. The suffix is from the Latin forma, meaning form, figure, or appearance.

  • Registration number. 930.

  • Reference phylogeny. A phylogeny inferred from a combined analysis of phylogenomic data (UCEs), Sanger-sequenced mtDNA and nuclear genes, and morphology (Martin et al. 2018, fig. 4). Phylogenetic relationships of living and fossil lineages of Myctophiformes are presented in Figure 12. The placements of fossil taxa in the phylogeny are on the basis of inferences from morphology (Prokofiev 2006a; Dietze 2009).

  • Phylogenetics. The first phylogenies of Myctophiformes inferred from morphological characters included nearly every genus (Paxton et al. 1984; Stiassny 1996), but did not include outgroups to test monophyly of the taxon. A phylogenetic analysis that sampled 60 morphological characters from taxa representing Acanthomorpha, Aulopiformes, Myctophidae (lanternfishes), Neoscopelidae (blackchins), and Stomiiformes resolves Myctophiformes as monophyletic (Dietze 2009). The monophyly of Myctophiformes is supported in morphological analyses (Stiassny 1986, 1996; Yamaguchi 2000). A combined analysis of morphology, mtDNA, Sanger-sequenced nuclear genes, and next-generation sequenced UCE loci strongly supports the monophyly of Myctophiformes (Martin et al. 2018). One analysis of partial mtDNA rRNA genes resolves Myctophiformes as paraphyletic (Colgan et al. 2000); however, all other phylogenetic analyses of molecular data result in myctophiform monophyly (e.g., Miya et al. 2001, 2003; Davis 2010; Near, Eytan, et al. 2012, 2013; Poulsen et al. 2013; W.-J. Chen, Santini, et al. 2014; Davis et al. 2014, 2016; Denton 2014; W. L. Smith et al. 2016; Martin et al. 2018).

  • Composition. Myctophiformes currently contains 258 living species (Paxton and Hulley 1999a, 1999b; Fricke et al. 2023), classified in Myctophidae and Neoscopelidae. Fossil lineages of Myctophiformes include the pan-neoscopelid †Beckerophotus and the pan-myctophid †Eomyctophum. Details of the ages and locations of the fossil taxa are given in Appendix 1. Over the past 10 years there was one new species of Myctophiformes described (Fricke et al. 2023), comprising 0.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Myctophiformes include (1) median dorsal keel present on mesethmoid (Stiassny 1986, 1996; Wiley and Johnson 2010), (2) median maxilla-premaxillary ligaments (VIII) insert on the contralateral buccal elements (Stiassny 1986, 1996; Wiley and Johnson 2010), (3) large tooth plate fused to proximal end of fourth ceratobranchial (Stiassny 1996; Wiley and Johnson 2010), (4) absence or reduction of first levator externus (Stiassny 1996; Wiley and Johnson 2010), (5) parapophyses on first vertebral centrum are cone-like and enlarged and meet at ventral midline (Stiassny 1986, 1996; Wiley and Johnson 2010), (6) adipose fin support inserted ventrally into supracarinalis posterior muscle mass (Stiassny 1996; Wiley and Johnson 2010), (7) presence of tranversus paryngobranchiales 2a and 2b (Springer and Johnson 2004; Wiley and Johnson 2010), (8) single fused extrascapular (Martin et al. 2018), and (9) narrow pubic plate (Martin et al. 2018).

  • Synonyms. Myctophata (Wiley and Johnson 2010:146; Betancur-R et al. 2017:19) and Scopelomorpha (Rosen and Patterson 1969:460; Rosen 1973:509; J. S. Nelson et al. 2016:276) are ambiguous synonyms of Myctophiformes. Myctophoidei (Greenwood et al. 1966:395) is a partial synonym of Myctophiformes.

  • Comments. Prior to Rosen's (1973) proposal limiting Myctophiformes to Myctophidae and Neoscopelidae, earlier classifications considered Myctophiformes or Iniomi to include Aulopiformes, Myctophidae, and Neoscopelidae (e.g., Regan 1911a; Greenwood et al. 1966; Gosline 1971). Scientists questioned the reality of Iniomi as early as the late 19th century (e.g., T. N. Gill 1893). In the first 10 years after Rosen (1973) some authors continued to recognize this heterogeneous concept of Myctophiformes (R. K. Johnson 1982; Okiyama 1984). Convincing evidence for the delimitation of Myctophiformes followed here came from detailed and thorough morphological analyses (Stiassny 1986, 1996). Essentially all molecular analyses have supported the monophyly of Myctophiformes (e.g., Near, Eytan, et al. 2012; Poulsen et al. 2013; Davis et al. 2014), demonstrating that as much as molecular phylogenies dramatically affect teleost classifications, they are also corroborative for well-supported but contentious hypotheses proposed as a result of analysis of morphological data. The name Myctophiformes was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest fossil Myctophiformes are two pan-myctophids: the otolith taxon †Eokrefftia prediaphus from the Thanetian (59.2–56.0 Ma) of South Australia and the skeletal taxon †Eomyctophum broncus from the Ypresian (56.0–48.1 Ma) of New Zealand (Schwarzhans 2019; Schwarzhans and Carnevale 2021). Bayesian relaxed molecular clock analyses of Myctophiformes result in an average posterior crown age estimate of 69.0 million years ago, with the credible interval ranging between 60.1 and 78.7 million years ago (Near et al. 2013).

  • img-z90-3_03.gif

    Acanthomorpha D. E. Rosen 1973:510
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Lampris guttatus (Brünnich 1788), Polymixia lowei Günther 1859, Percopsis omiscomaycus (Walbaum 1792), Zeus faber Linnaeus 1758, Stylephorus chordatus Shaw 1791, Gadus morhua Linnaeus 1758, Diretmus argenteus Johnson 1864, Beryx decadactylus Cuvier 1829 in Cuvier and Valenciennes (1829a), Carapus bermudensis (Jones 1874), and Micropterus salmoides (Lacépède 1802). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek μκανθα (æk̍ænθә), meaning thorn or spine, and µoρϕή (m̍ͻ͡ʊfiz), meaning form or shape.

  • Registration number. 931.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, figs. S1–S25). Phylogenetic relationships among the major living and fossil lineages of Acanthomorpha are presented in Figures 2 and 12. Phylogenetic placements of fossil taxa are on the basis of inferences from morphological analyses (Davesne et al. 2014, 2016; Delbarre et al. 2016; Cantalice et al. 2021).

  • Phylogenetics. Phylogenetic analyses of discretely coded morphological character state changes resolve Acanthomorpha as monophyletic (Stiassny 1986; Stiassny and Moore 1992; G. D. Johnson and Patterson 1993; Davesne et al. 2016; Cantalice et al. 2021). Relationships within acanthomorphs differ among morphological analyses, but several studies resolve Lampriformes as the sister lineage to all other acanthomorphs and Holocentridae as the sister lineage of Percomorpha (Stiassny and Moore 1992; Olney et al. 1993; Davesne et al. 2016; Cantalice et al. 2021).

  • Molecular phylogenetic analyses using mtDNA, nuclear genes, or combinations of the two and phylogenomic analyses consistently resolve Acanthomorpha as monophyletic (W.-J. Chen et al. 2003; Miya et al. 2003, 2005; W. L. Smith and Wheeler 2006; Alfaro, Santini, et al. 2009; Santini et al. 2009; Davis 2010; Near, Eytan, et al. 2012, 2013; Betancur-R, Broughton, et al. 2013; T. Grande et al. 2013; W.-J. Chen, Santini, et al. 2014; Davis et al. 2016; Malmstrøm et al. 2016; W. L. Smith et al. 2016; Betancur-R et al. 2017; Hughes et al. 2018; Musilova et al. 2019; Roth et al. 2020; Ghezelayagh et al. 2022; Mu et al. 2022). However, relationships among the major lineages of Acanthomorpha vary across studies. Phylogenomic analyses consistently resolve three major clades: Lampriformes, Paracanthopterygii, and Acanthopterygii (Alfaro et al. 2018; Hughes et al. 2018; Musilova et al. 2019; Ghezelayagh et al. 2022). Trees inferred from phylogenomic analyses of exons resolve Lampriformes as the sister lineage of Acanthopterygii (Hughes et al. 2018; Musilova et al. 2019; Roth et al. 2020), while phylogenomic analyses of UCE loci and a set of 82 exons place Lampriformes as the sister lineage of Paracanthopterygii (Alfaro et al. 2018; Ghezelayagh et al. 2022; Mu et al. 2022). Bayesian concordance factors estimated using UCE data find that resolution of Lampriformes as sister of Paracanthopterygii is supported by the greatest proportion of sampled loci; however, the 95% highest posterior density of the concordance factors overlaps with that of the phylogeny that resolves Lampriformes and Acanthopterygii as sister lineages (Ghezelayagh et al. 2022), suggesting relationships among lineages of Acanthomorpha are not confidently resolved.

  • Composition. Acanthomorpha currently includes more than 19,895 living species (Fricke et al. 2023), classified in the subclades Lampriformes, Paracanthopterygii, and Acanthopterygii. Fossil lineages include the pan-lampriforms †Aipichthys, †Aipichthyoides, †Nardovelifer, and †Zoqueichthys (Patterson 1964; Alvarado-Ortega and Than-Marchese 2012; Murray and Wilson 2014; Davesne et al. 2016; Delbarre et al. 2016; Cantalice et al. 2021); the pan-paracanthopterygian †Pycnosteroides (Patterson 1964, 1993; Davesne et al. 2016; Cantalice et al. 2021); and the pan-acanthopterygian †Choichix (Cantalice et al. 2021). Details of the ages and locations of the fossil taxa are given in Appendix 1. Over the past 10 years 1,680 new species of Acanthomorpha have been described (Fricke et al. 2023), comprising 8.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Acanthomorpha include (1) anterior facets on the first vertebral centrum that articulate with the exoccipital condyles (Rosen 1985; G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), (2) maxillo-rostroid ligament originates from inner portion of maxillary median process and inserts onto rostral cartilage (Stiassny 1986; Olney et al. 1993; Wiley and Johnson 2010), (3) spina occipitalis extends ventrally, forming dorsal margin of the foramen magnum (Stiassny 1986; Olney et al. 1993), (4) anterior extension of lateral ethmoid located close to, or sutured with, lateral process projecting from ventral stalk of vomer (Stiassny 1986; Olney et al. 1993; Davesne et al. 2016; Cantalice et al. 2021), (5) upper limb of posttemporal bound to epioccipital with a reduced posttemporal-epioccipital ligament (Stiassny 1986; Olney et al. 1993), (6) distal ossification of medial pelvic process (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), (7) separated medial and anterior infracarinales muscles (G. D. Johnson and Patterson 1993; Stiassny 1993; Wiley and Johnson 2010), (8) presence of unsegmented, bilaterally fused dorsal and anal fin spines (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010; Davesne et al. 2016; Cantalice et al. 2021), (9) absence of median caudal cartilages (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), and (10) antorbital bone absent (Cantalice et al. 2021).

  • Synonyms. Acanthomorphata (Wiley and Johnson 2010:127, 146–147; Betancur-R et al. 2017:20) is an ambiguous synonym of Acanthomorpha.

  • Comments. Acanthomorpha, or spiny-rayed fishes, comprise one of the major inclusive lineages of teleost fishes and the name Acanthomorpha is here defined as applying to the clade originating in their most recent common ancestor. Since the recognition and delimitation of Acanthomorpha by Rosen (1973), the major living lineages that comprise this taxon have not changed. The discovery of support for acanthomorph monophyly and the phylogenetic relationships of its constituent lineages (Stiassny 1986; G. D. Johnson and Patterson 1993; Near, Eytan, et al. 2012) remains an active area of research. Recent Sanger sequencing and phylogenomic studies provide unprecedented taxon sampling and resolution for acanthomorph phylogenetic relationships (Betancur-R, Broughton, et al. 2013; Near et al. 2013; Alfaro et al. 2018; Ghezelayagh et al. 2022).

  • The earliest fossils of Acanthomorpha date to the Cenomanian (100.5–93.9 Ma) (Patterson 1993; Friedman 2010; Murray 2016). Bayesian relaxed molecular clock analyses of Acanthomorpha result in an average posterior crown age estimate of 144.8 million years ago, with the credible interval ranging between 136.9 and 152.3 million years ago (Ghezelayagh et al. 2022).

  • img-z91-8_03.gif

    Lampriformes G. C. Steyskal 1980:171
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Lampris guttatus (Brünnich 1788), Metavelifer multiradiatus (Regan 1907a), and Regalecus russelii (Cuvier 1816), but not Stylephorus chordatus Shaw 1791. This is a minimum-crown-clade definition with an external specifier.

  • Etymology. From the ancient Greek λαµπῥoς (l̍æmpɹo͡Ʊz) meaning bright, brilliant, or radiant. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 932.

  • Reference phylogeny. A phylogeny of Lampriformes inferred from analysis of seven Sanger-sequenced nuclear genes (Brownstein and Near 2023, fig. 4). Phylogenetic relationships of the living and fossil lineages of Lampriformes are presented in Figure 12. The placements of the fossil lampriform taxa in the phylogeny are on the basis of inferences from morphology (Bannikov 1999; Gottfried et al. 2006; Brownstein and Near 2023).

  • Phylogenetics. The phylogenetic relationships within Lampriformes have been investigated with analyses of morphological and molecular datasets (Oelschläger 1983; Olney et al. 1993; Wiley et al. 1998; T. R. Roberts 2012; Martin 2015). The morphological phylogenies presented in Olney et al. (1993) and Martin (2015), and the molecular phylogeny in Wiley et al. (1998) are congruent in the resolution of Veliferidae (velifers) as the sister lineage of all other Lampriformes and Lampris (opahs) as the sister lineage of a clade containing Lophotidae (crest-fishes), Radiicephalus (tapertails), Regalecidae (oarfishes), and Trachipteridae (ribbonfishes). A molecular phylogeny inferred from mtDNA and nuclear genes resolves a clade containing Lampris and Veliferidae that is the sister lineage of all other Lampriformes (Rabosky et al. 2018; J. Chang et al. 2019), which is consistent with the classification that grouped Lampris and Veliferidae in Bathysomi and all other lineages of lampriforms in Taeniosomi (Regan 1907b). A series of morphological phylogenetic analyses that included multiple species of Lampriformes were aimed at investigating the relationships of several pan-lampriform fossil taxa (Davesne et al. 2014, 2016; Delbarre et al. 2016; Cantalice et al. 2021).

  • The two earliest morphological phylogenetic analyses of Oelschläger (1983) and Olney et al. (1993) include Stylephorus chordatus as this species was long classified with lineages of Lampriformes (Günther 1861:306; Regan 1908, 1924; Starks 1908; Goodrich 1909:475–477; Jordan 1923; Greenwood et al. 1966; McAllister 1968; J. S. Nelson 2006). Molecular phylogenetic analyses consistently resolve Lampriformes as monophyletic to the exclusion of Stylephorus, which is resolved as the sister lineage of all other Gadiformes (Miya et al. 2007; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Alfaro et al. 2018; Hughes et al. 2018; Ghezelayagh et al. 2022). Morphological phylogenetic analyses aimed at relationships among lineages of Acanthomorpha are congruent with molecular phylogenies in resolving Lampriformes and Stylephorus as distantly related (Davesne et al. 2016).

  • Composition. There are currently 30 living species of Lampriformes (Fricke et al. 2023) classified in Lampris, Lophotidae, Radiicephalus, Regalecidae, and Veliferidae. Fossil taxa of Lampriformes include the species of VeliferidaeVeronavelifer sorbini; the pan-veliferids †Palaeocentrotus boeggildi, †Turkmene finitimus, and †Danatinia casca (Bannikov 1990, 1999, 2014a); the species of LophotidaeBabelichthys olneyi (Davesne 2017); the pan-lophotids †Protolophotus elami, †Eolophotes lenis, and †Oligolophotes fragosus (Walters 1957; Bannikov 1999; Davesne 2017); and the pan-lamprid †Megalampris keyesi (Gottfried et al. 2006). Details of the ages and locations for the fossil taxa are given in Appendix 1. In the last 10 years, four new living species of Lampriformes have been described (Underkoffler et al. 2018; Koeda and Ho 2019; Fricke et al. 2023), comprising 13.3% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Lampriformes include (1) second ural centrum free from fused first ural and preural centra and fused posteriorly to upper hypural plate (Patterson 1968; Wiley and Johnson 2010; Davesne et al. 2014, 2016; Delbarre et al. 2016; Cantalice et al. 2021), (2) anterior palatine process and anterior palatomaxillary ligament absent (Olney et al. 1993; Wiley and Johnson 2010; Davesne et al. 2014), (3) mesethmoid posterior to lateral ethmoids (Olney et al. 1993; Wiley and Johnson 2010), (4) elongate ascending processes of premaxillae and large rostral cartilage insert into frontal vault or cradle (Olney et al. 1993; Wiley and Johnson 2010; Davesne et al. 2014, 2016; Cantalice et al. 2021), (5) first dorsal fin pterygiophore inserts anterior to first neural spine (Olney et al. 1993; Wiley and Johnson 2010; Davesne et al. 2014, 2016; Cantalice et al. 2021), (6) postcleithrum composed of a single bone (Otero and Gayet 1996; Davesne et al. 2014, 2016; Delbarre et al. 2016; Cantalice et al. 2021), (7) premaxillary free of dentition (Delbarre et al. 2016), (8) dentary free of dentition (Delbarre et al. 2016), (9) endopterygoid free of dentition (Delbarre et al. 2016), and (10) condylar articulation between anterior ceratohyal and ventral hypohyal (Davesne et al. 2016; Cantalice et al. 2021).

  • Synonyms. Allotriognathi is an ambiguous (Regan 1907b:638–640; Garstang 1931:259) and a partial (Jordan 1923:165–166) synonym of Lampriformes. Atelaxia (Starks 1908:1) is a partial synonym of Lampriformes. Lampridiformes (Goodrich 1909:475–477; Walters and Fitch 1960:442; Greenwood et al. 1966:398; McAllister 1968:106–108; J. S. Nelson 1976:179–180; Lauder and Liem 1983:166; Olney et al. 1993:137; Springer and Johnson 2004:80–81; Wiley and Johnson 2010:127, 147), Lampridacea (Wiley and Johnson 2010:127, 147), Lamprimorpha (J. S. Nelson et al. 2016:280), and Lampripterygii (Betancur-R et al. 2017:20) are ambiguous synonyms of Lampriformes.

  • Comments. Lampriformes is the group name most frequently applied to the clade as defined here in several classifications of acanthomorphs (Davis et al. 2016; Betancur-R et al. 2017; Dornburg and Near 2021; Ghezelayagh et al. 2022).

  • The earliest fossil taxa of Lampriformes include †Danatinia casca and †Turkmene finitimus from the Ypresian (56.0–48.1 Ma) of Turkmenistan (Bannikov 1999). Bayesian relaxed molecular clock analyses of Lampriformes result in an average posterior crown age estimate of 58.1 million years ago, with the credible interval ranging between 55.8 and 69.7 million years ago (Ghezelayagh et al. 2022).

  • img-z93-6_03.gif

    Paracanthopterygii P. H. Greenwood, D. E. Rosen, S. H. Weitzman, and G. S. Myers 1966:352, 396–397
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Percopsis omiscomaycus (Walbaum 1792) (Percopsiformes) and Gadus morhua Linnaeus 1758 (Gadiformes). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek παρά (p̍æɹә) meaning beside, ἇκανθα (æk̍ænθә) meaning thorn or spine, and πτερὀν (t̍εɹαːn) meaning fin or wing.

  • Registration number. 933.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, fig. S1). Phylogenetic relationships of the living lineages and fossil taxa of Paracanthopterygii are presented in Figure 13. Placements of the fossil taxa in the phylogeny are on the basis of inferences from morphology (Tyler and Santini 2005; Alvarado-Ortega and Than-Marchese 2012; Murray and Wilson 2014; Davesne et al. 2016, 2017; Cantalice et al. 2021; Schrøder et al. 2022).

  • Phylogenetics. Paracanthopterygii was first delimited as a named group in Greenwood et al. (1966). Among teleosts there is no other taxonomic group that has had a more fluid history of hypotheses aimed at its composition (Rosen and Patterson 1969; Patterson and Rosen 1989; A. C. Gill 1996; T. Grande et al. 2013). The varied delimitations of Paracanthopterygii before the advent of molecular phylogenetics included Myctophiformes (Fraser 1972) and the percomorphs Ophidiiformes, Batrachoididae, Gobiesocoidei, Lophioidei, and Zoarcoidei (Rosen and Patterson 1969; Fraser 1972; Lauder and Liem 1982; Patterson and Rosen 1989). All of the premolecular delimitations of Paracanthopterygii excluded Zeiformes because they were considered a lineage of Acanthopterygii (Greenwood et al. 1966; Rosen 1984; Patterson and Rosen 1989; G. D. Johnson and Patterson 1993), despite inferences from morphology that argued for common ancestry of zeiforms and paracanthopterygians (M. Gaudant 1979; Gayet 1980b).

  • The delimitation of Paracanthopterygii that includes Gadiformes, Percopsiformes, Polymixia, and Zeiformes was first proposed as a result of phylogenetic analyses of whole mtDNA genomes (Miya et al. 2003, 2005), and supported in subsequent molecular studies (W. L. Smith and Wheeler 2006; B. Li et al. 2009; T. Grande et al. 2013; W.-J. Chen Santini, et al. 2014; Malmstrøm et al. 2016; Alfaro et al. 2018; Hughes et al. 2018; Musilova et al. 2019; Roth et al. 2020; Roa-Varón et al. 2021; Ghezelayagh et al. 2022; Mu et al. 2022; J.-F. Wang et al. 2023) as well as a phylogenetic analysis of discretely coded morphological characters (Davesne et al. 2016). A number of molecular phylogenetic analyses do not resolve Paracanthopterygii as monophyletic (Wiley et al. 2000; Holcroft 2004; Sparks et al. 2005; Dettaï and Lecointre 2008; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Betancur-R et al. 2017; Near et al. 2013; Davis et al. 2016; Smith et al. 2016), but these studies are either on the basis of relatively small DNA sequence datasets or result in phylogenies with low support at nodes reflecting paracanthopterygian paraphyly.

  • Long classified in Lampriformes (Olney et al. 1993), Stylephorus chordatus is consistently resolved in molecular phylogenies as nested within Paracanthopterygii as the sister lineage of all other Gadiformes (Miya et al. 2007; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Malmstrøm et al. 2016, 2017; Alfaro et al. 2018; T. Grande et al. 2018; Hughes et al. 2018; Musilova et al. 2019; Roth et al. 2020; Ghezelayagh et al. 2022; J.-F. Wang et al. 2023). Within Paracanthopterygii, the results of phylogenetic analyses differ, with molecular and morphological studies resolving Polymixia and Percopsiformes as sister lineages (W.-J. Chen et al. 2003; Miya et al. 2005, 2007; W. L. Smith and Wheeler 2006; Dillman et al. 2011; Alvarado-Ortega and Than-Marchese 2012; Murray and Wilson 2014; Malmstrøm et al. 2016; Alfaro et al. 2018; Musilova et al. 2019; Roth et al. 2020; Cantalice et al. 2021; Ghezelayagh et al. 2022; J.-F. Wang et al. 2023), but other molecular studies resolving Polymixia as the sister lineage of all other paracanthopterygians (T. Grande et al. 2013; W.-J. Chen, Santini, et al. 2014; Hughes et al. 2018; Roa-Varón et al. 2021). Bayesian concordance factors estimated using UCE data find the hypothesis that Polymixia and Percopsiformes are sister lineages is supported by the greatest proportion of sampled loci, and the phylogeny that depicts Polymixia as the sister lineage of all other Paracanthopterygii is identified as less optimal (Ghezelayagh et al. 2022). Phylogenetic analyses of morphological datasets provide resolution for several fossil lineages of Paracanthopterygii (Murray and Wilson 1999; Tyler and Santini 2005; Alvarado-Ortega and Than-Marchese 2012; Murray and Wilson 2014; Davesne et al. 2016, 2017; Cantalice et al. 2021; Schrøder et al. 2022).

  • Composition. Paracanthopterygii currently includes 681 species (Fricke et al. 2023), classified in Gadiformes, Percopsiformes, Polymixia, and Zeiformes. Fossil lineages include the pan-polymixiid †Polyspinatus (Schrøder et al. 2022), the pan-percopsiforms †Sphenocephalidae and †Omosomopsis (Patterson 1964; M. Gaudant 1978a; Murray and Wilson 1999; Newbrey et al. 2013; Davesne et al. 2016; Cantalice et al. 2021), and the pan-zeiforms †Archaeozeus, †Bajaichthys, and †Protozeus (Tyler and Santini 2005; Davesne et al. 2017). Details of the ages and locations for the fossil taxa are given in Appendix 1. Over the past 10 years, 35 new species of Paracanthopterygii have been described (Fricke et al. 2023), comprising 5.1% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Paracanthopterygii include (1) presence of a full-length spine on dorsal surface of preural centrum 2 (Borden et al. 2013; T. Grande et al. 2013), (2) insertion sites of interradials on principal caudal and other rays (Borden et al. 2013), (3) first dorsal pterygiophore inserts posterior to neural spine 4 (Davesne et al. 2016; Cantalice et al. 2021), (4) no contact of pelvic girdle posterior to pectoral girdle (Davesne et al. 2016; Cantalice et al. 2021), and (5) base of pelvic fin spine asymmetrical (Cantalice et al. 2021).

  • Synonyms. Paracanthomorphacea (Betancur-R, Broughton, et al. 2013:12–13) is a partial synonym of Paracanthopterygii.

  • Comments. A consistent delimitation of Paracanthopterygii that includes Gadiformes, Percopsiformes, Polymixia, and Zeiformes in morphological and molecular studies is an important development in the resolution of phylogenetic relationships within Acanthomorpha. Remaining issues in the phylogenetics of Paracanthopterygii include the relationships of Percopsiformes and Polymixia and the resolution of the Cretaceous fossil lineages †Berycopsis, †Berycopsia, †Dalmatichthys, †Homonotichthys, and †Omosoma long classified with Polymixia in Polymixiiformes (Patterson 1964, 1993; Radovcic 1975; Bannikov and Bacchia 2005; Murray and Cumbaa 2013; Newbrey et al. 2013; Friedman et al. 2016).

  • The earliest fossils of Paracanthopterygii all date to the Cenomanian (100.5–93.9 Ma), including the species of †Sphenocephalidae, †Xenyllion zonensis from Canada, and the pan-percopsiform †Omosomopsis simum from Morocco (Otero and Gayet 1995; M. V. H. Wilson and Murray 1996; Newbrey et al. 2013; Murray and Wilson 2014; Davesne et al. 2016; Cantalice et al. 2021). Bayesian relaxed molecular clock analyses of Paracanthopterygii result in an average posterior crown age estimate of 120.7 million years ago, with the credible interval ranging between 101.9 and 135.0 million years ago (Ghezelayagh et al. 2022).

  • img-z96-5_03.gif

    FIGURE 13.

    Phylogenetic relationships of the major living lineages and fossil taxa of Paracanthopterygii, Percopsiformes, Zeiformes, Gadiformes, and Gadoidei. Filled circles identify the common ancestor of clades with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z94-1_03.jpg

    Percopsiformes L. S. Berg 1937:1279
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Percopsis omiscomaycus (Walbaum 1792), Aphredoderus sayanus (Gilliams 1824), and Chologaster cornuta Agassiz 1853. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek πέρκη (pˈke͡I) meaning perch, specifically the freshwater European Perch, Perca fluviatilis or the marine Painted Comber, Serranus scriba (D. W. Thompson 1947:194–197), and ὂΨις (ˈαːpsIs) meaning a vision or apparition. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 934.

  • Reference phylogeny. A phylogeny inferred using DNA sequences from one mtDNA gene and nine nuclear genes (Niemiller et al. 2013, fig. 1). Phylogenetic relationships of the living and fossil lineages of Percopsiformes are presented in Figure 13. The placements of fossil taxa in the phylogeny are on the basis of several phylogenetic studies (Murray and Wilson 1999; Borden et al. 2013; T. Grande et al. 2013; Guinot and Cavin 2018; Murray et al. 2020).

  • Phylogenetics. The delimitation of Percopsiformes presented here is consistent with several pre-Hennigian phylogenetic studies based on morphology (Rosen 1962; Gosline 1963a; Greenwood et al. 1966; McAllister 1968; Rosen and Patterson 1969). Phylogenetic analyses of morphological characters are incongruent, with some studies not supporting the monophyly of Percopsiformes (Rosen 1985; Patterson and Rosen 1989; Murray and Wilson 1999; Murray et al. 2020), but other analyses resolving Percopsiformes as a clade (Springer and Orrell 2004; Davesne et al. 2016; Cantalice et al. 2021). In contrast to the lack of agreement among morphological studies, molecular phylogenetic analyses consistently resolve Percopsiformes as monophyletic with Percopsis (troutperches) as the sister lineage of a clade containing Amblyopsidae (cavefishes) and Aphredoderus sayanus (Pirate Perch) (W. L. Smith and Wheeler 2006; Dillman et al. 2011; Near, Eytan, et al. 2012; T. Grande et al. 2013; Near et al. 2013; Davis et al. 2016; Smith et al. 2016; Betancur-R et al. 2017; T. Grande et al. 2018; Ghezelayagh et al. 2022). The presumed monophyly of Percopsiformes was the basis for the selection of Percopsis and Aphredoderus sayanus as the sole outgroups in morphological and molecular phylogenetic analyses of Amblyopsidae (Niemiller et al. 2013; Armbruster et al. 2016; Hart et al. 2020).

  • Composition. There are currently 13 living species of Percopsiformes that include Aphredoderus sayanus and species classified in Percopsis and Amblyopsidae (Poly 2004a, 2004b; Poly and Proudlove 2004; Fricke et al. 2023). Fossil taxa of Percopsiformes include †Lindoeichthys albertensis from the Maastrichtian Scollard Formation, Canada (Murray et al. 2020), †Mcconichthys longipinnis from the Danian Tullock Member, USA (L. Grande 1988), †Amphiplaga brachyptera and †Erismatopterus levatus from the Ypresian Green River Formation, USA (Cope 1871c, 1877a; L. Grande 1984), †Libotonius blakeburnensis from the Ypresian Blakeburn Mine, Canada (M. V. H. Wilson 1977), †Lateopisciculus turrifumosus and †Massamorichthys wilsoni from the Selandian–Thanetian Paskapoo Formation, Canada (Murray 1996; Murray and Wilson 1996), and †Tricophanes foliarum from the Priabonian Florissant, USA (Cope 1878; Meyer 2003:179). Details of the ages and locations for the fossil taxa are given in Appendix 1. In the last 10 years, a single new living species of Percopsiformes has been described (Chakrabarty et al. 2014; Fricke et al. 2023), comprising 8.3% of the living species diversity in the clade. Analyses of morphology and genomic data indicate that there are three additional species of Aphredoderus awaiting formal taxonomic description (Muller 2023).

  • Diagnostic apomorphies. Morphological apomorphies for Percopsiformes include (1) absence of postmaxillary process on premaxilla (Patterson and Rosen 1989; Murray and Wilson 1999; Davesne et al. 2016; Cantalice et al. 2021), (2) six branchiostegal rays (Murray and Wilson 1999), (3) presence of opercular dorsal projection that is anteriorly truncated or excavated (Murray and Wilson 1999), (4) transverses dorsales and obliqui dorsalis are combined and have a trapezoidal shape in dorsal view (Springer and Johnson 2004; Wiley and Johnson 2010), (5) obliquus dorsalis 4 extends posteriorly to insert on levator process of epibranchial 4 (Springer and Johnson 2004; Wiley and Johnson 2010), (6) two hypural plates do not contact any ural centra (Borden et al. 2013), (7) presence of a two-headed cranio-hyomandibular articulation (Davesne et al. 2016; Cantalice et al. 2021), (8) posterior and anterior ceratohyals sutured (Davesne et al. 2016; Cantalice et al. 2021), (9) metapterygoid contacts quadrate (Cantalice et al. 2021), and (10) condylar articulation between the anterior ceratohyal and ventral hypohyal (Cantalice et al. 2021).

  • Synonyms. Percopsacea (Wiley and Johnson 2010:147) and Percopsaria (Betancur-R et al. 2017:20) are ambiguous synonyms of Percopsiformes. Salmopercae (Goodrich 1909:425–426; Regan 1909a:79, 84–85, 1911b:294, 1929:305, 318) is a partial synonym of Percopsiformes.

  • Comments. Percopsiformes is the group name consistently applied to the clade as defined here (Rosen and Patterson 1969; Wiley and Johnson 2010; Davis et al. 2016; J. S. Nelson et al. 2016:287–289; Betancur-R et al. 2017; Dornburg and Near 2021; Ghezelayagh et al. 2022).

  • The earliest fossil Percopsiformes is †Lindoeichthys albertensis from Canada (Murray et al. 2020). Bayesian relaxed molecular clock analyses of Percopsiformes result in an average posterior crown age estimate of 53.5 million years ago, with the credible interval ranging between 40.1 and 72.4 million years ago (Ghezelayagh et al. 2022).

  • img-z97-8_03.gif

    Zeiformes L. S. Berg 1937:1279
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Zeus faber Linnaeus 1758, Cyttus australis (Richardson 1843), Cyttopsis rosea (Lowe 1843), and Macrurocyttus acanthopodus (Fowler 1934). This is a minimum-crown-clade definition.

  • Etymology. Zeus is the god of thunder and the sky in ancient Greek religion. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 940.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of three mtDNA regions and five nuclear genes (T. Grande et al. 2018, fig. 3). Phylogenetic relationships of the living lineages and fossil taxa of Zeiformes are presented in Figure 13. Placements of the fossil taxa in the phylogeny are on the basis of analyses of morphological characters (Tyler and Santini 2005; Davesne et al. 2017).

  • Phylogenetics. Prior to molecular phylogenetic analyses, Zeiformes was classified as a lineage of Acanthopterygii (Greenwood et al. 1966; Rosen 1984; G. D. Johnson and Patterson 1993). Molecular analyses consistently resolve Zeiformes and Gadiformes as sister lineages (Wiley et al. 2000; W.-J. Chen et al. 2003; Miya et al. 2003, 2005, 2007; Sparks et al. 2005; W. L. Smith and Wheeler 2006; Dettaï and Lecointre 2008; B. Li et al. 2009; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; T. Grande et al. 2013; Near et al. 2013; W.-J. Chen, Santini, et al. 2014; Davis et al. 2016; Malmstrøm et al. 2016; Smith et al. 2016; Betancur-R et al. 2017; Alfaro et al. 2018; Hughes et al. 2018; Roth et al. 2020; Roa-Varón et al. 2021; Ghezelayagh et al. 2022; Mu et al. 2022).

  • There are two sets of phylogenetic analyses of Zeiformes based on morphological characters that result in very different phylogenetic trees. A group of two morphological phylogenies work from the premise that Zeiformes are acanthopterygians and consequently use species of Beryciformes, Trachichthyiformes, Antigonia, Tetraodontoidei, Moronidae, and Capros aper as outgroups (Tyler et al. 2003; Tyler and Santini 2005). The other morphological phylogeny follows the inferences stemming from molecular phylogenetic analyses and uses species of Polymixia, Percopsiformes, and Gadiformes as outgroup taxa (T. Grande et al. 2018). Both sets of phylogenetic analyses resolve most of the major lineages (e.g., Cyttus [lookdown dories], Oreosomatidae [oreos], Parazenidae [smooth dories], Zeidae [dories], and Zeniontidae [armoreye dories]) of Zeiformes as monophyletic (Tyler et al. 2003; Tyler and Santini 2005; T. Grande et al. 2018), but the resolution of Macrurocyttus acanthopodus (Dwarf Dory) renders Grammicolepididae (tinselfishes) as paraphyletic in one of the studies (T. Grande et al. 2018). The two morphological phylogenies are completely incongruent with regard to the relationships among the major lineages of Zeiformes (Tyler et al. 2003; Tyler and Santini 2005; T. Grande et al. 2018), perhaps a result of using acanthopterygian rather than paracanthopterygian outgroups (T. Grande et al. 2018).

  • Relationships within Zeiformes inferred from a molecular phylogenetic analysis are not congruent with either of the morphological inferred phylogenies, but are more similar to the trees resulting from analyses using Paracanthopterygii as outgroups (T. Grande et al. 2018). In the molecular phylogeny, Zeidae is the sister lineage of all other Zeiformes, Zeniontidae is paraphyletic because Capromimus is resolved as the sister lineage of Oreosomatidae, and Grammicolepididae, Parazenidae, and Zenion are resolved as monophyletic. Relationships among the lineages of Zeiformes resolved in the molecular phylogeny are not strongly supported and are reasonably interpreted as a polytomy near the inferred common ancestor of the clade, which is indicative of a period of rapid lineage diversification early in the evolutionary history of Zeiformes (T. Grande et al. 2018). The enigmatic and infrequently encountered Macrurocyttus is not sampled in any molecular phylogeny, and analyses of combined molecular and morphological datasets resolve this lineage as a deeply branching sister lineage of all other Zeiformes (T. Grande et al. 2018).

  • Composition. There are currently 33 living species of Zeiformes classified in Cyttus, Grammicolepididae, Oreosomatidae, Parazenidae, Zeidae, and Zeniontidae (Fricke et al. 2023). Fossil taxa of Zeiformes include the pan-parazenid †Cretazeus and several species from the Oligocene and Miocene classified as Zeus and Zenopsis (Tyler et al. 2000, 2003; Tyler and Santini 2005; Santini et al. 2006). Details of the ages and locations for the fossil taxa are given in Appendix 1. In the last 10 years, a single new living species of Zeiformes has been described (Kai and Tashiro 2019; Fricke et al. 2023), comprising 3.0% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Zeiformes include (1) distal portions of proximal-middle dorsal fin radials laterally expanded (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), (2) distal radials of spinous portion of dorsal fin absent or reduced to minuscule cartilaginous or incompletely ossified elements (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), (3) palatine has a mobile articulation with ectopterygoid that is dorsally truncated (G. D. Johnson and Patterson 1993; Tyler et al. 2003; Wiley and Johnson 2010; T. Grande et al. 2018), (4) reduced metapterygoid (G. D. Johnson and Patterson 1993; Tyler et al. 2003; Wiley and Johnson 2010; T. Grande et al. 2013, 2018), (5) flexible articulations on anterior vertebral centra; if ribs are present they are never anterior to fourth vertebra (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), (6) pharyngobranchials 2 and 3 with upright columnar processes (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), (7) absence of pharyngobranchial 4 and upper pharyngeal tooth plate (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010; T. Grande et al. 2018), (8) area below frontals from ethmoid cartilage to parasphenoid with a continuous medial cartilage (G. D. Johnson and Patterson 1993; Tyler et al. 2003; Wiley and Johnson 2010), (9) second pleural centrum with full neural spine (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), (10) proximally truncated parhypural (G. D. Johnson and Patterson 1993; Tyler et al. 2003; Wiley and Johnson 2010), (11) presence of 3.5 gills and seven hemibranchs (Tyler et al. 2003), (12) dorsal-, anal-, and pectoral-fin rays unbranched (Tyler et al. 2003; Wiley and Johnson 2010), (13) absence of uncinate process on epibranchial 1 (Tyler et al. 2003; T. Grande et al. 2018), (14) absence of open gill slit between branchial arches 4 and 5 (Tyler et al. 2003; T. Grande et al. 2018), (15) fusion of hypurals 1–2 and 3–4; both elements fused to centrum (Tyler et al. 2003; Wiley and Johnson 2010; T. Grande et al. 2018), (16) first proximal radial of dorsal fin and first neural arch and spine in contact (T. Grande et al. 2013), (17) principal caudal-fin rays the only insertion site of caudal fin interradialis muscle (Borden et al. 2013), presence of single procurrent caudal-fin ray (T. Grande et al. 2018), and (18) 12 principal caudal-fin rays (T. Grande et al. 2018).

  • Synonyms. Zeoidei (Regan 1909a:80; Jordan 1923:171), Zeomorphi (Regan 1910a:481–482; Rosen 1984:44; Zehren 1987, fig. 1), Zeacea (Wiley and Johnson 2010:150), and Zeiariae (Betancur-R et al. 2017:20) are ambiguous synonyms of Zeiformes.

  • Comments. Since the mid-20th century Zeiformes was consistently applied as the group name for the clade defined above and was selected as the clade name over its synonyms because it is the name most frequently applied to a taxon approximating the named clade (e.g., Greenwood et al. 1966; Wiley et al. 2000; Borden et al. 2013; T. Grande et al. 2013; Davis et al. 2016; Betancur-R et al. 2017; Ghezelayagh et al. 2022).

  • The earliest fossil Zeiformes is †Cretazeus rinaldii from the Campanian-Maastrichtian of Italy (Appendix 1; Tyler et al. 2000). Bayesian relaxed molecular clock analyses of Zeiformes result in an average posterior crown age estimate of 50.0 million years ago, with the credible interval ranging between 37.3 and 72.0 million years ago (Ghezelayagh et al. 2022).

  • img-z99-7_03.gif

    Gadiformes P. Bleeker 1859:xxvi

  • Definition. The least inclusive crown clade that contains Stylephorus chordatus Shaw 1791 and Gadus morhua Linnaeus 1758. This is a minimum-crown-clade definition, but the clade is not defined using the PhyloCode.

  • Etymology. From the ancient Greek γάδoϛ (ɡˈ̍αːdo͡Ʊz), which was a name applied to the European Hake, Merluccius merluccius (D. W. Thompson 1947:38). The suffix is from the Latin forma meaning form, figure, or appearance.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S1). Phylogenetic relationships of the major lineages of Gadiformes are presented in Figure 13.

  • Phylogenetics. Historically, Stylephorus chordatus was classified in Lampriformes (Regan 1908; Olney et al. 1993; J. S. Nelson 2006:228–229) and Gadoidei was hypothesized to be closely related to Batrachoididae and Lophioidei (Patterson and Rosen 1989). A more recent phylogenetic analysis of Acanthomorpha based on morphological characters resolves Gadoidei as the sister lineage to a clade containing Stylephorus and Zeiformes (Davesne et al. 2016). On the other hand, molecular phylogenetic analyses consistently resolve Stylephorus and Gadoidei as a monophyletic group (Miya et al. 2007; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Davis et al. 2016; Smith et al. 2016; Alfaro et al. 2018; T. Grande et al. 2018; Hughes et al. 2018; Roth et al. 2020; Ghezelayagh et al. 2022; Mu et al. 2022; J.-F. Wang et al. 2023).

  • Composition. There are currently 625 living species of Gadiformes (Fricke et al. 2023) that include Stylephorus chordatus and species classified in Gadoidei. Over the past 10 years, 32 new species of Gadiformes have been described (Cohen et al. 1990; Fricke et al. 2023), comprising 5.1% of the species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Gadiformes include (1) levator arcus palatini lies lateral to section A2 of adductor mandibulae (T. Grande et al. 2013), (2) interradialis located only between caudal-fin rays (Borden et al. 2013), (3) hypochondral longtudinalis absent (Borden et al. 2013), and (4) most ural centra and pleural centra 2–4 exposed after removal of body musculature (Borden et al. 2013).

  • Synonyms.Gadariae(Betancur-Retal.2017:20) is an ambiguous synonym of Gadiformes.

  • Comments. Since the first phylogenetic analysis that resolved Stylephorus and Gadoidei as a monophyletic group (Miya et al. 2007), the substantial molecular evidence is supported by the discovery of morphological apomorphies providing confidence to the resolution of a more inclusive Gadiformes that includes Stylephorus (Borden et al. 2013). Bayesian relaxed molecular clock analyses of Gadiformes result in an average posterior crown age estimate of 88.4 million years ago, with the credible interval ranging between 72.9 and 105.4 million years ago (Ghezelayagh et al. 2022).

  • img-z100-11_03.gif

    Gadoidei L. J. F. J. Fitzinger 1832:331

  • Definition. The least inclusive crown clade that contains Bregmaceros cantori Milliken and Houde 1984, Gadus morhua Linnaeus 1758, and Macruronus novaezelandiae (Hector 1871). This is a minimum-crown-clade definition, but the clade is not defined using the PhyloCode.

  • Etymology. From the ancient Greek γάδoϛ (ɡˈ̍αːdo͡Ʊz), which was a name applied to the European Hake, Merluccius merluccius (D. W. Thompson 1947:38).

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 14,208 exons (Roa-Varón et al. 2021, fig. 4). Phylogenetic relationships of the major lineages of Gadoidei are presented in Figure 13.

  • Phylogenetics. Precladistic studies of relationships within Gadoidei concluded that Melanonus (pelagic cods) (Marshall 1965, 1966; Marshall and Cohen 1973) or Muraenolepididae (eel cods) (Rosen and Patterson 1969; Cohen 1984) represented the lineage with the least derived morphology in the clade. A theme in the study of gadoid morphology is that lineages are characterized by combinations of ancestral and derived character states (Rosen and Patterson 1969; Cohen 1984; Okamura 1989; Endo 2002). The heterogeneous nature of gadoid morphology is reflected in the dramatically minimal congruence among the more than 10 morphological phylogenies that examined a wide range of osteological, myological, and otolith characters (J. R. Dunn 1989; Howes 1989, 1991a, 1993; Iwamoto 1989; Markle 1989; Nolf and Steurbaut 1989; Okamura 1989; Howes 1990; Siebert 1990; Endo 2002; Teletchea et al. 2006; Grand et al. 2014). Heterochronic evolution has been invoked to explain the characteristic mosaic of ancestral and derived morphology in Gadoidei (Endo 2002), highlighting the potential challenges of using morphological characters to resolve phylogenetic relationships within the clade.

  • The first set of molecular phylogenetic studies of Gadoidei utilized data from Sanger-sequenced mitochondrial and nuclear genes and resulted in phylogenies with relatively poor node support (Møller et al. 2002; Bakke and Johansen 2005; Teletchea et al. 2006; von der Heyden and Matthee 2008; Roa-Varón and Ortí 2009; Betancur-R et al. 2017), limiting the ability of these analyses to resolve the deepest nodes in the gadoid phylogeny. Despite the challenge of limited resolution, the first group of gadoid molecular phylogenies demonstrated that previous delimitations of Merlucciidae (merluccid hakes) (Inada 1989; Cohen et al. 1990; Lloris et al. 2005) are not monophyletic, motivating the recognition of the monogeneric taxonomic families Macruronidae (southern grenadiers), Lyconidae (Atlantic hakes), and Steindachneriidae (Steindachneria argentea, Luminous Hake) (von der Heyden and Matthee 2008; Roa-Varón and Ortí 2009).

  • Next-generation phylogenomic analyses vary in the level of resolution and node support, but all result in phylogenies in which Bregmaceros (codlets) is placed as the sister lineage of all other Gadoidei (Malmstrøm et al. 2016; Hughes et al. 2018; Han et al. 2021; Roa-Varón et al. 2021; Ghezelayagh et al. 2022). The phylogenetic resolution of Bregmaceros is consistent with the observation that this lineage is “fundamentally different myologically and osteologically from other gadoids” (Rosen and Patterson 1969:427). The phylogenomic analyses agree with several earlier molecular studies in resolving Gadidae (cods), Lotidae (burbots), and Phycidae (hakes), all previously classified as Gadidae, as a monophyletic group (von der Heyden and Matthee 2008; Roa-Varón and Ortí 2009; Betancur-R et al. 2017). There is appreciable congruence among the trees generated from phylogenomic analyses but there is disagreement regarding the relationships of Muraenolepididae, Trachyrincidae (armored grenadiers), Melanonus, and Merlucciidae (Malmstrøm et al. 2016; Hughes et al. 2018; Han et al. 2021; Roa-Varón et al. 2021; Ghezelayagh et al. 2022).

  • Composition. There are currently 624 living species of Gadoidei that include Raniceps raninus (Tadpole Fish), Steindachneria argentea, and species classified in Bathygadidae (rattails), Bregmaceros, Euclichthys (eucla cods), Gadidae, Gaidropsaridae (rocklings), Lotidae, Lyconus, Macrouridae (grenadiers), Macruronus (blue grenadiers), Melanonus, Merlucciidae, Moridae (morid cods), Muraenolepididae, Phycidae, and Trachyrincidae (Cohen et al. 1990; Lloris et al. 2005; Roa-Varón et al. 2021; Fricke et al. 2023). Over the past 10 years, 32 new species of Gadoidei have been described (Fricke et al. 2023), comprising 5.1% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Gadoidei include (1) presence of X and Y bones in caudal skeleton (Cohen 1984; Fahay and Markle 1984; Markle 1989; Patterson and Rosen 1989), (2) first neural spine joined to occipital crest (Cohen 1984; Patterson and Rosen 1989), (3) larvae with anus that exits through the finfold (Fahay and Markle 1984; Markle 1989), (4) absence of ribs or epipleurals on vertebrae 1 and 2 (Markle 1989; Patterson and Rosen 1989; Howes 1993), (5) scapular foramen located between scapula and coracoid (Markle 1989; Patterson and Rosen 1989; Howes 1993; Endo 2002; Wiley and Johnson 2010), (6) notch in the front of the prootic is exit point for trigeminal and facial nerves from braincase, with absence of lateral commissure or trigeminofacial chamber (Patterson and Rosen 1989), (7) anal and dorsal fins with three fin rays per segment (Patterson and Rosen 1989), (8) head canals with 33 neuromasts (Patterson and Rosen 1989), (9) presence of three struts on pharyngobranchial 3 (Markle 1989), (10) otolith with pince-nezshaped sulcus and lateral collicular (Nolf and Steurbaut 1989; Endo 2002; Wiley and Johnson 2010), (11) absence of jugular foramen (Howes 1991a, 1993), (12) attrition of anterior border of lateral face of hyomandibular, exposing pathway of the hyoid branch of the facial nerve (Howes 1993), (13) dorsal hyomandibular with a single condyle (Endo 2002; Wiley and Johnson 2010), (14) basihyal absent (Endo 2002; Wiley and Johnson 2010; T. Grande et al. 2013), (15) flexor dorsalis and flexor ventralis separate with some bundles serving a single ray compounded (Borden et al. 2013), (16) flexor dorsalis and flexor dorsalis superior are a single muscle mass (Borden et al. 2013), and (17) flexor ventralis and flexor ventralis inferior are a single muscle mass (Borden et al. 2013).

  • Synonyms. Gadiformes (e.g., Cohen 1984; Patterson and Rosen 1989:13–19; Cohen et al. 1990, fig. 1; A. C. Gill and Mooi 2002, tbl. 2.3; Wiley and Johnson 2010:148–149; J. S. Nelson et al. 2016:293–302; Betancur-R et al. 2017:20) is an ambiguous synonym of Gadoidei.

  • Comments. The application of phylogenetic systematics has contributed to a flux in the classification of lineages that comprise Gadoidei since the mid-1980s (Cohen 1984; Markle 1989; Endo 2002; Roa-Varón and Ortí 2009; Roa-Varón et al. 2021). A more recent Linnaean classification of gadoids has an abundance of redundant group names as it recognizes 17 families of which seven comprise a single genus and five suborders of which three contain a single family (Roa-Varón et al. 2021).

  • The early fossil record of Gadoidei is dominated by otoliths (Kriwet and Hecht 2008). The earliest otolith fossils of Gadoidei include †Rhinocephalus cretaceus and †Archaemacruroides vanknippenbergi from the Maastrichtian (72.2–66.0 Ma) of Belgium and Netherlands (Schwarzhans and Jagt 2021), and †Dakotaichthys hogansoni, †Palaeogadus weltoni, and †Archaemacruroides bratishkoi from the Maastrichtian of Texas, USA (Schwarzhans and Stringer 2020). Bayesian relaxed molecular clock analyses of Gadoidei result in an average posterior crown age estimate of 77.0 million years ago, with the credible interval ranging between 61.5 and 98.2 million years ago (Ghezelayagh et al. 2022).

  • img-z102-6_03.gif

    Acanthopterygii P. Artedi 1738:26
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Diretmus argenteus Johnson 1864 (Trachichthyiformes), Beryx decadactylus Cuvier 1829 in Cuvier and Valenciennes (1829b) (Beryciformes), Holocentrus rufus (Walbaum 1792) (Beryciformes), Carapus bermudensis (Jones 1874) (Ophidiiformes), and Micropterus salmoides (Lacépède 1802) (Centrarchiformes), but not Percopsis omiscomaycus (Walbaum 1792) (Percopsiformes) nor Gadus morhua Linnaeus 1758 (Gadiformes). This is a minimum-crown-clade definition with external specifiers.

  • Etymology. From the ancient Greek ἇκανθα (æk̍ænθә) meaning thorn or spine and πτερόν (t̍εɹαːn) meaning feather, wing, or any winged animal.

  • Registration number. 943.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, figs. S2–S25). Phylogenetic relationships of the major lineages of Acanthopterygii are presented in Figures 2 and 14. The placements of the pan-trachichthyiforms †Judeoberyx and †Lissoberyx, and the pan-percomorph †Pepemkay in the phylogeny are on the basis of inferences from morphological studies (Moore 1993a, 1993b; Patterson 1993; Friedman 2009; Cantalice et al. 2021).

  • Phylogenetics. Morphological and molecular phylogenetic analyses consistently support the monophyly of Acanthopterygii (Miya et al. 2003, 2005; W. L. Smith and Wheeler 2006; Alfaro, Santini, et al. 2009; Santini et al. 2009; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; T. Grande et al. 2013; Near et al. 2013; W.-J. Chen, Santini, et al. 2014; Davesne et al. 2016; Davis et al. 2016; Malmstrøm et al. 2016, 2017; Smith et al. 2016; Betancur-R et al. 2017; Hughes et al. 2018; Musilova et al. 2019; Roth et al. 2020; Cantalice et al. 2021; Ghezelayagh et al. 2022; Mu et al. 2022; J.-F. Wang et al. 2023). However, earlier morphological studies nested the paracanthopterygian Zeiformes in Acanthopterygii (Stiassny and Moore 1992; G. D. Johnson and Patterson 1993).

  • Phylogenetic analyses of Acanthopterygii differ on the relationships among Trachichthyiformes, Beryciformes, and Percomorpha. Morphological phylogenetic studies led to delimitations of Trachichthyiformes and Beryciformes that differ from current classifications and deviate from one another primarily in the relationships of Holocentridae and Berycidae (Rosen 1973; Stiassny and Moore 1992; G. D. Johnson and Patterson 1993; Moore 1993b). Molecular studies result in four sets of phylogenies of Acanthopterygii: a clade containing Beryciformes and Trachichthyiformes that is the sister lineage of Percomorpha (W. L. Smith and Wheeler 2006; Alfaro, Santini, et al. 2009; Santini et al. 2009; Near, Eytan, et al. 2012; T. Grande et al. 2013; Near et al. 2013; Malmstrøm et al. 2017; Mu et al. 2022), Beryciformes (excluding Holocentridae) and Trachichthyiformes as a monophyletic group that is the sister lineage of Holocentridae (Near et al. 2013; Rabosky et al. 2018; J.-F. Wang et al. 2023), Beryciformes (excluding Holocentridae) and Trachichthyiformes as a monophyletic group that is the sister lineage of a clade containing Holocentridae and Percomorpha (Betancur-R, Broughton, et al. 2013; Smith et al. 2016; Betancur-R et al. 2017), and Trachichthyiformes as the sister lineage of a clade containing Beryciformes and Percomorpha (Figures 2 and 14; Miya et al. 2003, 2005; Thacker 2009; W.-J. Chen, Santini, et al. 2014; Malmstrøm et al. 2016; Dornburg et al. 2017; Hughes et al. 2018, fig. S2; Musilova et al. 2019; Roth et al. 2020; Ghezelayagh et al. 2022).

  • Morphological studies aimed at resolving the relationships of several Cretaceous acanthomorph fossil lineages place Hoplostethus (Trachichthyiformes) as the sister lineage of a clade containing Sargocentron (Beryciformes) and Percomorpha, but these studies are limited in taxon sampling and do not test the monophyly of Trachichthyiformes or Beryciformes (Davesne et al. 2016; Cantalice et al. 2021). Bayesian concordance factors estimated in a phylogenomic analysis of UCE loci find the hypothesis that Beryciformes and Percomorpha are sister lineages is supported by the greatest proportion of sampled loci, and phylogenies that depict either Holocentridae or a clade containing Beryciformes and Trachichthyiformes as the sister lineage of Percomorpha are less optimal (Ghezelayagh et al. 2022).

  • Composition. There are currently more than 19,185 living species of Acanthopterygii classified in Trachichthyiformes, Beryciformes, and Percomorpha. Fossil lineages include the pan-trachichthyiforms †Judeoberyx and †Lissoberyx (Patterson 1967; Gayet 1980b), and the pan-percomorph †Pepemkay (Alvarado-Ortega and Than-Marchese 2013). Details of the ages and locations of fossil taxa are presented in Appendix 1. Over the past 10 years, 1,641 new living species of Acanthopterygii have been described (Fricke et al. 2023), comprising 8.6% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Acanthopterygii include (1) retractor dorsalis muscle inserts primarily on pharyngobranchial 3 (Rosen 1973), (2) reduction of surface of epibranchial 4 and enlargement of epibranchials 2 and 3, which form the primary support for the upper pharyngeal jaw dentition (Rosen 1973), (3) presence of two hyomandibular articulation facets (Davesne et al. 2016), (4) proximal insertion of Baudelot's ligament onto the basioccipital (Davesne et al. 2016), and (5) presence of an antero-median pelvic process (Davesne et al. 2016).

  • Synonyms. Euacanthomorphacea (Betancur-R, Broughton, et al. 2013, app. 2) is an ambiguous synonym of Acanthopterygii. Euacanthopterygii (G. D. Johnson and Patterson 1993:607) is an approximate synonym of Acanthopterygii.

  • Comments. Classifications of Acanthomorpha differ in the application of the group name Acanthopterygii: (1) to the paraphyletic group that includes Zeiformes, Lampriformes, Trachichthyiformes, Beryciformes, and Percomorpha (Greenwood et al. 1966); (2) the likely paraphyletic group containing Lampriformes, Trachichthyiformes, Beryciformes, and Percomorpha (Davis et al. 2016); and (3) the clade containing Trachichthyiformes, Beryciformes, and Percomorpha as defined here (J. S. Nelson et al. 2016:302–303; Betancur-R et al. 2017; Hughes et al. 2018; Ghezelayagh et al. 2022). The name Acanthopterygii was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The morphological characterization of Acanthopterygii was hampered by previous phylogenetic studies and resulting classifications that placed the acanthopterygian lineages Ophidiiformes, Batrachoididae, and Lophioidei into Paracanthopterygii and treated the paracanthopterygian Zeiformes as an acanthopterygian (Lauder and Liem 1983; J. S. Nelson 1984, 1994, 2006; Rosen 1984; Patterson and Rosen 1989; Stiassny and Moore 1992; G. D. Johnson and Patterson 1993). The concept of Acanthopterygii as limited to Trachichthyiformes, Beryciformes, and Percomorpha originated from many molecular phylogenetic analyses (e.g., Miya et al. 2003; Alfaro, Santini, et al. 2009; Near, Eytan, et al. 2012; W.-J. Chen, Santini, et al. 2014; Malmstrøm et al. 2016; Ghezelayagh et al. 2022) and is validated in phylogenetic analyses of morphological characters (Davesne et al. 2016; Cantalice et al. 2021).

  • The earliest acanthopterygian fossils date to the Cenomanian (100.5–93.9 Ma) (Patterson 1993; Friedman 2009, 2010; Murray 2016). Bayesian relaxed molecular clock analyses of Actinopterygii result in an average posterior age estimate of 137.4 million years ago, with the credible interval ranging between 129.1 and 147.3 million years ago (Ghezelayagh et al. 2022).

  • img-z105-6_03.gif

    FIGURE 14.

    Phylogenetic relationships of the major living lineages and fossil taxa of Acanthopterygii, Trachichthyiformes, Beryciformes, Berycoidei, Percomorpha, Ophidiiformes, Bythitoidei, Gobiiformes, Apogonoidei, Gobioidei, Ovalentaria, and Eupercaria. Filled circles identify the common ancestor of clades with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z104-1_03.jpg

    Trachichthyiformes M. L. J. Stiassny and J. A. Moore 1992:212, figs. 14, 15, and 16
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Trachichthys australis Shaw 1799 in Shaw and Nodder (1799), Diretmus argenteus J. Y. Johnson 1864, and Aulotrachichthys prosthemius (Jordan and Fowler 1902), but not Beryx decadactylus Cuvier in Cuvier and Valenciennes (1829b) nor Holocentrus rufus (Walbaum 1792). This is a minimum-crown-clade definition with external specifiers.

  • Etymology. From the ancient Greek τρᾱχὐς (tɹˈ̍e͡Ikәs) meaning rough and ἰχθὐς (̍kIkθuːs) meaning fish. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 944.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S2). Although Trachichthys australis is not included in the reference phylogeny, it resolves in a clade with other species of Trachichthyidae in a phylogenetic analysis of morphological characters (Zehren 1979, figs. 4, 5). Phylogenetic relationships of the major lineages of Trachichthyiformes are presented in Figure 14.

  • Phylogenetics. Morphological analyses resolved the prephylogenetic delimitation of Beryciformes as paraphyletic relative to Zeiformes and Percomorpha (Stiassny and Moore 1992; G. D. Johnson and Patterson 1993). The relationships of these lineages differed among morphological phylogenetic analyses. One set of studies introduced Trachichthyiformes as a clade that includes Anomalopidae (flashlight fishes), Anoplogaster (fangtooths), Diretmidae (spinyfins), Monocentridae (pinecone fishes), Trachichthyidae (roughies), and all lineages delimited here as Beryciformes except for Berycidae (alfonsinos) and Holocentridae (squirrelfishes) (Stiassny and Moore 1992; Moore 1993b). In a different analysis of morphological characters, a clade Stephanoberyciformes was resolved that includes all lineages delimited here as Beryciformes to the exclusion of Berycidae and Holocentridae and a definition of Beryciformes that included what is delimited here as Trachichthyiformes with the addition of Berycidae and Holocentridae (G. D. Johnson and Patterson 1993).

  • The monophyly of Trachichthyiformes is supported in morphological (Zehren 1979; Moore 1993b; Baldwin and Johnson 1995; Konishi and Okiyama 1997) and molecular phylogenetic studies (Miya et al. 2003, 2005; T. Grande et al. 2013; Near et al. 2013; Davis et al. 2016; Betancur-R et al. 2017; Dornburg et al. 2017; Malmstrøm et al. 2017; Musilova et al. 2019; Ghedotti et al. 2021; Ghezelayagh et al. 2022; Mu et al. 2022). Phylogenetic relationships among lineages of Trachichthyiformes inferred from morphological and molecular data differ substantially. Morphological inferences resolve Anoplogaster and Diretmidae as a clade that is the sister lineage of all other Trachichthyiformes (Moore 1993b; Konishi and Okiyama 1997). Molecular phylogenies and analyses of combined molecular and morphological datasets consistently place Diretmidae as the sister lineage of all other Trachichthyiformes (Miya et al. 2003, 2005; Near et al. 2013; Betancur-R et al. 2017; Musilova et al. 2019; Ghedotti et al. 2021; Ghezelayagh et al. 2022).

  • Composition. There are currently 71 living species of Trachichthyiformes (Fricke et al. 2023) classified in Anomalopidae, Anoplogaster, Diretmidae, Monocentridae, and Trachichthyidae. Over the past 10 years, four new living species of Trachichthyiformes have been described (Su et al. 2022a, 2022b; Fricke et al. 2023), comprising 5.6% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies of Trachichthyiformes include (1) presence of X pattern on the frontal (Zehren 1979; Moore 1993b; Ghedotti et al. 2021), (2) ethmoid very small and confined to area between upper portions of lateral ethmoids (Zehren 1979; Moore 1993b; Ghedotti et al. 2021), (3) presence of bony arches over infraorbitals (Moore 1993b), (4) presence of tack-like scales on larvae (Baldwin and Johnson 1995; Konishi and Okiyama 1997), (5) presence of ornamentation on lateral face of opercle in larvae (Baldwin and Johnson 1995), and (6) presence of spicules on rays of dorsal, anal, caudal, and pectoral fins of larvae (Konishi and Okiyama 1997; Ghedotti et al. 2021).

  • Synonyms. Trachichthyoidei (Moore 1993b:115) is an ambiguous synonym of Trachichthyiformes.

  • Comments. The group name Trachichthyiformes was initially applied to the paraphyletic group that included all species of Trachichthyiformes and Beryciformes to the exclusion of Berycidae (Stiassny and Moore 1992; Moore 1993b). Trachichthyiformes was used as the group name for the clade defined here in classifications resulting from molecular phylogenetic analyses (Betancur-R et al. 2017; Ghezelayagh et al. 2022) and was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest fossil Trachichthyiformes is the trachichthyid †Gephyroberyx robustus from the Rupelian (33.9–27.82 Ma) of the Caucasus area of Russia (Danil'chenko 1960). Bayesian relaxed molecular clock analyses of Trachichthyiformes result in an average posterior crown age estimate of 46.6 million years ago, with the credible interval ranging between 24.7 and 75.5 million years ago (Ghezelayagh et al. 2022).

  • img-z106-9_03.gif

    Beryciformes A. C. L. G. Günther 1880:419
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Beryx decadactylus Cuvier in Cuvier and Valenciennes (1829b) and Holocentrus rufus (Walbaum 1792), but not Diretmus argenteus J. Y. Johnson 1864 nor Aulotrachichthys prosthemius (Jordan and Fowler 1902). This is a minimum-crown-clade definition with external specifiers.

  • Etymology. From the ancient Greek βρρυς (b̍e͡I͡ɪɹuːz) meaning fish. The word is known primarily from the lexicon of the fifth- or sixth-century CE grammarian Hesychius of Alexandria (D. W. Thompson 1947:32). The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 945.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S2). Phylogenetic relationships of the major living and fossil lineages of Beryciformes are presented in Figure 14. The phylogenetic placements of the pan-holocentrids †Berybolcensis, †Iridopristis, †Plesioberyx, †Stichocentrus, and †Tenuicentrum, and the pan-berycoid †Berycomorus are on the basis of inferences from morphology (Friedman 2009; Andrews et al. 2023).

  • Phylogenetics. No phylogenetic analysis of morphological characters has resolved Beryciformes as a monophyletic group (Stiassny and Moore 1992; G. D. Johnson and Patterson 1993; Moore 1993b). Molecular phylogenetic analyses differ on the monophyly of Beryciformes, but the incongruence is limited to the identity of the sister lineage of Holocentridae (squirrelfishes). One set of molecular analyses resolves Beryciformes as paraphyletic, with Holocentridae and Percomorpha as sister lineages (Betancur-R, Broughton, et al. 2013; Smith et al. 2016; Betancur-R et al. 2017). Alternatively, another group of analyses results in phylogenies in which a monophyletic Beryciformes is placed as the sister group of Percomorpha (Figures 2 and 14; Miya et al. 2003; Thacker 2009; W.-J. Chen, Santini, et al. 2014; Malmstrøm et al. 2016; Dornburg et al. 2017; Hughes et al. 2018, fig. S2; Musilova et al. 2019; Ghezelayagh et al. 2022). Within Beryciformes, Berycoidei and Holocentridae are resolved as sister lineages (Figure 14; Miya et al. 2003, 2005; Thacker 2009; Near, Eytan, et al. 2012; W.-J. Chen, Santini, et al. 2014; Malmstrøm et al. 2016; Dornburg et al. 2017; Hughes et al. 2018, fig. S2; Musilova et al. 2019; Roth et al. 2020; Ghedotti et al. 2021; Ghezelayagh et al. 2022).

  • Composition. There are 213 living species of Beryciformes (Fricke et al. 2023) classified in Berycoidei and Holocentridae. Fossil lineages include the pan-holocentrids †Berybolcensis, †Iridopristis, †Plesioberyx, †Stichocentrus, and †Tenuicentrum (Patterson 1967; Gayet 1980a; Andrews et al. 2023), and the pan-berycoid †Berycomorus (Arambourg 1966). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, 20 new living species of Beryciformes have been described (Fricke et al. 2023), comprising 9.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. There are no known morphological apomorphies for Beryciformes (Moore 1993b; Ghedotti et al. 2021).

  • Synonyms. Berycimorphaceae (Betancur-R et al. 2017:21) is a partial synonym of Beryciformes.

  • Comments. Since the introduction of Beryciformes as a taxonomic group (Günther 1880), its composition has included Polymixia, Caristiidae (Scombriformes), Ostracoberyx (Acropomatiformes), and lineages now classified as Trachichthyiformes (Starks 1904; Regan 1911c; Berg 1940:467–468; Patterson 1964:432–434; McAllister 1968; Gosline 1971:147–148; Rosen 1973; J. S. Nelson 1984:232–240; G. D. Johnson and Patterson 1993; Near, Eytan, et al. 2012). A more restricted composition of Beryciformes came after the mid-20th century in a series of morphological phylogenetic studies (Zehren 1979; Stiassny and Moore 1992; Moore 1993b). Molecular phylogenetic analyses consistently support the monophyly of Beryciformes (e.g., Hughes et al. 2018, fig. S2; Ghezelayagh et al. 2022), highlighting the need for morphological studies to continue testing this hypothesis with the aim of discovering morphological apomorphies for the clade. The name Beryciformes was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest fossils of Beryciformes include the Cenomanian (100.5–93.9 Ma) pan-holocentrids †Stichocentrus liratus, †S. elegans, †S. spinulosus, †Plesioberyx maximus, and †P. discoides from Lebanon (Patterson 1967; M. Gaudant 1969; Gayet 1980a; Forey et al. 2003). Bayesian relaxed molecular clock analyses of Beryciformes result in an average posterior crown age estimate of 95.8 million years ago, with the credible interval ranging between 71.6 and 117.6 million years ago (Andrews et al. 2023).

  • img-z108-3_03.gif

    Berycoidei P. Bleeker 1874:15
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Beryx decadactylus Cuvier in Cuvier and Valenciennes (1829b) and Cetostoma regani Zugmayer 1914. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek word βˈρυς (be͡Iɪɹuːz) meaning fish. The word is known primarily from the lexicon of the fifth- or sixth-century CE grammarian Hesychius of Alexandria (D. W. Thompson 1947:32).

  • Registration number. 946.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, fig. S2). Phylogenetic relationships of the major lineages of Berycoidei are presented in Figure 14. The relationships of Gibberichthys follow Kobyliansky et al. (2020) and Hispidoberyx follow Moore (1993b) and Ghedotti et al. (2021).

  • Phylogenetics. No phylogenetic analysis of morphological characters has resolved Berycoidei as monophyletic (Stiassny and Moore 1992; G. D. Johnson and Patterson 1993; Moore 1993b); however, the resolution of Stephanoberycoidei in a morphological phylogeny differs from the composition of Berycoidei in the exclusion of Berycidae (alfonsinos) (Moore 1993b). With a notable exception (Colgan et al. 2000), molecular phylogenetic analyses consistently resolve Berycoidei as a clade (Miya et al. 2003, 2005; W. L. Smith and Wheeler 2006; Dettaï and Lecointre 2008; Thacker 2009; Near, Eytan, et al. 2012; T. Grande et al. 2013; Near et al. 2013; Davis et al. 2016; Betancur-R et al. 2017; Hughes et al. 2018; Rabosky et al. 2018; Ghedotti et al. 2021; Ghezelayagh et al. 2022). Within Berycoidei, Berycidae and Melamphaidae (ridgeheads) comprise a clade that is the sister lineage to all other berycoids (e.g., Miya et al. 2003; Near et al. 2013; Betancur-R et al. 2017; Ghezelayagh et al. 2022). Molecular phylogenies differ on the relationships of Barbourisia rufa (Velvet Whalefish), Cetomimidae (flabby whalefishes), and Stephanoberycidae (pricklefishes). Phylogenies inferred from mtDNA or combinations of mtDNA and nuclear genes resolve Barbourisia and Cetomimidae as sister lineages (Near et al. 2013; Rabosky et al. 2018; Kobyliansky et al. 2020; Ghedotti et al. 2021), consistent with inferences from morphology (Moore 1993b). However, phylogenetic analyses of a supermatrix of Sanger-sequenced genes and a dataset comprising more than 980 UCE loci resolve Barbourisia and Stephanoberycidae as a clade (Betancur-R et al. 2017; Ghezelayagh et al. 2022).

  • Morphological phylogenies resolve Gibberichthys (gibberfishes) as the sister lineage of a clade containing Hispidoberyx ambagiosus (Bristlyskin) and Stephanoberycidae, and resolve the deepsea whalefishes as a monophyletic group comprising Barbourisia, Cetomimidae, and Rondeletia (red mouth whalefishes) (Moore 1993b). However, the presence of Tominaga's organ, a large globular mass of tissue below the nasal rosette with a potential secretory function, was presented as morphological evidence that Gibberichthys and Rondeletia are sister lineages (Paxton et al. 2001), a result supported in a phylogenetic analysis of mtDNA gene sequences (Kobyliansky et al. 2020). The previously recognized lineages Mirapinnidae (tapetails) and Megalomycteridae (bignose fishes) are larvae and males, respectively, of species of Cetomimidae (G. D. Johnson et al. 2009).

  • Composition. There are currently 123 living species of Berycoidei that include Barbourisia rufa, Hispidoberyx ambagiosus, and species classified in Berycidae, Cetomimidae, Gibberichthys, Melamphaidae, Rondeletia, and Stephanoberycidae. Over the past 10 years, 13 new living species of Berycoidei have been described (Fricke et al. 2023), comprising 10.6% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Berycoidei are found in all lineages except Berycidae and include (1) ocular sclera absent (Moore 1993b; Ghedotti et al. 2021), (2) orbitosphenoid absent (Moore 1993b; Ghedotti et al. 2021), (3) cranium with thinly ossified bones consisting mostly of cartilage and connective tissue (Moore 1993b; Ghedotti et al. 2021), and (4) lower branchial tooth patches absent (Moore 1993b; Ghedotti et al. 2021).

  • Synonyms. Stephanoberycoidei (Moore 1993b, fig. 5; J. S. Nelson et al. 2016:308–309; Betancur-R et al. 2017:21; Afonso et al. 2021) is a partial synonym of Berycoidei.

  • Comments. The group name Berycoidei has been applied to several para- and polyphyletic groups, including: (1) Trachichthyidae and Holocentridae (Patterson 1964); (2) Berycidae, Trachichthyidae, Diretmidae, Anoplogaster, Anomalopidae, and Holocentridae (Greenwood et al. 1966); (3) Berycidae and Melamphaidae (J. S. Nelson et al. 2016:313–314; Betancur-R et al. 2017); or (4) limited to Berycidae (Nelson 1994:288, 2006:302–303). The composition of Berycoidei as defined here follows the results of several molecular phylogenetic analyses (e.g., Davis et al. 2016; Betancur-R et al. 2017; Hughes et al. 2018; Rabosky et al. 2018; Ghedotti et al. 2021; Ghezelayagh et al. 2022) and was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • Bayesian relaxed molecular clock analyses of Berycoidei result in an average posterior crown age estimate of 85.8 million years ago, with the credible interval ranging between 65.8 and 101.9 million years ago (Ghezelayagh et al. 2022).

  • img-z109-8_03.gif

    Percomorpha O. P. Hay 1903:693
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade that contains Carapus bermudensis (Jones 1874) (Ophidiiformes), Perca fluviatilis Linnaeus 1758 (Perciformes), and Micropterus salmoides (Lacépède 1802) (Centrarchiformes), but not Diretmoides pauciradiatus (Woods 1973) in Woods and Sonoda (1973) (Trachichthyiformes) nor Beryx decadactylus Cuvier 1829 in Cuvier and Valenciennes (1829b) (Beryciformes). This is a minimum-crown-clade definition with external specifiers.

  • Etymology. From the ancient Greek πέρκη (p̍ːke͡I), a name applied to many species of fishes by ancient authors (D. W. Thompson 1947:195–197) and µoρϕή (m̍ͻfiː) meaning form or shape.

  • Registration number. 947.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, figs. S3–S25). Phylogenetic relationships of Percomorpha are shown in Figures 2 and 14. In the phylogeny, the placement of the fossil pan-ophidiiform †Pastorius is the more conservative of two hypotheses presented by Carnevale and Johnson (2015) and resolution of the pan-batrachoid †Bacchiaichthys follows Carnevale and Collette (2014).

  • Phylogenetics. Percomorpha was first delimited as a result of comparative morphological studies and included all lineages currently classified in Acanthopterygii except Atheriniformes (=Atherinomorpha), Batrachoididae, Lophioidei (=Lophiiformes), and Ophidiiformes (Rosen and Patterson 1969). Over the next two decades, Percomorpha and Atheriniformes were presented as sister lineages in several phylogenetic trees (Hinegardner and Rosen 1972; Rosen 1973, 1982; C. L. Smith 1975; Rosen and Parenti 1981; Lauder 1983; Lauder and Liem 1983). During this period, Percomorpha was identified as a clade that was inadequately diagnosed with morphological characters and contained many lineages with unresolved relationships (Rosen 1982; Lauder and Liem 1983). Subsequent morphological phylogenetic studies indicated that Rosen and Patterson's (1969) concept of Percomorpha was paraphyletic due to the resolution of Mugilidae as the sister lineage to Atheriniformes (Stiassny 1990, 1993). A subsequent review and investigation of acanthomorph phylogeny based on 34 morphological characters led to a redefinition of Percomorpha to include Atheriniformes and exclude Trachichthyiformes and Beryciformes (G. D. Johnson and Patterson 1993).

  • Phylogenies resulting from analyses of molecular data offer a refined delimitation of Percomorpha that not only includes Atheriniformes, but also the lineages Batrachoididae, Lophioidei, and Ophidiiformes that were previously classified in Paracanthopterygii (W.-J. Chen et al. 2003; Miya et al. 2003, 2005). Subsequent molecular phylogenetic analyses consistently support the monophyly of this revised delimitation of Percomorpha (W. L. Smith and Wheeler 2006; Davis 2010; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; T. Grande et al. 2013; Near et al. 2013; W.-J. Chen, Santini, et al. 2014; Davis et al. 2016; Malmstrøm et al. 2016, 2017; Smith et al. 2016; Betancur-R et al. 2017; Alfaro et al. 2018; Hughes et al. 2018; Roth et al. 2020; Ghezelayagh et al. 2022; Mu et al. 2022; J.-F. Wang et al. 2023). Molecular studies with inclusive taxon sampling resolve 13 major clades within Percomorpha, with Ophidiiformes and Batrachoididae as the first of two successive branching lineages in the clade; Scombriformes and Syngnathiformes as sister lineages; a clade containing Ovalentaria, Synbranchiformes, and Carangiformes; and Eupercaria as a clade containing Perciformes, Centrarchiformes, Labriformes, Acropomatiformes, and Acanthuriformes (Near et al. 2013; Smith et al. 2016; Betancur-R et al. 2017; Dornburg and Near 2021; Ghezelayagh et al. 2022). Phylogenetic analyses of morphological characters support the monophyly of Percomorpha (Davesne et al. 2016; Cantalice et al. 2021), but these studies limit taxon sampling to four species, one representing each of Acanthuriformes, Batrachoididae, Carangiformes, and Ophidiiformes.

  • Composition. Percomorpha currently includes more than 18,900 living species (Fricke et al. 2023), classified in the subclades Ophidiiformes, Batrachoididae, Syngnathiformes, Scombriformes, Ovalentaria, Gobiiformes, Synbranchiformes, Carangiformes, and Eupercaria. Fossil lineages include the pan-ophidiiform †Pastorius (Carnevale and Johnson 2015) and the pan-batrachoid †Bacchiaichthys (Bannikov and Sorbini 2000; Carnevale and Collette 2014). Details of the ages and locations for the fossil taxa are given in Appendix 1. Over the past 10 years, 1,090 new living species of Percomorpha have been described (Fricke et al. 2023), comprising 5.8% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Percomorpha include (1) external dorsal pelvic wing equal in size to external ventral wing (Stiassny and Moore 1992; Davesne et al. 2016), (2) first epibranchial and second pharyngobranchial with rodlike interarcual cartilage present between separated uncinate processes (G. D. Johnson and Patterson 1993; W. L. Smith 2005; Wiley and Johnson 2010), (3) absence of second ural centrum (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), (4) five or fewer hypurals (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), (5) fewer than six rays in pelvic fin (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010; Davesne et al. 2016), (6) absence of free pelvic radials (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010; Davesne et al. 2016), (7) all but the first two epineurals have a point of origin that is displaced ventrally with distal parts of all epineurals displaced ventrally into the horizontal septum (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), (8) 17 principal caudal rays arranged as I,8,7,I (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010; Davesne et al. 2016; Cantalice et al. 2021), (9) absence of anterior supramaxilla (Davesne et al. 2016; Cantalice et al. 2021), (10) absence of orbitosphenoid (Davesne et al. 2016; Cantalice et al. 2021), (11) anterior and posterior ceratohyals sutured (Davesne et al. 2016), and (12) the first dorsal pterygiophore inserts between neural spines 2 and 4 (Davesne et al. 2016).

  • Synonyms. Percomorphacea (Wiley and Johnson 2010:127, 151–152; Betancur-R et al. 2017:22) is an ambiguous synonym of Percomorpha.

  • Comments. Percomorpha was famously referred to the “bush at the top of the tree” in reference to the limited phylogenetic resolution among the more than 18,900 species and at least 288 taxonomic families in the clade (G. J. Nelson 1989:328). This was later restated as the “percomorph problem” in reference to the lack of morphological apomorphies diagnosing the group and the fact that Percomorpha represented the largest polytomy in the phylogeny of living vertebrates, a consequence of too many lineages and too few morphological characters to resolve relationships (G. D. Johnson and Patterson 1993; Chakrabarty 2010). Despite impressive efforts that involve careful and elegant studies of comparative morphology (G. D. Johnson and Patterson 1993; Patterson and Johnson 1995; Datovo et al. 2014; Pastana et al. 2022), the status of efforts using morphology to resolve the phylogeny of Percomorpha is summarized as “any tree can be justified by special pleading, by insisting that certain characters are uniquely derived but others are more labile or plastic” as “very few of the characters found among percomorphs and their relatives are uniquely derived” (G. D. Johnson and Patterson 1993:555). Molecular phylogenetics has not only led to a dramatic increase in the resolution of relationships within Percomorpha, but has also provided a mechanism for the development of exciting and surprising hypotheses of relationships that were undiscovered and wholly unanticipated from the study of morphology (Dornburg and Near 2021). The future of phylogenetic studies of Percomorpha likely involves a full integration of molecular phylogenetics and comparative morphology as evidenced by studies that lead to reinterpretations of morphological traits in the context of phylogenies resulting from analysis of molecular data (e.g., Chanet et al. 2013; Ghedotti et al. 2018; M. G. Girard et al. 2020).

  • Since the turn of the 21st century, Percomorpha is consistently delimited as including Ophidiiformes and Batrachoididae and excluding Beryciformes and Trachichthyiformes (Miya et al. 2003, 2005; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; T. Grande et al. 2013; Near et al. 2013; W.-J. Chen, Santini, et al. 2014; Davis et al. 2016; J. S. Nelson et al. 2016:314–315; Smith et al. 2016; Betancur-R et al. 2017; Alfaro et al. 2018; Hughes et al. 2018; Dornburg and Near 2021; Ghezelayagh et al. 2022; Mu et al. 2022; J.-F. Wang et al. 2023). The name Percomorpha was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest fossils of Percomorpha all date to the Campanian and Maastrichtian (83.6–72.1, 72.1–66.0 Ma) of the Late Cretaceous and include the pan-ophidiiform †Pastorius (Carnevale and Johnson 2015), the pan-batrachoid †Bacchiaichthys (Bannikov and Sorbini 2000), and the pan-centriscoid †Gasterorhamphosus (Sorbini 1981). Bayesian relaxed molecular clock analyses of Percomorpha result in an average posterior crown age estimate of 126.8 million years ago, with the credible interval ranging between 116.9 and 135.6 million years ago (Ghezelayagh et al. 2022).

  • img-z111-7_03.gif

    Ophidiiformes P. Bleeker 1859:xxv
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Ophidion barbatum Linnaeus 1758, Dinematichthys iluocoeteoides Bleeker 1855, Aphyonus gelatinosus Günther 1878a, Brotula barbata (Bloch and Schneider 1801), Carapus acus (Brünnich 1768), and Dicrolene introniger Goode and Bean 1883. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek όφίς (̍o͡ƱfIs) meaning snake. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 948.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S3). Although Ophidion barbatum is not included in the reference phylogeny, it resolves in a clade with other species of Ophidion in a phylogenomic analysis of Sanger-sequenced mitochondrial and nuclear genes (Betancur-R et al. 2017, fig. S6; Rabosky et al. 2018). Phylogenetic relationships of the major living lineages and fossil taxa of Ophidiiformes are presented in Figure 14. Placements of the fossil taxa in the phylogeny are on the basis of inferences from morphology (Patterson and Rosen 1989; Schwarzhans 2003, 2010; Møller et al. 2016; Schwarzhans and Stringer 2020).

  • Phylogenetics. Ophidiiformes was previously classified in Paracanthopterygii on the basis of studies of morphology (e.g., Greenwood et al. 1966; Rosen and Patterson 1969; Patterson and Rosen 1989; J. S. Nelson 2006:243–248), but they are distantly related to paracanthopterygians and are resolved as the sister group of all other Percomorpha in molecular phylogenetic analyses (Miya et al. 2003, 2005; W. L. Smith and Wheeler 2006; Davis 2010; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; T. Grande et al. 2013; Near et al. 2013; W.-J. Chen, Santini, et al. 2014; Davis et al. 2016; Malmstrøm et al. 2016; Betancur-R et al. 2017; Alfaro et al. 2018; Hughes et al. 2018; Roth et al. 2020; Ghezelayagh et al. 2022; Mu et al. 2022). Despite being resolved as monophyletic in analyses of molecular data (Miya et al. 2003; Near et al. 2013; Møller et al. 2016; Betancur-R et al. 2017; Campbell, Nielsen, et al. 2017; Ghezelayagh et al. 2022), there is little evidence from morphology for the monophyly of Ophidiiformes (Rosen 1985; Patterson and Rosen 1989; Howes 1992; Nielsen et al. 1999).

  • The mode of reproduction is an important trait in classifying Ophidiiformes into the oviparous Ophidiidae (cusk eels) and viviparous Bythitoidei (Cohen and Nielsen 1978; Nielsen et al. 1999; J. S. Nelson et al. 2016). Phylogenies inferred from molecular data result in paraphyly of the traditional delimitation of Ophidiidae because of the resolution of Carapidae (pearlfishes) (Miya et al. 2003, 2005; Near et al. 2013; Betancur-R et al. 2017; Rabosky et al. 2018; Ghezelayagh et al. 2022; M. G. Girard et al. 2023), prompting the delineation of a more inclusive Ophidiidae to include species previously classified in Carapidae (Betancur-R et al. 2017). Molecular phylogenetic analyses resolve both the more inclusive Ophidiidae and Bythitoidei as monophyletic groups (Near et al. 2013; Møller et al. 2016; Betancur-R et al. 2017; Campbell, Nielsen, et al. 2017; Rabosky et al. 2018; Ghezelayagh et al. 2022).

  • Composition. There are currently 569 living species of Ophidiiformes (Nielsen et al. 1999; Fricke et al. 2023) classified in Ophidiidae and Bythitoidei. Fossil lineages of Ophidiiformes include the pan-bythitoid †“Bidenichthyscrepidatus and the pan-ophiid †Ampheristus americanus (Schwarzhans 2003, 2010; Schwarzhans and Stringer 2020). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, there have been 43 new living species of Ophidiiformes described (Fricke et al. 2023), comprising 7.6% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Ophidiiformes include (1) supraoccipital excluded from posterior cranial margin by posterodorsal extension of exoccipitals (Howes 1992; Carnevale and Johnson 2015), (2) presence of angled bursa-like cavity between exoccipitals and basioccipital, and (3) posterior portion of first infraorbital covered by second infraorbital (Ohashi 2018).

  • Synonyms. Ophidiicae (Hubbs 1952:51, fig. 1), Ophidiimorpharia (Betancur-R, Broughton, et al. 2013:13), Ophidiida (J. S. Nelson et al. 2016:315), and Ophidiaria (Sanciangco et al. 2016, fig. 1; Betancur-R et al. 2017:22) are ambiguous synonyms of Ophidiiformes.

  • Comments. Ophidiiformes is a diverse clade with more than 560 species classified among 121 genera (Fricke et al. 2023), but very little of this rich diversity has been integrated into phylogenetic studies (Møller et al. 2016; Rabosky et al. 2018). The migration of Ophidiiformes, Batrachoididae, and Lophioidei from Paracanthopterygii to Percomorpha speaks to the effect of molecular data on inferring the phylogeny of ray-finned fishes and is “akin to placing a morphologically established lineage of marsupials as the sister lineage of rodents or vipers as the sister lineage of Anolis” (Dornburg and Near 2021:441).

  • The earliest fossil Ophidiiformes are the pan-bythitoid †“Bidenichthyscrepidatus and the pan-ophiid †Ampheristus americanus from the Maastrichtian (72.2–66.0 Ma) in the Cretaceous (Appendix 2; Voigt 1926; Schwarzhans 2010; Schwarzhans and Stringer 2020). Bayesian relaxed molecular clock analyses of Ophidiiformes result in an average posterior crown age estimate of 84.5 million years ago, with the credible interval ranging between 59.3 and 111.3 million years ago (Ghezelayagh et al. 2022).

  • img-z113-5_03.gif

    Bythitoidei D. M. Cohen and J. G. Nielsen 1978:42
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Bythites fuscus Reinhardt 1837, Dinematichthys iluocoeteoides Bleeker 1855, and Aphyonus gelatinosus Günther 1878a. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek βῠθὀς (bˈuːθo͡Ʊz) meaning the depths of the sea.

  • Registration number. 949.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S3). Although Bythites fuscus is not included in the reference phylogeny, morphological studies indicate that B. fuscus, species of Grammonus, and species of Cataetyx share common ancestry (Cohen and Nielsen 1978). Phylogenetic relationships of the major living lineages and fossil taxa of Bythitoidei are presented in Figure 14. Placements of the fossil taxa in the phylogeny are on the basis of inferences from morphology (Schwarzhans 2003, 2010; Møller et al. 2016; Schwarzhans and Stringer 2020).

  • Phylogenetics. Bythitoidei was delimited to include Bythitidae (livebearing brotulas) and Aphyonidae (aphyonids) based on the presence of an intromittent organ in males and the placement of the anterior nostril well above the upper lip (Cohen and Nielsen 1978). From the late 1960s through the 1990s, Parabrotulidae (false brotulas) was classified in Zoarcoidei on the basis of the presence of a one-to-one ratio of vertebrae to fin pterygiophores, an eel-shaped body, ventral fins, lack of fin spines, and a confluent dorsal and anal fin (Nielsen 1968; Cohen and Nielsen 1978; Nielsen et al. 1990; Miya and Nielsen 1991). It was argued that the presence of paired nostrils, a bilobed ovary, and a well-developed intromittent organ in Parabrotulidae is evidence for their shared ancestry with Ophidiiformes, specifically Bythitoidei, and not Zoarcoidei (Anderson 1994; Nelson 1994:227). A detailed analysis of the osteology of Parabrotula plagiophthalmus highlighted the morphology of the intromittent organ and the presence of six caudal rays as consistent with shared common ancestry of Parabrotulidae and Bythitidae (Hilton et al. 2021).

  • Molecular phylogenetic analyses consistently resolve Bythitoidei as monophyletic (Near et al. 2013; Møller et al. 2016; Betancur-R et al. 2017; Campbell, Nielsen, et al. 2017; Evseenko et al. 2018; Arroyave et al. 2022; Ghezelayagh et al. 2022). Parabrotulidae and the abyssal Aphyonidae are phylogenetically nested in Bythitidae, while Bythitidae and Dinematichthyidae (viviparous brotulas) are resolved as sister lineages making up the more inclusive clade Bythitoidei (Møller et al. 2016; Betancur-R et al. 2017; Campbell, Nielsen, et al. 2017; Ghezelayagh et al. 2022). The results from molecular phylogenetic analyses were the basis for the reclassification of Aphyonidae and Parabrotulidae within Bythitidae and the elevation of Dinematichthyidae from lineages formerly classified in Dinematichthyini (Møller et al. 2016).

  • Composition. There are currently 246 living species of Bythitoidei (Møller et al. 2016; Fricke et al. 2023) classified in Bythitidae and Dinematichthyidae. Fossil lineages of Bythitoidei include †Bythitidarum rasmussenae from the Danian (66.0–61.7 Ma) of Denmark (Appendix 1; Schwarzhans 2003; Møller et al. 2016). Over the past 10 years, there have been 13 new living species of Bythitoidei described (Fricke et al. 2023), comprising 5.3% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Bythitoidei include (1) presence of a male intromittent organ (Cohen and Nielsen 1978; Patterson and Rosen 1989; Wiley and Johnson 2010), (2) anterior nostril positioned low on snout and close to upper lip (Cohen and Nielsen 1978; Patterson and Rosen 1989), and (3) reduction of pelvic fin to a single ray or entirely absent (Møller et al. 2016).

  • Synonyms. There are no synonyms of Bythitoidei.

  • Comments. The earliest fossil taxon of Bythitoidei is the otolith species †Bythitidarum rasmussenae from the Danian (66.0–61.7 Ma) of Denmark (Schwarzhans 2003; Møller et al. 2016). Bayesian relaxed molecular clock analyses of Bythitoidei result in an average posterior crown age estimate of 46.0 million years ago, with the credible interval ranging between 28.1 and 69.3 million years ago (Ghezelayagh et al. 2022).

  • img-z114-6_03.gif

    Batrachoididae D. S. Jordan 1896:231
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Opsanus tau (Linnaeus 1766), Batrachoides pacifici (Günther 1861), and Halobatrachus didactylus (Bloch and Schneider 1801). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek word βάτρχoς (bætɹ̍æko͡Ʊz) meaning frog.

  • Registration number. 950.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S3). The phylogenetic resolution of Batrachoididae relative to other lineages of Percomorpha is presented in Figures 2 and 14.

  • Phylogenetics. Batrachoididae (toadfishes) was previously classified in Paracanthopterygii (e.g., Greenwood et al. 1966; Rosen and Patterson 1969; Patterson and Rosen 1989; J. S. Nelson 2006:243–248), and viewed as closely related to Lophioidei (Regan 1912c; Patterson and Rosen 1989; Datovo et al. 2014). Molecular phylogenetic analyses resolve Batrachoididae as nested within Percomorpha, and most studies place batrachoids as the sister lineage of an inclusive clade that contains all other percomorphs except for Ophidiiformes (Miya et al. 2005; W. L. Smith and Wheeler 2006; Davis 2010; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; T. Grande et al. 2013; Near et al. 2013; W.-J. Chen, Santini, et al. 2014; Davis et al. 2016; Malmstrøm et al. 2016; Betancur-R et al. 2017; Alfaro et al. 2018; Hughes et al. 2018; Roth et al. 2020; Ghezelayagh et al. 2022; Mu et al. 2022).

  • The monophyly of Batrachoididae is supported in several molecular and morphological phylogenetic analyses (W. L. Smith and Wheeler 2006; Near et al. 2013; Betancur-R et al. 2017; Hughes et al. 2018; Rabosky et al. 2018; Vaz 2020; Ghezelayagh et al. 2022). Morphological and molecular phylogenetic studies infer relationships within Batrachoididae that are congruent, with Halophryninae resolved as the sister group of a clade containing Batrachoidinae, Porichthyinae, and Thalassophryninae (Greenfield et al. 2008; Rice and Bass 2009; Rabosky et al. 2018). A detailed description of the caudal skeleton identified potential apomorphies for Batrachoididae and several subclades (Vaz and Hilton 2020). A phylogenetic analysis of 191 morphological characters with extensive taxon sampling resolves lineages of Batrachoididae in a polytomy containing Triathalassothia, a clade of lineages traditionally classified in Halophryninae (Barchatus, Batrichthys, Bifax, Chatrabus, Colletteichthys, Halobatrachus, Perulibatrachus, and Riekertia), and a clade containing Halophryninae (Allenbatrachus, Batrachomoeus, and Halophryne), Thalassophryninae (Daector and Thalassophryne), Porichthyinae (Aphos and Porichthys), and Batrachoidinae (Amphichthys, Batrachoides, Opsanus, Sanopus, and Vladichthys) (Vaz 2020).

  • Composition. There are currently 84 living species of Batrachoididae (Fricke et al. 2023) that include Bifax lacinia, Halobatrachus didactylus, Riekertia ellisi, and species classified in Batrachoidinae, Halophryninae, Porichthyinae, Thalassophryninae, and Triathalassothia (Greenfield et al. 2008; Vaz 2020; Fricke et al. 2023). Over the past 10 years, one new living species of Batrachoididae was described (Fricke et al. 2023), comprising 1.2% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Batrachoididae include (1) large yolk sac with a ventral adhesive disc present in larvae (Wiley and Johnson 2010), (2) tightly packed configuration in the dorsal spine and pterygiophore complex (Wiley and Johnson 2010), (3) robust and hypertrophied epineural bound to medial surface of cleithrum (Wiley and Johnson 2010; Vaz 2020), (4) supracleithrum articulates with ankylosed posttemporal (Wiley and Johnson 2010; Vaz 2020), (5) parietals absent (Wiley and Johnson 2010), (6) mesethmoid unossified (Wiley and Johnson 2010; Vaz 2020), (7) swimbladder heart shaped with anterior portion separated in two lobes with bands of musculature along the lateral surface of each lobe (Wiley and Johnson 2010), (8) dorsal edge of metapterygoid with a trapezoidal shape (Vaz 2020), (9) subopercle with one, two, or three spines (Vaz 2020), (10) urohyal with lateral projections giving a T-shape (Vaz 2020), (11) uncinate process longer than the anterior half of epibranchial process (Vaz 2020), (12) fifth ceratobranchial is one-half the length of fourth ceratobranchial (Vaz 2020), (13) origin of first epineural bone articulates with the neural spine of the first vertebra (Vaz 2020), (14) the origin of third epineural at the level of neural arch of third vertebra (Vaz 2020), (15) ventral limb of posttemporal reduced to a knob (Vaz 2020), (16) presence of anterodorsal process of the supracleithrum, (17) propterygium hypertrophied as long pectoral radials (Vaz 2020; Vaz and Hilton 2023), (18) propterygium rod shaped (Vaz 2020), and (19) presence of a filamentous cushion organ on the pelvic spine and lateralmost soft ray (Vaz 2020).

  • Synonyms. Haplodoci (Cope 1871a:458), Batrachoidiformes (Berg 1937:1279; Greenwood et al. 1966:396; Lauder and Liem 1983, fig. 37; Patterson and Rosen 1989:23–24; Wiley and Johnson 2010:159–160; J. S. Nelson et al. 2016:320–321; Betancur-R et al. 2017:22), Batrachoidimorpharia (Betancur-R et al. 2013a:13), Batrachoidida (J. S. Nelson et al. 2016:320), and Batrachoidaria (Betancur-R et al. 2017:22) are ambiguous synonyms of Batrachoididae.

  • Comments. Batrachoididae is a valid family-group name under the International Code of Zoological Nomenclature (Van der Laan et al. 2014:64), has long been applied as the group name for the clade presented in the definition (Jordan 1923:238; McAllister 1968:164; J. S. Nelson et al. 2016:321–323), and was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest fossil taxon of Batrachoididae is the otolith-based species †Batrachoididarum trapezoidalis from the Ypresian (56.0–48.1 Ma) of France (Nolf 1988; Carnevale and Collette 2014), and the earliest skeletal fossil is †Louckaichthys novosadi from the Rupelian (33.9–27.3 Ma) of the Czech Republic (Přikryl and Carnevale 2017). Bayesian relaxed molecular clock analyses of Batrachoididae result in an average posterior crown age estimate of 49.1 million years ago, with the credible interval ranging between 25.3 and 76.3 million years ago (Ghezelayagh et al. 2022).

  • img-z116-2_03.gif

    Gobiiformes P. Bleeker 1859:xxv
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Gobius niger Linnaeus 1758, Lythrypnus dalli (Gilbert 1890), Trichonotus filamentosus (Steindachner 1867), Ostorhinchus doederleini (Jordan and Snyder 1901), and Kurtus indicus Bloch 1786. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek κωβιός (k̍o͡ƱbIo͡Ʊz) meaning small insignificant fish. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 951.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, figs. S3, S4). Although Gobius niger is not included in the reference phylogeny, it resolves in a clade with other species of Gobiidae in phylogenetic analyses of Sanger-sequenced mitochondrial and nuclear genes (Tornabene et al. 2013, fig. 2; McCraney et al. 2020, fig. 6). Phylogenetic relationships among the major lineages of Gobiiformes are presented in Figure 14. Placement of the fossil pan-gobioid †Paralates in the phylogeny is on the basis of an analysis of morphological characters (Gierl et al. 2022).

  • Phylogenetics. One of the most remarkable results from molecular phylogenetic analyses of Percomorpha is the discovery that Apogonidae, Gobioidei, Kurtus, and Trichonotus resolve in a strongly supported clade, delimited here as Gobiiformes, that is the sister lineage of a clade containing all other lineages of Percomorpha exclusive of Ophidiiformes and Batrachoididae (Thacker and Hardman 2005; W. L. Smith and Wheeler 2006; W. L. Smith and Craig 2007; Thacker 2009; Thacker and Roje 2009; Chakrabarty et al. 2012; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Thacker et al. 2015; Betancur-R et al. 2017; Alfaro et al. 2018; Hughes et al. 2018; Kuang et al. 2018; McCraney et al. 2020; Ghezelayagh et al. 2022; Satoh and Katayama 2022). Within Gobiiformes, a clade containing Gobioidei and Trichonotus is the sister lineage of Apogonoidei, including Apogonidae and Kurtus (Near et al. 2013; Thacker et al. 2015; Betancur-R et al. 2017; Rabosky et al. 2018; McCraney et al. 2020; Ghezelayagh et al. 2022). Alternative phylogenetic relationships among Gobiiformes resulting from molecular phylogenetic analyses include the resolution of Kurtus as the sister lineage of all other Gobiiformes (Thacker 2009; Chakrabarty et al. 2012; Alfaro et al. 2018; Kuang et al. 2018) and a clade containing Apogonidae, Kurtus, and Trichonotus as the sister lineage of Gobioidei (Satoh and Katayama 2022).

  • Composition. There are currently 2,740 living species of Gobiiformes (Fricke et al. 2023) classified in Apogonoidei, Gobioidei, and Trichonotus. Fossil lineages include the pan-gobioid †Paralates (Gierl and Reichenbacher 2017; Gierl et al. 2022). Over the past 10 years, 368 new living species of Gobiiformes have been described (Fricke et al. 2023), comprising 13.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Gobiiformes include (1) presence of large gap between symplectic and preopercle in Gobioidei and Trichonotus (J. S. Nelson 1986; Winterbottom 1993b), (2) presence of sensory papillae rows on the head and body in Gobioidei, Kurtus, and Apogonidae (G. D. Johnson 1993; Thacker 2009), but see Sato (2022), and (3) presence of eggs with adhesive filaments around the micropyle in Gobioidei, Kurtus, and Apogonidae (G. D. Johnson 1993; Thacker et al. 2015). All Gobiiformes engage in egg guarding or brooding by the male, either in a benthic nest (Gobioidei), in the mouth (Apogonidae), on the forehead (Kurtus), or in the gill chamber (Trichonotus) (Clark and Pohle 1996; Berra and Humphrey 2002; Östlund-Nilsson and Nilsson 2004; Thacker et al. 2015).

  • Synonyms. Gobiomorpharia (Betancur-R, Broughton, et al. 2013, fig. 1) and Gobiaria (Betancur-R et al. 2017:23) are ambiguous synonyms of Gobiiformes. Gobiida (J. S. Nelson et al. 2016:323) is a partial synonym of Gobiiformes.

  • Comments. Gobiiformes has been applied as a group name for (1) a group that included Apogonidae, Gobioidei, Kurtus, and Pempheridae (Thacker 2009); (2) a clade containing Trichonotus and Gobioidei (Betancur-R et al. 2017); and (3) a clade containing Apogonoidei, Gobioidei, and Trichonotus as presented here in the definition (Thacker et al. 2015; Davis et al. 2016; Dornburg and Near 2021; Ghezelayagh et al. 2022). The name Gobiiformes was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The resolution of a monophyletic Gobiiformes is one of many unexpected results stemming from molecular phylogenetic analyses of Percomorpha. Morphological studies exploring the phylogenetic affinities of Apogonidae, Gobioidei, Kurtus, and Trichonotus all predate the resolution of these lineages as a clade in molecular studies (G. D. Johnson 1993; Winterbottom 1993b; D. G. Smith and Johnson 2007). A potentially fruitful area of future research is the exploration of comparative morphological and anatomical studies among the seemingly disparate lineages that comprise Gobiiformes, with the goal of understanding their history of phenotypic trait diversification and the discovery of additional morphological apomorphies.

  • The earliest fossils of Gobiiformes are the otolith-based species of †Apogonidarum classified as Apogonidae from the Maastrichtian (72.2–66.0 Ma) in the Cretaceous of India and North Dakota, USA (Khajuria and Prasad 1998; Hoganson et al. 2019). The earliest skeletal fossils of Gobiiformes include the gobioid †Carlomonnius and the apogonids †Apogoniscus, †Bolcapogon, †Eoapogon, †Eosphaeramia, and †Leptolumamia all from the Ypresian (56.0–48.1 Ma) of Monte Bolca, Italy (Bannikov and Carnevale 2016; Bannikov and Fraser 2016). Bayesian relaxed molecular clock analyses of Gobiiformes result in an average posterior crown age estimate of 109.6 million years ago, with the credible interval ranging between 98.9 and 119.9 million years ago (Ghezelayagh et al. 2022).

  • img-z117-7_03.gif

    Gobioidei Bleeker 1849:4
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Gobius niger Linnaeus 1758, Lythrypnus dalli (Gilbert 1890), Periophthalmus barbarus (Linnaeus 1766), Eleotris pisonis (Gmelin 1789), Milyeringa veritas Whitley 1945, and Rhyacichthys aspro (Valenciennes in Cuvier and Valenciennes 1837). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek κωβιός (k̍o͡ƱbIˌo͡Ʊz) meaning small insignificant fish.

  • Registration number. 952.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, fig. S4). Although Gobius niger is not included in the reference phylogeny, it resolves in a clade with other species of Gobiidae in phylogenetic analyses of Sanger-sequenced mitochondrial and nuclear genes (Tornabene et al. 2013, fig. 2; McCraney et al. 2020, fig. 6). Phylogenetic relationships among the living and fossil lineages of Gobioidei are presented in Figure 14. Placements of the fossil pan-butids †Carlomonnius and †Lepidocottus and the pan-thalasseleotrids †Eleogobius and †Pirskenius follow Gierl et al. (2022).

  • Phylogenetics. Prior to the application of molecular data to the study of fish phylogeny, the relationships of Gobioidei among Percomorpha were unresolved (P. J. Miller 1973, 1986; Springer 1983; Hoese 1984). In a study of osteological characters, Winterbottom (1993b) concluded that Hoplichthys, Gobiesocidae, Callionymidae, and various “trachinoids” that included Creediidae, Hemerocoetidae, and Trichonotus were the lineages with the greatest number of character states shared with Gobioidei. Despite many morphological apomorphies diagnosing Gobioidei (Springer 1983; Hoese 1984; P. J. Miller 1992; G. D. Johnson and Brothers 1993; Winterbottom 1993b), comparative morphological studies did not provide a strong hypothesis for the phylogenetic affinities of gobioids among percomorphs. Morphology of the dorsal gill arches was cited as evidence of shared common ancestry for Apogonidae and Kurtus (G. D. Johnson 1993).

  • Gobioidei is consistently resolved as monophyletic in molecular phylogenetic studies (W. L. Smith and Wheeler 2006; Thacker 2009; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Thacker et al. 2015, 2023; Betancur-R et al. 2017; Hughes et al. 2018; Kuang et al. 2018; McCraney et al. 2020; Ghezelayagh et al. 2022). Within Gobioidei, molecular phylogenies resolve a clade containing Rhyacichthyidae (loach gobies) and Odontobutidae (freshwater sleepers) as the sister lineage of all other gobioids, with Milyeringidae (blind cave gobies), Eleotridae (spinycheek sleepers), Butidae (butid sleepers), and Thalasseleotrididae (ocean sleepers) as successive branching lineages leading to a clade containing Gobiidae (gobies) and Oxudercidae (mudskippers and relatives) (Thacker et al. 2015; McCraney et al. 2020; Ghezelayagh et al. 2022; Goatley and Tornabene 2022). A phylogenomic analysis of Gobioidei using UCE loci resolves Xenisthmus and Butidae as sister lineages, prompting the elevation of Xenisthmidae (collared wrigglers) out of synonymy with Eleotridae (Thacker 2003; McCraney 2019).

  • Phylogenies inferred from morphological characters are fairly congruent with relationships inferred from molecular data (Hoese and Gill 1993), specifically in resolving Rhyacichthyidae and Odontobutidae as the sister lineage of all other gobioids and supporting Thalasseleotrididae as the sister lineage of a clade containing Gobiidae and Oxudercidae (A. C. Gill and Mooi 2012; Reichenbacher et al. 2020; Gierl et al. 2022). The presence of five branchiostegal rays is consistent with the monophyly of a clade containing the gobioid lineages Thalasseleotrididae, Oxudercidae, and Gobiidae (Hoese 1984; Hoese and Gill 1993; A. C. Gill and Mooi 2012; Reichenbacher et al. 2020); the remaining lineages Rhyacichthyidae, Odontobutidae, Milyeringidae, Xenisthmidae, Eleotridae, and Butidae all have six branchiostegal rays. Molecular phylogenetic studies focusing on specific gobioid lineages have attempted to resolve relationships within Rhyacichthys (Haÿ et al. 2022), Odontobutidae (H. Li et al. 2018), Butidae, and Eleotridae (Thacker and Hardman 2005; Thacker 2017; Thacker, Shelley, McCraney, Adams, et al. 2022; Thacker, Shelley, McCraney, Unmack, et al. 2022), Oxudercidae (Yamada et al. 2009; Thacker 2013; Thacker et al. 2019; McMahan et al. 2021), and Gobiidae (Rüber et al. 2003; Herler et al. 2009; Neilson and Stepien 2009; Thacker and Roje 2011; Tornabene et al. 2013, 2023).

  • Composition. There are currently 2,347 living species of Gobioidei (Fricke et al. 2023) classified in Rhyacichthyidae, Odontobutidae, Milyeringidae, Xenisthmidae, Butidae, Eleotridae, Thalasseleotrididae, Oxudercidae, and Gobiidae. Fossil Gobioidei include the pan-butids †Carlomonnius and †Lepidocottus and the pan-thalasseleotrids †Eleogobius and †Pirskenius (Gierl et al. 2013, 2022; Přikryl 2014; Gierl and Reichenbacher 2015; Bannikov and Carnevale 2016; Reichenbacher et al. 2020). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, 349 new living species of Gobioidei have been described (Fricke et al. 2023), comprising 14.9% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Gobioidei include (1) parietals absent (Springer 1983; G. D. Johnson and Brothers 1993; Winterbottom 1993b), (2) basisphenoid absent (Springer 1983; G. D. Johnson and Brothers 1993; Winterbottom 1993b), (3) two or fewer (usually zero) infraorbitals (Springer 1983; G. D. Johnson and Brothers 1993), (4) interhyal attached to preopercle by a ligament, not articulating at junction of symplectic and hyomandibular, resulting in gap between symplectic and preopercle (Springer 1983; G. D. Johnson and Brothers 1993; Winterbottom 1993b), (5) basibranchial 1 cartilaginous (Springer 1983; Winterbottom 1993b), (6) pelvic intercleithral cartilage present (Springer 1983; Winterbottom 1993b), (7) ventral intercleithral cartilage present (Springer 1983; Winterbottom 1993b), (8) sagittae and lapilli with elongate primordia (Brothers 1984; G. D. Johnson and Brothers 1993; Winterbottom 1993b), (9) accessory sperm-duct glands present in males (P. J. Miller 1992; G. D. Johnson and Brothers 1993), (10) supraneurals absent (Springer 1983; G. D. Johnson and Brothers 1993), (11) neural and haemal arches and spines developing as membrane bones with little to no cartilaginous precursors (G. D. Johnson and Brothers 1993), (12) first neural arch fused to first centrum at earliest appearance in ontogeny (G. D. Johnson and Brothers 1993), (13) dorsalmost pectoral ray articulating with posterior margin of dorsalmost actinost or radial cartilage rather than with scapula, medial part of ray lacking enlarged articular base and in early ontogeny not embracing ovoid cartilage lying at posterodorsal corner of scapulocoracoid cartilage (G. D. Johnson and Brothers 1993), and (14) hypurals 1+2 and 3+4 fused to one another and to the urostyle (G. D. Johnson and Brothers 1993; Winterbottom 1993b).

  • Synonyms. Gobiiformes (Betancur-R, Broughton, et al. 2013, fig. 3; J. S. Nelson et al. 2016:326) is an ambiguous synonym of Gobioidei.

  • Comments. Gobioidei has long been applied as the group name for the clade presented in the definition and was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade (McAllister 1968:146–147; P. J. Miller 1973; Nelson 1994:412–418; Thacker 2009; McCraney et al. 2020).

  • Time-calibrated phylogenies of acanthopterygians have repeatedly identified Gobioidei as containing clades with significantly elevated rates of lineage diversification (Near et al. 2013; Rabosky et al. 2013, 2018; Ghezelayagh et al. 2022). Comparative studies have deployed phylogenies of Gobioidei to investigate the history of phenotypic diversification (Thacker 2014, 2017; Thacker and Gkenas 2019; Huie et al. 2020) and the biogeography of near-shore marine habitats (Thacker 2015, 2017; Tornabene et al. 2016).

  • The earliest Gobioidei fossil is the pan-butid †Carlomonnius from the Ypresian (56.0–47.8 Ma) of Monte Bolca, Italy (Bannikov and Carnevale 2016). Bayesian relaxed molecular clock analyses of Gobioidei result in an average posterior crown age estimate of 93.6 million years ago, with the credible interval ranging between 82.1 and 104.6 million years ago (Ghezelayagh et al. 2022).

  • img-z119-7_03.gif

    Apogonoidei C. E. Thacker 2009:100
    [C. E. Thacker and T. J. Near], converted clade name.

  • Definition. The least inclusive crown clade that contains Kurtus indicus Bloch 1786, Pseudamia gelatinosa J. B. L. Smith 1955a, Apogon imberbis (Linnaeus 1758), and Cheilodipterus quinquelineatus Cuvier 1828 in Cuvier and Valenciennes (1828). This is a minimum-crown-clade definition.

  • Etymology. From the Greek prefix α- (a-) meaning without, and the ancient Greek πώγων (p̍o͡Ʊαːn) meaning beard.

  • Registration number. 953.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S3). AlthoughApogonimberbisisnotincludedinthe reference phylogeny, it resolves in a clade with other species of Apogonidae in phylogenetic analyses of Sanger-sequenced mitochondrial and nuclear genes (Mabuchi et al. 2014, figs. 2–6). The phylogenetic relationships of Apogonoidei are presented in Figure 14.

  • Phylogenetics. The monophyly of Apogonoidei is supported in molecular phylogenetic analyses (W. L. Smith and Craig 2007; Near et al. 2013; Thacker et al. 2015; Betancur-R et al. 2017; Rabosky et al. 2018; McCraney et al. 2020; Ghezelayagh et al. 2022) and is consistent with suggestions of a close relationship between Apogonidae (cardinalfishes) and Kurtus (nurseryfishes) based on morphological characters of the gill arches, axial skeleton, and fine structures of the egg micropyle and filaments (G. D. Johnson 1993; Prokofiev 2006b).

  • Composition. There are currently 383 living species of Apogonoidei (Fricke et al. 2023) classified in Apogonidae and Kurtus. Over the past 10 years, 19 new living species of Apogonoidei have been described (Fricke et al. 2023), comprising 5.0% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Apogonoidei include (1) second epibranchial articulates with third rather than second pharyngobranchial (G. D. Johnson 1993), (2) head of third pharyngobranchial expanded and much larger than fourth (G. D. Johnson 1993), (3) fourth pharyngobranchial cartilage absent (G. D. Johnson 1993), and (4) radial ridges of simple or bifid filaments around the micropyle of the eggs (G. D. Johnson 1993).

  • Synonyms. Kurtiformes (Betancur-R, Broughton, et al. 2013, fig. 3; J. S. Nelson et al. 2016:324; Betancur-R et al. 2017:23) is an ambiguous synonym of Apogonoidei.

  • Comments. The group name Apogonoidei has been applied to (1) a group containing Apogonidae and Pempheridae (Thacker 2009; Thacker and Roje 2009), (2) limited to Apogonidae (Betancur-R, Broughton, et al. 2013; Betancur-R et al. 2017; McCraney et al. 2020), and (3) a clade containing Apogonidae and Kurtus as presented here in the definition (Thacker et al. 2015; Ghezelayagh et al. 2022). The name Apogonoidei was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • Apogonidae and Kurtus each have highly derived egg brooding behaviors in which the eggs bear filaments that allow them to adhere into a ball that is guarded in the mouth of male Apogonidae or on a forehead hook extending from the supraoccipital in Kurtus (Berra and Humphrey 2002; Berra 2003; Östlund-Nilsson and Nilsson 2004; Mabuchi et al. 2014). Several lineages of Apogonidae, including Jaydia, Rhabdamia, Siphamia, and Taeniamia, contain species with bioluminescent organs elaborated from the gut, which may host symbiotic luminescent bacteria or generate light endogenously; this luminescence has evolved multiple times within Apogonidae (Thacker 2009; Fraser 2013).

  • The earliest fossils of Apogonoidei are the otolith-based species of †Apogonidarum that are listed as Apogonidae from the Maastrichtian (72.2–66.0 Ma) in the Cretaceous of India and North Dakota, USA (Khajuria and Prasad 1998; Hoganson et al. 2019). The earliest skeletal fossils of Apogonoidei include the apogonids †Apogoniscus, †Bolcapogon, †Eoapogon, †Eosphaeramia, and †Leptolumamia, all from the Ypresian (56.0–48.1 Ma) of Monte Bolca, Italy (Bannikov and Fraser 2016). Bayesian relaxed molecular clock analyses of Apogonoidei result in an average posterior crown age estimate of 81.9 million years ago, with the credible interval ranging between 46.6 and 110.4 million years ago (Ghezelayagh et al. 2022).

  • img-z120-10_03.gif

    Scombriformes A. S. Woodward 1901:418
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Arripis trutta (Bloch and Schneider 1801), Icosteus aenigmaticus Lockington 1880, Scomber scombrus Linnaeus 1758, Brama japonica Hilgendorf 1878, and Trichiurus lepturus Linnaeus 1758. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek σκόµβρoς (sk̍αːmbɹo͡Ʊz), which was the name for the Atlantic Mackerel, Scomber scombrus (D. W. Thompson 1947:243). The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 954.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, figs. S5, S6). See Figures 2 and 14 for the phylogenetic resolution of Scombriformes within Percomorpha and Figure 15 for a phylogeny of the living lineages and fossil taxa comprising Scombriformes. The placements of the fossil pan-trichiurid †Anenchelum, the pan trichiuroid †Argestichthys, the pan-chiasmodontid †Bannikovichthys, the pan-pomatomid †Carangopsis, and the pan-stromateid †Pinichthys are on the basis of inferences from morphology (Bannikov 1987, 1988, 2014b; Prokofiev 2002b; Carnevale 2007; Carnevale et al. 2014; Beckett et al. 2018b; Friedman et al. 2019; Collar et al. 2022).

  • Phylogenetics. Molecular phylogenetic analyses led to the discovery of the clade delimited here as Scombriformes (W.-J. Chen et al. 2003; W. L. Smith and Craig 2007; Dettaï and Lecointre 2008; B. Li et al. 2009; Yagishita et al. 2009; Wainwright et al. 2012; Betancur-R, Broughton, et al. 2013; Miya et al. 2013; Near et al. 2013; Davis et al. 2016; Sanciangco et al. 2016; Smith et al. 2016; Betancur-R et al. 2017; Alfaro et al. 2018; Campbell et al. 2018; Hughes et al. 2018; Friedman et al. 2019; Arcila et al. 2021; Harrington et al. 2021; Ghezelayagh et al. 2022), consisting of lineages that were never grouped together in classifications based on morphology (Greenwood et al. 1966; Wiley and Johnson 2010). Scombriformes includes lineages previously classified in Scombroidei (Scombridae [mackerels and tunas], Scombrolabrax heterolepis [Longfin Escolar], Gempylidae [snake mackerels], and Trichiuridae [cutlassfishes]) and Stromateoidei (Amarsipus carlsbergi [Amparsipas], Ariomma [ariommatids], Centrolophidae [medusafishes], Nomeidae [driftfishes], Stromateidae [butterfishes], and Tetragonuridae [squaretails]) (Greenwood et al. 1966; Haedrich 1967; Collette, Potthoff, et al. 1984; Horn 1984; G. D. Johnson 1986). The billfishes, Istiophoridae (marlins) and Xiphias gladius (Swordfish), have been classified in Scombroidei since the earliest 20th century (Regan 1909b), but are distantly related to Scombriformes in molecular phylogenies (e.g., Orrell et al. 2006; Little et al. 2010; Hughes et al. 2018; Ghezelayagh et al. 2022).

  • Relationships among major lineages of Scombriformes resulting from phylogenomic analyses are characterized by a lack of resolution among the earliest nodes in the phylogeny that is likely the result of gene tree discordance and short branch lengths (e.g., Friedman et al. 2019; Arcila et al. 2021; Harrington et al. 2021; Ghezelayagh et al. 2022). Despite the limited resolution, phylogenomic analyses resolve several clades in Scombriformes that include: a clade containing Stromateidae (butterfishes), Ariomma, and Nomeidae (driftfishes); a lineage that includes Amarsipus carlsbergi (Amarsipa) as the sister lineage of a clade containing Tetragonurus (squaretails) and Chiasmodontidae (swallowers); a clade that includes Scombrolabrax heterolepis (Longfin Escolar), Lepidocybium flavobrunneum (Escolar), a paraphyletic Gempylidae (snake mackerels), and Trichiuridae (cutlassfishes); and a lineage containing Caristiidae (manefishes) and Bramidae (pomfrets) (Friedman et al. 2019; Arcila et al. 2021; Harrington et al. 2021; Ghezelayagh et al. 2022). The relationships of Scombridae (mackerels and tunas), Icosteus aenigmaticus (Ragfish), Pomatomus saltatrix (Bluefish), and Arripis (Australian Salmon) are not well resolved within Scombriformes; however, molecular analyses consistently resolve the lineages traditionally classified in Stromateoidei (e.g., Haedrich 1967; Horn 1984) as paraphyletic (Friedman et al. 2019; Arcila et al. 2021; Harrington et al. 2021; Ghezelayagh et al. 2022).

  • The monophyly of Scombriformes is not supported in a morphological analysis of 207 characters that resolves the lineages traditionally classified in Stromateoidei as a monophyletic group (Pastana et al. 2022). The paraphyly of Stromateoidei consistently resolved in molecular phylogenetic analyses (e.g., Betancur-R, Broughton, et al. 2013; Near et al. 2013; Arcila et al. 2021; Harrington et al. 2021; Ghezelayagh et al. 2022) is dismissed on the basis of the subjective assessment that morphological characters supporting stromateoid monophyly are “unparalleled and highly complex anatomical features unlikely to have evolved multiple times independently” (Pastana et al. 2022:957). There is a degree of uncertainty in the phylogenetic relationships among the major lineages of Scombriformes inferred from molecular data, including phylogenomic datasets (Arcila et al. 2021; Harrington et al. 2021); however, there is no analysis of character evolution for the traits offered as evidence for stromateoid monophyly that accommodates different models of trait evolution and uncertainty in the phylogenetic relationships of scombriforms and stromateoids.

  • Other morphological phylogenetic analyses have focused on lineages within Scombriformes that included the previous delimitations of Scombroidei and Stromateoidei (Collette, Potthoff, et al. 1984; Horn 1984; G. D. Johnson 1986; Doiuchi et al. 2004), Gempylidae and Trichiuridae (Gago 1997, 1998; Beckett et al. 2018b), and Chiasmodontidae (Melo 2009). A phylogenetic analysis of 29 morphological characters focused on Scombroidei resolves Lepidocybium flavobrunneum, long classified in Gempylidae, as the sister lineage of a clade containing all other Gempylidae and Trichiuridae (G. D. Johnson 1986), a result that is congruent with several molecular phylogenetic analyses (Friedman et al. 2019; Arcila et al. 2021; Harrington et al. 2021; Ghezelayagh et al. 2022). Phylogenetic analyses of DNA sequences from 13 mtDNA protein coding regions seems to resolve Gempylidae as monophyletic (Mthethwa et al. 2023a, 2023b), but this result is likely an artifact of limiting the outgroups to two species of Trichiuridae. There is no available family-group name to classify Lepidocybium flavobrunneum.

  • Composition. There are currently 287 living species of Scombriformes (Collette and Nauen 1983; Fricke et al. 2023) that include Amarsipus carlsbergi, Icosteus aenigmaticus, Lepidocybium flavobrunneum, Pomatomus saltatrix, Scombrolabrax heterolepis, and species classified in Ariomma, Arripis, Bramidae, Caristiidae, Centrolophidae, Chiasmodontidae, Gempylidae, Nomeidae, Scombridae, Stromateidae, Tetragonurus, and Trichiuridae. Fossil lineages of Scombriformes include the pan-stromateid †Pinichthys pulcher (Bannikov 1988), the pan-chiasmodontid †Bannikovichthys paelignus (Carnevale 2007), the pan-pomatomid †Carangopsis maximus (Agassiz 1835:42), the pan-trichiuroid †Argestichthys vysotzkyi (Prokofiev 2002b), and the pan-trichiurid †Anenchelum eocaenicum (Danilit'chenko 1962; Monsch and Bannikov 2011). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, six new living species of Scombriformes have been described, comprising 2.1% of the living species diversity in the clade (Fricke et al. 2023).

  • Diagnostic apomorphies. There are no known morphologicalapomorphiesforScombriformes.

  • Synonyms. Stromateoidei (B. Li et al. 2009, tbl. 4), Pelagia (Miya et al. 2013:2; Campbell et al. 2018:172), and Pelagiaria (Betancur-R et al. 2017:22; Campbell et al. 2018:173; Friedman et al. 2019:1) are ambiguous synonyms of Scombriformes.

  • Comments. The name Scombriformes was applied to (1) the paraphyletic group containing Carangidae, Scombridae, Stromateidae, and Xiphias (Woodward 1901:418), (2) expanded to include Trichiuridae, Coryphaena, and Luvarus (Goodrich 1909:462–468), (3) limited to Scombridae (Regan 1909b), and (4) the monophyletic group as presented here in the definition (Betancur-R, Broughton, et al. 2013; Davis et al. 2016; Betancur-R et al. 2017; Dornburg and Near 2021; Ghezelayagh et al. 2022).

  • The earliest fossil Scombriformes is the scombrid †Landanichthys from the Danian (66.0–61.7 Ma) of Angola (Friedman et al. 2019). Bayesian relaxed molecular clock analyses of Scombriformes result in an average posterior crown age estimate of 72.8 million years ago, with the credible interval ranging between 66.4 and 81.7 million years ago (Friedman et al. 2019).

  • img-z123-9_03.gif

    FIGURE 15.

    Phylogenetic relationships of the major living lineages and fossil taxa of Scombriformes and Syngnathiformes. Filled circles identify the common ancestor of clades with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z122-1_03.jpg

    Syngnathiformes P. Bleeker 1859:xv
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Pegasus volitans Linnaeus 1758, Mullus auratus Jordan and Gilbert 1882b, Callionymus curvicornis Valenciennes 1837 in Cuvier and Valenciennes (1837), Centriscus scutatus Linnaeus 1758, and Syngnathus acus Linnaeus 1758. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek σύµφῠσις (s̍Imfuːsiz) meaning grown together or fused, especially in reference to bones, and γνάθoς (n̍æθo͡Ʊz) meaning jaw. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 955.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, figs. S7–S9). See Figures 2 and 14 for the phylogenetic resolution of Syngnathiformes within Percomorpha and Figure 15 for a phylogeny of the living lineages and fossil taxa comprising Syngnathiformes. The phylogenetic placement of the fossil pan-pegasid †Rhamphosus follows Pietsch (1978), Bannikov (2014b), Carnevale et al. (2014), and Calzoni et al. (2023); the pan-aulostomoid †Eekaulostomus follows Cantalice and Alvarado-Ortega (2016); the pan-aulostomids †Eoaulostomus, †Jurgensenichthys, †Macroaulostomus, and †Synhypuralis follows Blot (1980) and Orr (1995); the pan-fistularid †Urosphen follows Orr (1995); the pan centriscoid †Gasterorhamphosus follows Orr (1995) and Friedman (2009); the pan-centriscids †Paraeoliscus and †Paramphisile follows Blot (1980), Friedman (2009), and Brownstein (2023); the pan-solenostomids †Calamostoma and †Solenorhynchus follows Bannikov and Carnevale (2017) and Brownstein (2023); the pan-syngnathid †Prosolenostomus follows Orr (1995), A. B. Wilson and Orr (2011), and Brownstein (2023); the pan-callionymid †Gilmourella follows Carnevale and Bannikov (2019); and the pan-dactylopterid †Pterygocephalus follows Bannikov (2014b) and Carnevale et al. (2014). The phylogenetic placements of †Eekaulostomus and †Prosolenostomus differ from those presented in other phylogenetic analyses (Murray 2022).

  • Phylogenetics. Reflecting earlier classifications (e.g., Goodrich 1909:410–416), Greenwood et al. (1966) placed many lineages of Syngnathiformes, including pipefishes and seahorses, in Gasterosteiformes along with sticklebacks (e.g., Gasterosteidae) and Indostomus (armored sticklebacks). This delimitation of Gasterosteiformes was corroborated with several putative morphological synapomorphies (Pietsch 1978; G. D. Johnson and Patterson 1993; Orr 1995; Britz and Johnson 2002; Wiley and Johnson 2010).

  • The first set of molecular phylogenetic analyses aimed at relationships with Percomorpha resolved lineages traditionally classified in Gasterosteiformes into three disparately related clades (W.-J. Chen et al. 2003; Miya et al. 2003; W. L. Smith and Wheeler 2004, 2006; Dettaï and Lecointre 2005; W. L. Smith and Craig 2007; Kawahara et al. 2008; B. Li et al. 2009). Subsequent molecular phylogenetic studies with a broad taxon sampling of percomorph lineages consistently resolved Syngnathiformes as a clade containing a paraphyletic Syngnathoidei (e.g., Pegasidae, Syngnathidae, and Centriscidae), Callionymidae (dragonets), Draconettidae (slope dragonets), Mullidae (goatfishes), and Dactylopteridae (flying gurnards) (Betancur-R, Broughton, et al. 2013; Near et al. 2013; Song et al. 2014; Betancur-R et al. 2017; Alfaro et al. 2018; Hughes et al. 2018; Roth et al. 2020; Ghezelayagh et al. 2022). Molecular phylogenetic analyses of Syngnathiformes consistently resolve two lineages: a clade of benthic lineages that contains Pegasidae (seamoths), Dactylopteridae, Draconettidae, Callionymidae, and Mullidae, and a clade of the longsnouted lineages Syngnathidae (seahorses and pipefishes), Solenostomus (ghost pipefishes), Centriscidae (shrimpfishes), Macrorhamphosus (snipefishes), Aulostomus (trumpetfishes), and Fistularia (cornetfishes) (Longo et al. 2017; Santaquiteria et al. 2021; Ghezelayagh et al. 2022). Several molecular phylogenetic studies have focused on resolving relationships within Syngnathidae (Hamilton et al. 2017; Longo et al. 2017; Santaquiteria et al. 2021; Stiller et al. 2022).

  • Composition. There are currently 690 living species of Syngnathiformes (Fricke et al. 2023) classified in Aulostomus, Callionymidae, Centriscidae, Dactylopteridae, Draconettidae, Fistularia, Macroramphosidae, Mullidae, Pegasidae, Solenostomus, and Syngnathidae. Fossil lineages of Syngnathiformes include the pan-pegasid †Rhamphosus rastrum (Volta 1796); the pan-aulostomoid †Eekaulostomus cuevasae (Cantalice and Alvarado-Ortega 2016); the pan-aulostomids †Eoaulostomus bolcensis, †Jurgensenichthys elongatus, †Macroaulostomus veronensis, and †Synhypuralis banister (Blot 1980); the pan-fistularid †Urosphen dubius (Blainville 1818); the pan-centriscid †Gasterorhamphosus zuppichinii (Sorbini 1981); the pan-centriscids †Paraeoliscus robinetae and †Paramphisile weileri (Blot 1980); the pan-solenostomids †Calamostoma breviculum and †Solenorhynchus elegans (Blainville 1818; Heckel 1853); the pan-syngnathid †Prosolenostomus lessinii (Blot 1980); the pan-callionymid †Gilmourella minuta (Carnevale and Bannikov 2019); and the pan-dactylopterid †Pterygocephalus paradoxus (Agassiz 1835). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, 51 new living species of Syngnathiformes have been described, comprising 7.4% of the living species diversity in the clade (Fricke et al. 2023).

  • Diagnostic apomorphies. There are no known morphological apomorphies for Syngnathiformes.

  • Synonyms. Syngnatharia (Betancur-R et al. 2017:22) is an ambiguous synonym of Syngnathiformes. Gasterosteiformes (Goodrich 1909:410–416; Greenwood et al. 1966:398; G. D. Johnson and Patterson 1993:580; J. S. Nelson 2006:308–316; Wiley and Johnson 2010:154) and Gobiesociformes (Wiley and Johnson 2010:162–163) are partial synonyms of Syngnathiformes.

  • Comments. Syngnathiformes was the name applied to the clade containing Aulostomus, Centriscidae, Fistularia, Macroramphosidae, Solenostomus, and Syngnathidae (McAllister 1968:111–114; J. S. Nelson 1984:249–253). In recent classifications of percomorphs, the name Syngnathiformes was applied for a more inclusive clade presented here in the definition (Davis et al. 2016; Betancur-R et al. 2017; Dornburg and Near 2021; Ghezelayagh et al. 2022). The name Syngnathiformes was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest fossil Syngnathiformes is the pan-centriscoid †Gasterorhamphosus zuppichinii from the Campanian and Maastrichtian (83.6–66.0 Ma) of Italy. Bayesian relaxed molecular clock analyses of Syngnathiformes result in an average posterior crown age estimate of 104.6 million years ago, with the credible interval ranging between 95.0 and 114.8 million years ago (Ghezelayagh et al. 2022).

  • img-z125-8_03.gif

    Ovalentaria W. L. Smith and T. J. Near in
    Wainwright et al. 2012
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Ambassis urotaenia Bleeker 1852, Mugil cephalus Linnaeus 1758, Embiotoca lateralis Agassiz 1854, Pseudochromis fridmani Klausewitz 1968, Gobiesox maeandricus (C. Girard 1858a), Gillellus semicinctus Gilbert 1890, Polycentrus schomburgkii Müller and Troschel 1848, Pholidichthys leucotaenia Bleeker 1856, Cichla temensis Humboldt in Humboldt and Valenciennes 1821, Labidesthes sicculus (Cope 1865), Gambusia affinis (Baird and Girard 1853), and Oryzias latipes (Temminck and Schlegel 1846). This is a minimum-crown-clade definition.

  • Etymology. From the Latin ovum meaning egg and lentae meaning sticky or tenacious.

  • Registration number. 998.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 10 concatenated Sanger-sequenced nuclear genes (Wainwright et al. 2012, fig. 2). The phylogenetic resolution of Ovalentaria within Percomorpha is presented in Figure 2, and the phylogenetic relationships of the major lineages of Ovalentaria are presented in Figures 14 and 16.

  • Phylogenetics. Monophyly of Ovalentaria was discovered in early molecular analyses aimed at resolving relationships within Percomorpha (W.-J. Chen et al. 2003; Dettaï and Lecointre 2005; Miya et al. 2005; W. L. Smith and Wheeler 2006; Mabuchi et al. 2007; W. L. Smith and Craig 2007; Kawahara et al. 2008; Setiamarga et al. 2008). A phylogenetic analysis of DNA sequences from four nuclear genes resolved a clade comprising Mugilidae (mullets), Plesiopidae (roundheads), Blennioidei (blennies), Atheriniformes (silversides, needlefishes, and killifishes), Cichlidae (cichlids), Gobiesocidae (clingfishes), and Pomacentridae (damselfishes) (B. Li et al. 2009). A subsequent analysis of 10 exons expanded the clade to include Polycentridae (leaffishes), Pholidichthys (engineer blennies), Embiotocidae (surfperches), Congrogadidae (eel blennies), Pseudochromidae (dottybacks), Gramma and Lipogramma (basslets), and Opistognathidae (jawfishes) (Wainwright et al. 2012). Subsequent molecular analyses consistently support the monophyly of Ovalentaria (Betancur-R, Broughton, et al. 2013; Near et al. 2013; Collins et al. 2015; Betancur-R et al. 2017; Alfaro et al. 2018; Hughes et al. 2018; Ghezelayagh et al. 2022; Mu et al. 2022). Initially, the monophyly of Ovalentaria was discussed in the context of the presence of demersal eggs with adhesive filaments that characterizes many of the lineages in the clade (Breder and Rosen 1966; Semple 1985; Mooi 1990; Wirtz 1993; Britz 1997; Breining and Britz 2000).

  • A morphological phylogenetic analysis of Ovalentaria based on 38 characters scored from the caudal skeleton did not include other percomorph lineages and therefore did not test monophyly of the clade (Thieme et al. 2022). Relationships within Ovalentaria differed from molecular phylogenetic analyses in that Gramma and Lipogramma were resolved as a clade, Pholidichthys and Cichlidae were not resolved as sister lineages, Gobiesocidae and Blennioidei did not form a monophyletic group, and both Blennioidei and Atheriniformes were resolved as paraphyletic (Thieme et al. 2022).

  • Composition. There are currently 5,940 living species of Ovalentaria (Fricke et al. 2023) classified in Atheriniformes and Blenniiformes. Over the past 10 years, 527 new living species of Ovalentaria have been described (Fricke et al. 2023), comprising approximately 8.9% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Ovalentaria are currently limited to features of the caudal skeleton and include (1) fusion of two ural centra to form the compound centrum during development (Thieme et al. 2022), and (2) second uroneural present (Thieme et al. 2022).

  • Synonyms. Blenniiformes (Dornburg and Near 2021; Ghezelayagh et al. 2022) and Ovalentariae (Betancur-R, Broughton, et al. 2013:13) are ambiguous synonyms of Ovalentaria. Stiassnyiformes (B. Li et al. 2009, tbl. 4) is a partial synonym of Ovalentaria.

  • Comments. Ovalentaria is one of the most species-rich named clades of Percomorpha and similarly to nearly every percomorph clade was discovered primarily through molecular phylogenetic analyses (e.g., Wainwright et al. 2012). A study examining the morphology of the caudal skeleton in Ovalentaria illustrates the potential of applying molecular inferred phylogenies with novel results to understanding phenotypic evolution in large inclusive clades of Percomorpha (Thieme et al. 2022). The name Ovalentaria was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • Bayesian relaxed molecular clock analyses of Ovalentaria result in an average posterior crown age estimate of 96.2 million years ago, with the credible interval ranging between 86.2 and 106.0 million years ago (Ghezelayagh et al. 2022).

  • img-z128-3_03.gif

    FIGURE 16.

    Phylogenetic relationships of the major living lineages and fossil taxa of Ovalentaria, Atheriniformes, Atherinoidei, Belonoidei, Cyprinodontoidei, Blenniiformes, and Blennioidei. Filled circles identify the common ancestor of clades with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z127-1_03.jpg

    Atheriniformes J. Ferrer Aledo 1930:245

  • Definition. The least inclusive crown clade that contains Oryzias latipes (Temminck and Schlegel 1846), Atherina hepsetus Linnaeus 1758, and Cyprinodon variegatus Lacépède 1803. This is a minimum-crown-clade definition, but the clade is not defined using the PhyloCode.

  • Etymology. From the ancient Greek ἀθερίνη (æθɚɹˈiːnə), which is the name used by ancient authors (e.g., Aristotle and Oppian) in reference to the Mediterranean Sand Smelt, Atherina hepsetus Linnaeus (D. W. Thompson 1947:3–4).

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S10). Although Atherina hepsetus is not included in the reference phylogeny, it resolves in a clade with other species of Atherina in phylogenetic analyses of Sanger-sequenced mitochondrial and nuclear genes (Sparks and Smith 2004b, fig. 2; Astolfi et al. 2005, fig. 2; Francisco et al. 2008, fig. 2, 2011, fig. 2; Heras and Roldán 2011, fig. 2; Campanella et al. 2015, fig. 2B). See Figure 16 for a phylogeny of the lineages comprising Atheriniformes.

  • Phylogenetics. The delimitation of Atheriniformes that includes Atherinoidei, Belonoidei, and Cyprinodontoidei was first proposed in a pre-Hennigian study of osteology, musculature, and reproductive characters that aimed toward the identification of “a phylogenetically natural group” (Rosen 1964:260). Since this work, the monophyly of Atheriniformes has not been challenged. Analysis of the gill arch skeleton and hyoid apparatus led to the reclassification of Adrianichthyidae (ricefishes) from Cyprinodontoidei to Belonoidei, the discovery that Atherinoidei was not diagnosed by morphological synapomorphies, and the resolution of Belonoidei and Cyprinodontoidei as sister lineages (Rosen and Parenti 1981). Subsequent phylogenetic analyses using morphological characters consistently supported the monophyly of Atheriniformes and the sister lineage relationship between Belonoidei and Cyprinodontoidei (White 1985; Stiassny 1990; L. R. Parenti 1993, 2005; Saeed et al. 1994; Dyer and Chernoff 1996; Dyer H 2006).

  • Several of the earliest molecular phylogenies of Percomorpha resolved Atheriniformes as paraphyletic as a result of the placement of other lineages of Ovalentaria (W.-J. Chen et al. 2003; Dettaï and Lecointre 2005; Miya et al. 2005), but subsequent molecular phylogenetic studies support atheriniform monophyly (Mabuchi et al. 2007; Kawahara et al. 2008; Setiamarga et al. 2008; B. Li et al. 2009; Near, Eytan, et al. 2012; Wainwright et al. 2012; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Eytan et al. 2015; Davis et al. 2016; Smith et al. 2016; Betancur-R et al. 2017; Hughes et al. 2018; Rabosky et al. 2018; Ghezelayagh et al. 2022). Within Atheriniformes, molecular phylogenies have resolved all three possible relationships among Atherinoidei, Belonoidei, and Cyprinodontoidei: Belonoidei and Cyprinodontoidei as sister lineages (Miya et al. 2005; Mabuchi et al. 2007; Kawahara et al. 2008; Setiamarga et al. 2008; Wainwright et al. 2012; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Smith et al. 2016; Hughes et al. 2018; Rabosky et al. 2018), Atherinoidei and Belonoidei as sister lineages (B. Li et al. 2009; Eytan et al. 2015; Ghezelayagh et al. 2022), and Atherinoidei and Cyprinodontoidei as sister lineages (Davis et al. 2016; Betancur-R et al. 2017). Node support for these relationships is typically low, and changes in taxon sampling for similar sets of sampled genes seem to affect the resolution of relationships within Atheriniformes (e.g., Betancur-R, Broughton, et al. 2013; Betancur-R et al. 2017). While there is strong support from both morphological and molecular data for monophyly of Atheriniformes, there remains uncertainty in the relationships among Atherinoidei, Belonoidei, and Cyprinodontoidei.

  • Composition. There are currently 2,126 living species of Atheriniformes (Fricke et al. 2023) classified in Atherinoidei, Belonoidei, and Cyprinodontoidei. Over the past 10 years, 259 new living species of Atheriniformes have been described (Fricke et al. 2023), comprising approximately 12.2% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Atheriniformes include (1) separation of afferent and efferent circulation during development (Rosen and Parenti 1981; L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (2) a restricted lobular type of testis, in which the spermatogonia are present at the lobule ends only rather than throughout the entire length of the lobule (Rosen and Parenti 1981; L. R. Parenti 1993, 2005; L. R. Parenti and Grier 2004; Wiley and Johnson 2010; Uribe et al. 2014), (3) protrusible upper jaw mechanism with palatomaxillary ligaments crossed and with maxillary ligament to the cranium (Rosen and Parenti 1981), (4) dermal and endochondral disclike ethmoid ossifications (Rosen and Parenti 1981; L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (5) medial hooklike projection and ventral flange on fifth ceratobranchial (Stiassny 1990; L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (6) supraneurals absent (Stiassny 1990; L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (7) infraorbital series consisting of lacrimal, dermosphenotic, and two or fewer anterior infraorbital bones (Rosen and Parenti 1981; L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (8) dorsal portion of gill arch with large fourth epibranchial as the supporting bone (Rosen and Parenti 1981; L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (9) fourth pharyngobranchial absent in dorsal gill arch (Rosen and Parenti 1981; L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (10) saccus vasculosus absent (Tsuneki 1992; L. R. Parenti 2005; Wiley and Johnson 2010), (11) coupling during mating (L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (12) distal end of pleural rib and lateral process of pelvic bone in close association and sometimes attached with a ligament (L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (13) supracleithrum reduced or absent (L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (14) superficial (A1) division of adductor mandibulae with two tendons, one inserting on maxilla, second inserting on lacrimal (L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (15) olfactory sensory epithelium arranged in sensory islets (L. R. Parenti 1993, 2005; Wiley and Johnson 2010), (16) fluid (rather than granular) egg yolk (L. R. Parenti and Grier 2004; L. R. Parenti 2005; Wiley and Johnson 2010), (17) caudal fin supported by three preural centra (Thieme et al. 2022), (18) lower hypural plate fused with compound centrum (Thieme et al. 2022), (19) uroneural fused with compound centrum (Thieme et al. 2022), (20) haemal arch of preural centrum 2 fused with its centrum (Thieme et al. 2022), and (21) absence of interhaemal spine cartilage 2 (Thieme et al. 2022).

  • Synonyms. Atherinomorpha (Greenwood et al. 1966:397; Rosen 1973:510, fig. 129; Rosen and Parenti 1981:23; J. S. Nelson et al. 2016:353–354) and Atherinomorphae (Wiley and Johnson 2010:154; Betancur-R, Broughton, et al. 2013, fig. 8; Betancur-R et al. 2017:25) are ambiguous synonyms of Atheriniformes.

  • Comments. Atheriniformes was applied as the name of a taxonomic group that included species classified in Atherinoidei, Belonoidei, and Cyprinodontoidei (Rosen 1964; Greenwood et al. 1966:397–398).

  • The earliest fossils of Atheriniformes are all from the Ypresian (56.0–48.1 Ma) of Italy and include the pan-exocoetid †Rhamphexocoetus (Appendix 1; Bannikov et al. 1985) and the atherinoid †Latellagnathus (Bannikov et al. 1985; Bannikov 2008, 2014b; Carnevale et al. 2014). Bayesian relaxed molecular clock analyses of Atheriniformes result in an average posterior crown age estimate of 87.5 million years ago, with the credible interval ranging between 76.8 and 97.2 million years ago (Ghezelayagh et al. 2022).

  • img-z129-8_03.gif

    Atherinoidei P. Bleeker 1859:xxiv

  • Definition. The least inclusive crown clade that contains Atherinella panamensis Steindachner 1875, Atherinopsis californiensis C. Girard 1854, Atherina hepsetus Linnaeus 1758, and Atherion elymus Jordan and Starks 1901. This is a minimum-crown-clade definition, but the clade is not defined using the PhyloCode.

  • Etymology. From the ancient Greek ἀθερίνη (æθɚɹˈiːnə), which is the name for the Mediterranean Sand Smelt, Atherina hepsetus Linnaeus, used by Aristotle and Oppian (D. W. Thompson 1947:3–4).

  • Reference phylogeny. A phylogeny inferred from a dataset comprising eight Sanger-sequenced mtDNA and nuclear genes (Campanella et al. 2015, fig. 2). Although Atherina hepsetus is not included in the reference phylogeny, it resolves in a clade with other species of Atherina in phylogenetic analyses of Sanger-sequenced mitochondrial and nuclear genes (e.g., Campanella et al. 2015, fig. 2B). Phylogenetic relationships of the major lineages of Atherinoidei are presented in Figure 16.

  • Phylogenetics. Subsequent to the delimitation of Atherinoidei (Rosen 1964), several morphological studies did not support the monophyly of the group (Rosen and Parenti 1981; L. R. Parenti 1984, 1989, 1993; Ivantsoff et al. 1987; Saeed et al. 1994); however, studies based on adult and larval morphology provided evidence for the monophyly of the lineage (White et al. 1984; Dyer and Chernoff 1996; Aarn and Ivantsoff 1997). Molecular phylogenetic studies consistently resolve Atherinoidei as monophyletic (Setiamarga et al. 2008; Bloom et al. 2012; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Campanella et al. 2015; Betancur-R et al. 2017; Hughes et al. 2018; Rabosky et al. 2018; Ghezelayagh et al. 2022).

  • Morphological and molecular phylogenetic analyses are congruent in resolving Atherinopsidae (New World silversides) as the sister lineage of all other Atherinoidei (Aarn and Ivantsoff 1997; Bloom et al. 2012; Near et al. 2013; Campanella et al. 2015; Betancur-R et al. 2017; Ghezelayagh et al. 2022). One area of incongruence among phylogenetic analyses is the support for Notocheirus hubbsi (Surf Silverside) and species of Iso (surf sardines) as sister lineages in a morphological study (Dyer and Chernoff 1996); however, Notocheirus is nested well within Atherinopsidae and Iso is resolved as the sister lineage of a clade containing Atherinidae (silversides), Bedotiidae (Madagascar rainbowfishes), Melanotaeniidae (rainbowfishes), Telmatherinidae (Celebes rainbowfishes), and Pseudomugilidae (blue eyes) in molecular phylogenies (Bloom et al. 2012, 2013; Campanella et al. 2015; Rabosky et al. 2018; Ghezelayagh et al. 2022).

  • Phylogenetic analyses of Sanger-sequenced mtDNA and nuclear genes result in the paraphyly of Melanotaeniidae because Cairnsichthys is resolved as the sister lineage all other sampled species of Telmatherinidae and Pseudomugilidae (Bloom et al. 2012; Campanella et al. 2015; Rabosky et al. 2018); however, Melanotaeniidae is monophyletic in morphological and phylogenomic analyses (Aarn and Ivantsoff 1997; Aarn et al. 1998; Ghezelayagh et al. 2022). Following conclusions from a morphological phylogenetic analysis (Dyer and Chernoff 1996), Nelson et al. (2016:358–360) treat Bedotiidae, Pseudomugilidae, and Telmatherinidae as lineages of Melanotaeniidae.

  • A molecular phylogeny resolves Atherion (pricklenose silversides) and Phallostethidae (priapiumfishes) as sister lineages (Campanella et al. 2015). To date, there are no molecular data available for Dentatherina merceri (Mercer's Tusked Silverside), but several morphological studies place it as the sister lineage of Phallostethidae (L. R. Parenti 1984; Dyer and Chernoff 1996; Aarn and Ivantsoff 1997). This has prompted the classification of D. merceri in Phallostethidae (Dyer and Chernoff 1996; Aarn and Ivantsoff 1997; J. S. Nelson 2006:273); however, the group name Dentatherinidae, including only D. merceri, is endorsed by others (Ivantsoff et al. 1987; J. S. Nelson et al. 2016:360–361). Because D. merceri is convincingly resolved as the sister lineage of a clade containing all other priapiumfishes, we include it in Phallostethidae as an optimal reflection of phylogenetic relationships and an effort to reduce redundant group names in the classification of ray-finned fishes.

  • Composition. There are currently 385 living species of Atherinoidei (Fricke et al. 2023) classified in Atherinidae, Atherinopsidae, Atherion, Bedotiidae, Iso, Melanotaeniidae, Phallostethidae, Pseudomugilidae, and Telmatherinidae. Over the past 10 years, there have been 38 new species of Atherinoidei described (Fricke et al. 2023), comprising 9.9% of the species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Atherinoidei include (1) preanal length of flexion larvae short, approximately 33% of body length (White et al. 1984; Wiley and Johnson 2010), (2) single row of melanophores on dorsal midline of larvae (White et al. 1984; L. R. Parenti 2005; Wiley and Johnson 2010), (3) ventral face of vomer concave (Dyer and Chernoff 1996; Wiley and Johnson 2010), (4) adductor mandibulae A1 with long tendon to lacrimal (Dyer and Chernoff 1996; Wiley and Johnson 2010), (5) two anterior infraorbital bones (Dyer and Chernoff 1996; L. R. Parenti 2005; Wiley and Johnson 2010), (6) presence of pelvic rib ligament (Dyer and Chernoff 1996; Wiley and Johnson 2010), (7) pelvic plate does not extend to anterior tip of longitudinal shaft (Dyer and Chernoff 1996; Wiley and Johnson 2010), and (8) presence of a flexible second dorsal fin (Dyer and Chernoff 1996; Wiley and Johnson 2010).

  • Synonyms. Atherinidae (Jordan and Hubbs 1919:12–19; Schultz 1948:2–3) and Atheriniformes(Saeedetal.1994:47–48;DyerandChernoff 1996, tbl. 1; Wiley and Johnson 2010:155; Betancur-R, Broughton, et al. 2013, fig. 3; J. S. Nelson et al. 2016:354–355; Betancur-R et al. 2017:25) are ambiguous synonyms of Atherinoidei.

  • Comments. In the mid-20th century, Atherinoidei was applied as the name of a group containing Atherinidae, Bedotiidae, Isonidae, Melanotaeniidae, Phallostethidae, and Pseudomugilidae (Rosen 1964; Greenwood et al. 1966).

  • The earliest fossil taxa of Atherinoidei are all from the Ypresian (56.0–48.1 Ma) and include the otolith taxon †‘Atherinidarum’ from France and India (Nolf 1988; Nolf et al. 2006) and the skeletal fossils †Rhamphognathus, †Latellagnathus, and †Mesogaster from Italy (Bannikov 2008, 2014b; Carnevale et al. 2014). Bayesian relaxed molecular clock analyses of Atherinoidei result in an average posterior crown age estimate of 71.0 million years ago, with the credible interval ranging between 56.5 and 84.4 million years ago (Ghezelayagh et al. 2022).

  • img-z131-7_03.gif

    Belonoidei E. Postel 1959:150

  • Definition. The least inclusive crown clade that contains Adrianichthys oophorus (Kottelat 1990), Xenentodon cancila (Hamilton 1822), Hemiramphus far (Fabricius in Niebuhr 1775), and Belone belone (Linnaeus 1761). This is a minimum-crown-clade definition, but the clade is not defined using the PhyloCode.

  • Etymology. From the ancient Greek βελόνη (bƗl̍αːne͡I) meaning needle, but also the name applied to the Greater Pipefish (Syngnathus acus) and the Garfish (Belone belone) in the biological writings of Aristotle (D. W. Thompson 1947:29–32).

  • Reference phylogeny. A phylogeny of 123 species of Belonoidei inferred from a supermatrix of 27 nuclear and mitochondrial genes (Rabosky et al. 2018; J. Chang et al. 2019). The phylogeny is available on the Dryad data repository (Rabosky et al. 2019). Phylogenetic relationships among the major lineages of Belonoidei are presented in Figure 16. The placement of †Rhamphexocoetus in the phylogeny is on the basis of inferences from morphology (Bannikov et al. 1985; Benton et al. 2015).

  • Phylogenetics. Hemiramphidae (halfbeaks), Exocoetidae (flyingfishes), Belonidae (needlefishes), and Scomberesocidae (sauries) were grouped together in early 20th-century classifications (Schlesinger 1909; Regan 1911f). Phylogenetic analysis of morphology led to a delimitation of Belonoidei that includes those lineages plus Adrianichthyidae (ricefishes) (Rosen and Parenti 1981). Subsequent morphological and molecular studies provided additional support for the monophyly of Belonoidei and for the resolution of Adrianichthyidae as the sister lineage to all other belonoids (Collette, McGowen, et al. 1984; L. R. Parenti 1987, 1993, 2008; Miya et al. 2003, 2005; Kawahara et al. 2008; Setiamarga et al. 2008; Near, Eytan, et al. 2012; Wainwright et al. 2012; Near et al. 2013; Davis et al. 2016; Smith et al. 2016; Betancur-R et al. 2017; Hughes et al. 2018; Ghezelayagh et al. 2022; Ding et al. 2023).

  • Phylogenetic analyses of morphological and molecular data motivated changes to the traditional classification of Belonoidei (Lovejoy et al. 2004; Aschliman et al. 2005), but current classifications continue to include paraphyletic groups (J. S. Nelson et al. 2016:363–370; Betancur-R et al. 2017). For example, Belonidae as traditionally delimited is paraphyletic because species classified in Scomberesocidae (sauries), Cololabis, and Scomberesox are nested within Belonidae as the sister lineage of Belone (Lovejoy 2000; Lovejoy et al. 2004; Daane et al. 2021). Cololabis and Scomberesox are now placed in Belonidae (Betancur-R et al. 2017). Hemiramphidae is resolved as paraphyletic in both morphological (Tibbetts 1991; Aschliman et al. 2005) and molecular phylogenetic analyses (Lovejoy 2000; Lovejoy et al. 2004; Betancur-R et al. 2017; Daane et al. 2021; Ghezelayagh et al. 2022; Ding et al. 2023). The paraphyly of Hemiramphidae led to the recognition of Zenarchopteridae (viviparous halfbeaks) as a separate Linnean-ranked taxonomic family (Lovejoy et al. 2004); however, the remaining lineages of Hemiramphidae are paraphyletic relative to Exocoetidae. Three lineages comprise the current delimitation of Hemiramphidae (Lovejoy 2000; Lovejoy et al. 2004; Daane et al. 2021): a clade we refer to with the informal name hyporhamphids that contains Arrhamphus, Chriodorus atherinoides, Hyporhamphus, and Melapedalion breve for which there is no available family-group name (Van der Laan et al. 2014:77); Euleptorhamphus and Rhynchorhamphus that we delimit as Euleptorhamphidae, an elevation of Euleptorhamphinae (Fowler 1934:323); and Hemiramphidae that is limited here to species of Hemiramphus and Oxyporhamphus. The more exclusive Hemiramphidae and Exocoetidae are consistently resolved as sister lineages in molecular phylogenetic analyses (Lovejoy 2000; Lovejoy et al. 2004; Betancur-R et al. 2017; Daane et al. 2021; Ghezelayagh et al. 2022). Morphological phylogenetic analyses result in the resolution of most lineages traditionally classified in Hemiramphidae in a large polytomy with a clade containing Oxyporhamphus and Exocoetidae (Tibbetts1991; Aschliman et al. 2005).

  • Composition. There are currently 292 living species of Belonoidei (Collette 2003, 2004a, 2004b; Bemis and Collette 2019; Collette and Bemis 2019a, 2019b, 2019c; Parin et al. 2019; Fricke et al. 2023) that include Arrhamphus sclerolepis, Chriodorus atherinoides, Melapedalion breve, and species classified in Adrianichthyidae, Belonidae, Euleptorhamphidae, Exocoetidae, Hemiramphidae, Hyporhamphus, and Zenarchopteridae. Fossil Belonoidei include the pan-exocoetid †Rhamphexocoetus volans (Appendix 2; Bannikov et al. 1985). Over the past 10 years, 25 new species of Belonoidei have been described (Fricke et al. 2023), comprising 8.6% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Belonoidei include (1) interarcual cartilage absent (Rosen and Parenti 1981; L. R. Parenti 2005, 2008; Wiley and Johnson 2010), (2) relatively small second and third epibranchials (Rosen and Parenti 1981; L. R. Parenti 2005, 2008; Wiley and Johnson 2010), (3) vertically reoriented second pharyngobranchial (Rosen and Parenti 1981; L. R. Parenti 2005, 2008; Wiley and Johnson 2010), (4) dorsal hypohyal absent (Rosen and Parenti 1981; L. R. Parenti 2005, 2008; Wiley and Johnson 2010), (5) interhyal absent (Rosen and Parenti 1981; L. R. Parenti 1987, 2005, 2008; Wiley and Johnson 2010), (6) upper lobe of caudal fin with fewer principal fin rays than lower lobe (Rosen and Parenti 1981; L. R. Parenti 2005, 2008; Wiley and Johnson 2010), and (7) parietals extremely small or absent (L. R. Parenti 2008; Wiley and Johnson 2010).

  • Synonyms. Beloniformes (Rosen and Parenti 1981:23; Wiley and Johnson 2010:156; Betancur-R, Broughton, et al. 2013, fig. 8; J. S. Nelson et al. 2016:363–370; Betancur-R et al. 2017:25) is an ambiguous synonym of Belonoidei. Synentognathi (Regan 1911f:331–335) and Exocoetoidei (Greenwood et al. 1966:397) are partial synonyms of Belonoidei.

  • Comments. In earlier classifications, Belonoidei is used as a group name for all belonoids to the exclusion of Adrianichthyidae (Nelson 1994:266; Betancur-R et al. 2017). The earliest fossil taxa of Belonoidei are all from the Ypresian (56.0–48.1 Ma) of Italy and include the pan-exocoetid †Rhamphexocoetus and the taxa †“Engraulisevolans and †“Hemiramphusedwardsi of uncertain phylogenetic resolution within the clade (Bannikov et al. 1985; Bannikov 2014b; Carnevale et al. 2014). Bayesian relaxed molecular clock analyses result in an average posterior crown age estimate of the Belonoidei crown of 68.0 million years ago, with the credible interval ranging between 56.8 and 81.2 million years ago (Ghezelayagh et al. 2022). Euleptorhamphidae is a valid family-group name under the International Code of Zoological Nomenclature (Van der Laan et al. 2014:77).

  • img-z133-3_03.gif

    Cyprinodontoidei P. Bleeker 1859:xxix

  • Definition. The least inclusive crown clade that contains Cyprinodon variegatus Lacépède 1803, Poecilia velifera (Regan 1914a), Pantanodon stuhlmanni (Ahl 1924), Austrolebias nigripinnis (Regan 1912f), and Aplocheilus lineatus (Valenciennes in Cuvier and Valenciennes 1846). This is a minimum-crown-clade definition, but the clade is not defined using the PhyloCode.

  • Etymology. From the ancient Greek κυπρῖνος (kuːpɹˈiːno͡ʊz), frequently applied to the Eurasian Carp, Cyprinus carpio (D. W. Thompson 1947:135–136), and δών (̍o͡Ʊdαːn) meaning tooth.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 295 genes captured using anchored hybrid enrichment (Piller et al. 2022, figs. 3–7). Although Cyprinodon variegatus is not included in the reference phylogeny, it resolves in a clade with other species of Cyprinodon in phylogenetic analyses of mtDNA (Echelle et al. 2005, 2006; Martin and Wainwright 2011). Phylogenetic relationships of the major living and fossil lineages of Cyprinodontoidei are presented in Figure 16. The phylogenetic resolutions of the pan-rivulid †Kenyaichthys, the pan-orestiid †Carrionellus, and the pan-valenciids †Francolebias and †Prolebias are on the basis of inferences from morphology (Costa 2011, 2012b; Altner and Reichenbacher 2015).

  • Phylogenetics. The lineages that comprise Cyprinodontoidei were grouped together in many pre-Hennigian classifications of teleost fishes (e.g., Garman 1895), but were thought to be related to such disparate lineages as Esocidae and Amblyopsidae (T. N. Gill 1872; Boulenger 1904a; Goodrich 1909:400–401; Regan 1909a, 1911g; Hubbs 1924; Gosline 1963a). Subsequent studies identified Atheriniformes as a clade containing Cyprinodontoidei, Atherinoidei, and Belonoidei (Rosen 1964; Greenwood et al. 1966; Rosen and Parenti 1981). Morphological and molecular phylogenetic analyses of relationships within Cyprinodontoidei are broadly congruent in supporting monophyly of the lineage and the resolution of two clades: the aplocheiloids and cyprinodontoids (L. R. Parenti 1981; Costa 1998, 2012a, 2012b; Hertwig 2008; Pohl et al. 2015; Helmstetter et al. 2016; Costa et al. 2017; Reznick et al. 2017; Amorim and Costa 2018; Bragança et al. 2018; Ghezelayagh et al. 2022; Piller et al. 2022).

  • Relationships among the aplocheiloids, including Aplocheilidae (Asian rivulines), Nothobranchiidae (African rivulines), and Rivulidae (New World rivulines), vary among different phylogenetic analyses. Studies using mtDNA and morphology resolve the traditional delimitation of Aplocheilidae (e.g., L. R. Parenti 1981) as paraphyletic (Murphy and Collier 1997; Costa 2004, 2012a, 2012b), with African lineages and the South American Rivulidae forming a clade that is the sister lineage of the Asian-Malagasy Aplocheilidae (sensu stricto). Because of the apparent paraphyly of Aplocheilidae, the African aplocheiloid lineages are now classified in Nothobranchiidae (Costa 2004, 2016). Subsequent phylogenetic analyses of morphology (Hertwig 2008), Sanger-sequenced mtDNA and nuclear genes (Pohl et al. 2015; Costa et al. 2017; Reznick et al. 2017; Amorim and Costa 2018; Bragança et al. 2018), and phylogenomic datasets (Ghezelayagh et al. 2022; Piller et al. 2022) resolve Rivulidae as the sister lineage of a clade containing Aplocheilidae and Nothobranchiidae.

  • Molecular phylogenies resolve Pantanodon (spine killifishes) as the sister lineage of all other cyprinodontoids (Pohl et al. 2015; Bragança et al. 2018; Piller et al. 2022). The remaining cyprinodontoid lineages resolve into three clades (Amorim and Costa 2018; Bragança and Costa 2019; Ghezelayagh et al. 2022; Piller et al. 2022): (1) Cubanichthyidae (Caribbean killifishes), Cyprinodontidae (pupfishes), Fundulidae (topminnows), Goodeidae (goodeids), and Profundulidae (Middle American killifishes) (Webb et al. 2004; Reznick et al. 2017); (2) Anablepidae (four-eyed fishes), Fluviphylax (American lampeyes), and Poeciliidae (livebearers) (Reznick et al. 2017; Bragança and Costa 2018); and (3) Aphaniidae (Asian killifishes), Procatopodidae (African lampeyes), Orestiidae (Andean pupfishes), and Valencia (toothcarps) (A. Parker and Kornfield 1995; Pohl et al. 2015; Helmstetter et al. 2016; Reznick et al. 2017; Bragança and Costa 2019). The traditional delimitations of Cyprinodontidae and Poeciliidae (L. R. Parenti 1981; Ghedotti 2000) are paraphyletic. Fluviphylax, Pantanodon, and Procatopodidae do not share common ancestry with Poeciliidae; Aphaniidae, Cubanichthyidae, and Orestiidae are distantly related to Cyprinodontidae (Freyhof et al. 2017; Bragança and Costa 2019; Piller et al. 2022).

  • Composition. There are currently 1,449 living species of Cyprinodontoidei (Fricke et al. 2023) classified in Aplocheilidae, Nothobranchiidae, Rivulidae, Anablepidae, Aphaniidae, Cubanichthyidae, Cyprinodontidae, Fluviphylax, Fundulidae, Goodeidae, Orestiidae, Pantanodontidae (Meinema and Huber 2023), Poeciliidae, Profundulidae, Procatopodidae, and Valencia. Fossil taxa include the pan-rivulid †Kenyaichthys kipkechi (Altner and Reichenbacher 2015), the pan-orestiid †Carrionellus diumortuus (Costa 2011), and the pan-valenciids †Francolebias aymardi and †Prolebias stenoura (J. Gaudant 1988; Costa 2012b). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, 196 new species of Cyprinodontoidei have been described (Fricke et al. 2023), comprising 13.5% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Cyprinodontoidei include (1) caudal fin endoskeleton with one epural symmetrically opposing parhypural (L. R. Parenti 1981; Rosen and Parenti 1981; Costa 2012a), (2) caudal fin unlobed, truncate or rounded (L. R. Parenti 1981; Rosen and Parenti 1981; Costa 1998), (3) first rib attached to second rather than third vertebra (L. R. Parenti 1981; Costa 1998), (4) pectoral fin set low on body, with large scalelike postcleithrum (L. R. Parenti 1981; Rosen and Parenti 1981; Costa 1998), (5) elongate interarcual cartilage joining expanded base of first epibranchial with shaft of second pharyngobranchial (Rosen and Parenti 1981), (6) presence of anterior expansion on the alveolar arm of premaxilla (Costa 1998), (7) tendon of the A1 division of adductor mandibulae attached to the lacrimal (Costa 1998; Hertwig 2008), (8) dorsal edge of mesopterygoid reduced (Costa 1998), (9) urohyal deep (Costa 1998), (10) ventral process of lateral portion of epibranchial 2 absent (Costa 1998), (11) mesethmoid slightly anterior to lateral ethmoid (Costa 1998), (12) anteromedial process of pelvic girdle absent (Costa 1998), (13) subdivision of A1 adductor mandibulae into two heads (Hertwig 2008), (14) many muscle fibers arising from the tendon or aponeurosis of the A2 /A3 adductor mandibulae (Hertwig 2008), (15) the A2 /A3 adductor mandibulae subdivided into three distinct heads by the ramus mandibularis (Hertwig 2008), (16) a separate section of the adductor mandibulae (AωQ) originates with a single tendon from the medial face of the quadrate (Hertwig 2008), (17) adductor arcus palatini inserts medially on the mesopterygoid (Hertwig 2008), (18) epural with bladelike shape (Costa 2012a), (19) caudal fin rays continuously arranged between upper and lower hypural plates (Costa 2012a), (20) distal tip of well-developed preural vertebra 2 acting in support of caudal fin rays (Costa 2012a), (21) stegural minute (Costa 2012a), (22) neural spine of preural vertebra 2 wider than neural spines of preural vertebrae 4 and 5 (Costa 2012a), and (23) complete ankylosis of upper hypurals and compound caudal centrum (Costa 2012a).

  • Synonyms. Cyprinodontiformes (L. R. Parenti 1981:462–463, 1993, tbl. 2; Rosen and Parenti 1981:23; Wiley and Johnson 2010:157; Betancur-R, Broughton, et al. 2013, fig. 8; J. S. Nelson et al. 2016:369–380; Betancur-R et al. 2017:26) is an ambiguous synonym of Cyprinodontoidei. Microcyprini (Regan 1911g:321–322) is a partial synonym of Cyprinodontoidei.

  • Comments. In the mid-20th century, Cyprinodontoidei was applied as the name of a group containing Adrianichthyidae, Anablepidae, Cyprinodontidae, Goodeidae, and Poeciliidae (Rosen 1964; Greenwood et al. 1966).

  • Time-calibrated molecular phylogenies estimate divergence times for clades in Cyprinodontoidei that are too young for Gondwanan fragmentation to explain the disjunct geographic distribution of Aplocheilidae (Near et al. 2013; Amorim and Costa 2018; Hughes et al. 2018; Ghezelayagh et al. 2022; Piller et al. 2022). Initial phylogenetic analyses of mtDNA and morphological characters (Murphy and Collier 1997; Costa 2004, 2012a, 2012b) resolved Nothobranchiidae and Rivulidae as sister lineages to the exclusion of Aplocheilidae, a relationship consistent with vicariance-driven diversification resulting from Gondwanan fragmentation. However, both the consistent resolution of Aplocheilidae and Nothobranchiidae as sister lineages (e.g., Amorim and Costa 2018; Piller et al. 2022) and relaxed molecular clock age estimates that date the diversification of Cyprinodontoidei to the latest part of the Cretaceous (e.g., Amorim and Costa 2018; Ghezelayagh et al. 2022; Piller et al. 2022) contradict the Gondwanan vicariance scenario.

  • Bayesian relaxed molecular clock analyses of Cyprinodontoidei result in an average posterior crown age estimate of 76.2 million years ago, with the credible interval ranging between 65.0 and 88.6 million years ago (Ghezelayagh et al. 2022).

  • img-z135-7_03.gif

    Blenniiformes P. Bleeker 1859:xxv
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The most inclusive crown clade that contains Lamprologus callipterus Boulenger 1906, Chromis chromis (Linnaeus 1758), Crenimugil crenilabis (Forsskål in Niebuhr 1775), Embiotoca jacksoni Agassiz 1853, Gobiesox maeandricus (C. Girard 1858a), Scartella cristata (Linnaeus 1758), Blennius ocellaris Linnaeus 1758, and Gibbonsia metzi Hubbs 1927, but not Atherina presbyter Cuvier 1829. This is a minimum-crown-clade definition with an external specifier.

  • Etymology. From the ancient Greek βλέννoς (bl̍εno͡Ʊz) used in reference to blennies by ancient Mediterranean authors and also meaning slime or spittle (D. W. Thompson 1947:32–33). The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 956.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, fig. S11). Although Blennius ocellaris is not included in the reference phylogeny, it resolves in a clade with other species of Blenniidae in phylogenetic analyses of Sanger-sequenced mitochondrial and nuclear genes (Almada et al. 2005, fig. 1; Hundt et al. 2014, fig. 2; Hundt and Simons 2018, figs. 4–6; Vecchioni et al. 2019, fig. 1). Phylogenetic relationships of the major lineages of Blenniiformes are presented in Figure 16. Placement of the fossil pan-pomacentrid †Chaychanus in the phylogeny is on the basis of analysis of morphological characters (Cantalice et al. 2022).

  • Phylogenetics. Monophyly of Blenniiformes was supported in early molecular analyses, but with limited taxon sampling (W. L. Smith and Wheeler 2006; Kawahara et al. 2008). The first resolution of blenniiform monophyly with strong support was a phylogenomic analysis of UCE loci (Ghezelayagh et al. 2022). Within Blenniiformes molecular analyses resolve a clade containing Cichlidae, Pholidichthys, and Polycentridae (Betancur-R et al. 2017; Ghezelayagh et al. 2022; Astudillo-Clavijo et al. 2023), with Cichlidae and Pholidichthys consistently resolved as sister lineages (Wainwright et al. 2012; Friedman, Keck, et al. 2013; Near et al. 2013; Collins et al. 2015; Eytan et al. 2015; Betancur-R et al. 2017; Rabosky et al. 2018; Ghezelayagh et al. 2022; Astudillo-Clavijo et al. 2023). Previous morphological studies led to differing conclusions regarding the relationships of Mugilidae within Percomorpha: analysis of branchial musculature supported a hypothesis that Mugilidae and Atheriniformes are sister lineages (Stiassny 1990), but analysis of pelvic girdle morphology suggested a phylogenetic relationship of mullets with “higher” percomorphs (Stiassny 1993). Both of these analyses were conducted in the context of an Acanthopterygii that placed Atheriniformes outside of Percomorpha (Rosen 1973; Rosen and Parenti 1981; Lauder and Liem 1983), so these seemingly inconsistent conclusions from two different anatomical systems seem to be clarified in the context of a phylogeny in which Atheriniformes is nested within Percomorpha (G. D. Johnson and Patterson 1993; Miya et al. 2003). In some molecular phylogenies, Mugilidae is resolved as either the sister lineage of Embiotocidae or Ambassidae (Asiatic glassfishes) (Wainwright et al. 2012; Near et al. 2013; Collins et al. 2015; Betancur-R et al. 2017; Hughes et al. 2018; Ghezelayagh et al. 2022). Congrogadidae is distantly related to Pseudochromidae in molecular phylogenies (Near et al. 2013; Betancur-R et al. 2017; Ghezelayagh et al. 2022), despite being well nested in Pseudochromidae in phylogenetic analyses of egg morphology, osteology, and external morphological characters (Godkin and Winterbottom 1985; Mooi 1990; A. C. Gill 2013).

  • A notable result of molecular phylogenetic analyses of Blenniiformes is the consistent resolution of a clade containing Gramma, Opistognathidae, Gobiesocidae, and Blennioidei, but exclusive of Lipogramma (Wainwright et al. 2012; Near et al. 2013; Collins et al. 2015; Eytan et al. 2015; Betancur-R et al. 2017; Hughes et al. 2018; Ghezelayagh et al. 2022). Grammatidae traditionally included Gramma and Lipogramma (G. D. Johnson 1984; J. S. Nelson 1984:281), and morphology of the adductor mandibulae muscles and a phylogenetic analysis of 38 caudal fin skeleton characters offers evidence for monophyly of Grammatidae (A. C. Gill and Mooi 1993; Thieme et al. 2022); however, Lipogramma and Gramma are not resolved as a monophyletic group in phylogenetic analyses of Sanger-sequenced mtDNA and nuclear genes (Betancur-R et al. 2017). Gobiesocidae and Blennioidei are resolved as sister lineages in many molecular phylogenetic studies (W.-J. Chen et al. 2003; Dettaï and Lecointre 2005; Miya et al. 2005; Mabuchi et al. 2007; Kawahara et al. 2008; Setiamarga et al. 2008; Wainwright et al. 2012; Lin and Hastings 2013; Near et al. 2013; Collins et al. 2015; Eytan et al. 2015; Smith et al. 2016; Betancur-R et al. 2017; Fricke et al. 2017; Hughes et al. 2018; Ghezelayagh et al. 2022), supporting conclusions from morphological analyses of gill arch musculature and skeletal anatomy (Rosen and Patterson 1990; Springer and Johnson 2004; Springer and Orrell 2004).

  • Composition. There are currently 3,814 living species of Blenniiformes (Fricke et al. 2023) classified in Ambassidae, Blennioidei, Cichlidae, Congrogadidae, Embiotocidae, Gobiesocidae, Gramma, Lipogramma, Mugilidae, Opistognathidae, Pholidichthys, Plesiopidae, Polycentridae, Pomacentridae, and Pseudochromidae. Fossil blenniiforms include the pan-pomacentrid †Chaychanus gonzalezorum (Appendix 1; Cantalice et al. 2020). Over the past 10 years, 268 living species of Blenniiformes have been described (Fricke et al. 2023), comprising approximately 7.0% of the living species diversity in the clade.

  • Diagnostic apomorphies. There are no known morphological apomorphies for Blenniiformes.

  • Synonyms. There are no synonyms of Blenniiformes.

  • Comments. Alternative classifications apply the group name Blenniiformes to a less inclusive clade we define as Blennioidei (Lin and Hastings 2013; Betancur-R et al. 2017). The earliest fossil blenniiform is the pan-pomacentrid †Chaychanus gonzalezorum from the Danian (66.0–61.7 Ma) of Mexico (Cantalice et al. 2020). Bayesian relaxed molecular clock analyses of Blenniiformes result in an average posterior crown age estimate of 88.7 million years ago, with the credible interval ranging between 77.2 and 100.5 million years ago (Ghezelayagh et al. 2022).

  • img-z137-5_03.gif

    Blennioidei P. Bleeker 1853a:114
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Lepidonectes corallicola (Kendall and Radcliffe 1912), Dactyloscopus lacteus (Myers and Wade 1946), Blennius ocellaris Linnaeus 1758, and Gibbonsia metzi Hubbs 1927. This is a minimum-crown-clade definition.

  • Etymology. Derived from the ancient Greek βλέννoς (bl̍εno͡Ʊz) used in reference to blennies by ancient Mediterranean authors and also meaning slime or spittle (D. W. Thompson 1947:32–33).

  • Registration number. 957.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S11). Although Blennius ocellaris is not included in the reference phylogeny, it resolves in a clade with other species of Blenniidae in phylogenetic analyses of Sanger-sequenced mitochondrial and nuclear genes (e.g., Vecchioni et al. 2019, fig. 1). See Figure 16 for a phylogeny of the lineages comprising Blennioidei.

  • Phylogenetics. Prephylogenetic hypotheses of the relationships of Blennioidei include many disparately related lineages such as Ammodytidae, Congrogadidae, Notothenioidei, Ophidiiformes, Uranoscopidae, and Zoarcoidei (Regan 1912b; Jordan 1923:228–238; Gosline 1968). The delimitation of Blennioidei presented here was first proposed in studies investigating the systematics of Pholidichthys and Clinidae (Springer and Freihofer 1976; George and Springer 1980) and validated in a review of morphological evidence for blennioid monophyly (Springer 1993). Morphological apomorphies were presented for each of the lineages of Blennioidei, but there is no morphological evidence for the monophyly of Labrisomidae (labrosomid blennies) (Springer 1993). The shape of the cartilage of the third infrapharyngobranchials was presented as a possible synapomorphy for a clade within Blennioidei containing Chaenopsidae (true blennies), Dactyloscopidae (sand stargazers), Labrisomidae, and Clinidae (kelp blennies) (J. T. Williams 1990; Springer 1993).

  • Blennioidei is resolved as monophyletic in morphological (Springer and Orrell 2004) and molecular phylogenetic analyses (Lin and Hastings 2013; Near et al. 2013; Betancur-R et al. 2017; Rabosky et al. 2018; Ghezelayagh et al. 2022). Phylogenetic relationships within Blennioidei resulting from morphological and molecular analyses are congruent. A phylogeny based on dorsal gill arch morphology resolved a paraphyletic Tripterygiidae (triplefin blennies) as successive branching lineages with Lepidoblennius as the sister lineage of all other blennioids and Blenniidae (combtooth blennies) as the sister lineage to a clade containing Clinidae, Chaenopsidae, Dactyloscopidae, and Labrisomidae (Springer and Orrell 2004). Molecular phylogenetic analyses resulted in a very similar phylogeny, except Tripterygiidae is monophyletic and relationships within the clade containing Clinidae, Labrisomidae (sensu stricto), Calliclinus, Chaenopsidae (sensu stricto), and Dactylopteridae are fully resolved (Lin and Hastings 2013; Near et al. 2013; Betancur-R et al. 2017; Rabosky et al. 2018; Ghezelayagh et al. 2022).

  • Molecular phylogenetic analyses of Blennioidei with dense taxon sampling reveal that Chaenopsidae and Labrisomidae are paraphyletic (Lin and Hastings 2013; Rabosky et al. 2018). Stathmonotus, traditionally classified in Chaenopsidae, is phylogenetically nested in Labrisomidae. Neoclinini, containing Neoclinus and Mccoskerichthys and traditionally classified in Chaenopsidae, and Cryptotremini, traditionally classified in Labrisomidae, are resolved as successive branching sister lineages to a clade that contains Labrisomidae (sensu stricto), Chaenopsidae (sensu stricto), and Dactyloscopidae. Calliclinus, traditionally classified in Cryptotremini and Labrisomidae, is the sister lineage of the inclusive clade containing Neoclinini, Cryptotremini, Labrisomidae (sensu stricto), Chaenopsidae (sensu stricto), and Dactyloscopidae (Lin and Hastings 2013; Rabosky et al. 2018). There are available family group-names for Calliclinus, Neoclinini, and Cryptotremini (Van der Laan et al. 2014), but we leave the establishment of taxonomic families for these lineages to future research.

  • Composition. There are currently 948 living species of Blennioidei (Fricke et al. 2023) classified in Blenniidae, Calliclinus, Chaenopsidae, Clinidae, Cryptotremini, Dactyloscopidae, Labrisomidae, Neoclinini, and Tripterygiidae. Over the past 10 years, 24 new living species of Blennioidei have been described, comprising 2.5% of the living species diversity in the clade (Fricke et al. 2023).

  • Diagnostic apomorphies. Morphological apomorphies for Blennioidei include (1) first pharyngobranchial cartilaginous or absent (Springer 1993; Wiley and Johnson 2010), (2) second and fourth pharyngobranchials absent (Springer 1993; Wiley and Johnson 2010), (3) uncinate process or associated interarcual cartilage of first epibranchial absent (Springer 1993; Wiley and Johnson 2010), (4) unique pelvic girdle with bean-shaped pelvis (Springer 1993; Wiley and Johnson 2010), (5) unique, simplified caudal fin (Springer 1993; Wiley and Johnson 2010), (6) neural spines lacking on first vertebrae or several of the anteriormost vertebrae (G. D. Johnson 1993; Wiley and Johnson 2010), (7) first external levator and fourth transversus ventralis absent (Springer and Orrell 2004; Wiley and Johnson 2010), (8) proximal pectoral fin radials longer than wide (Lin and Hastings 2013), (9) unbranched pectoral fin rays (Lin and Hastings 2013), and (10) haemal arch of preural centrum 2 fused with its centrum (Thieme et al. 2022).

  • Synonyms. Blenniicae (Hubbs 1952:51, fig. 1) andBlenniiformes(WileyandJohnson2010:160; J. S. Nelson et al. 2016:346; Betancur-R et al. 2017:26) are ambiguous synonyms of Blennioidei. Blenniiformes (Betancur-R, Broughton, et al. 2013, fig. 8) is a partial synonym of Blennioidei.

  • Comments. The name Blennioidei has long been applied to a group that includes Blenniidae, Chaenopsidae, Clinidae, Dactyloscopidae, Labrisomidae, and Tripterygiidae (Springer and Freihofer 1976; George and Springer 1980; Springer 1993; Hastings and Springer 2009). Since the mid-20th century (Hubbs 1952; Springer and Freihofer 1976; George and Springer 1980), the monophyly of Blennioidei has never been seriously questioned; however, the close relationship between Blennioidei, Opistognathidae, Grammatidae, Embiotocidae, Pomacentridae, and Pseudochromidae is a novel phylogenetic resolution derived from analyses of molecular data (Wainwright et al. 2012; Ghezelayagh et al. 2022). The lineages not currently placed in Linnaean families are listed with generic and tribe names in the classification outlined in Appendix 2 and in the Constituent lineages section below.

  • The earliest fossil Blennioidei is the otolith species †Exallias vectensis from the Ypresian (56.0–48.1 Ma) of France (Nolf 1972; Nolf and Lapierre 1979). The earliest skeletal fossils of Blennioidei are from the Serravallian (13.82–11.63 Ma) of Azerbaijan, Bosnia, Croatia, and Moldova (Anđelković 1989; Bannikov 1998). Bayesian relaxed molecular clock analyses of Blennioidei result in an average posterior crown age estimate of 48.5 million years ago, with the credible interval ranging between 38.6 and 60.2 million years ago (Ghezelayagh et al. 2022).

  • img-z139-2_03.gif

    Carangiformes D. S. Jordan 1923:183
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Centropomus medius Günther 1864b, Polynemus melanochir Valenciennes in Cuvier and Valenciennes (1831), Psettodes erumei (Bloch and Schneider 1801), Pleuronichthys cornutus (Temminck and Schlegel 1846), Xiphias gladius Linnaeus 1758, Caranx melampygus Cuvier in Cuvier and Valenciennes (1833), and Caranx melampygus (Linnaeus 1766). This is a minimum-crown-clade definition.

  • Etymology. From the French carangue, referring to a Caribbean flatfish. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 962.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, figs. S14–S15). Although Caranx hippos is not in the reference phylogeny, it resolves in a monophyletic Carangidae with other species of Caranx in phylogenies inferred from Sanger-sequenced genes and UCE loci (Reed et al. 2002, fig. 3; Damerau et al. 2018, fig. 1; Glass et al. 2023, fig. 2B). Phylogenetic relationships among the major lineages of Carangiformes are presented in Figure 17. The placement of the fossil pan-latid †Eolates in the phylogeny of Carangiformes is on the basis of phylogenetic analysis of morphological characters (Otero 2004).

  • Phylogenetics. The resolution of Carangiformes as a monophyletic group is one of several surprising results in the phylogenetics of Percomorpha to emerge over the past two decades (Miya et al. 2003, 2005; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Musilova et al. 2019; Dornburg and Near 2021; Ghezelayagh et al. 2022). Carangiformes includes biologically and phenotypically disparate lineages, many of which have long evaded confident phylogenetic resolution. For example, Pleuronectoidei (flatfishes) are morphologically among the most atypical of all teleosts and prior to the application of molecular data had not been confidently placed among major lineages of percomorphs (Figure 1; Regan 1913b, 1929; Norman 1934; Chapleau 1993). On the other hand, the billfishes Istiophoridae (marlins) and Xiphias gladius (Swordfish) were classified with tunas in Scombroidei throughout the 20th century on the basis of presumably strong morphological evidence (Regan 1909b; Greenwood et al. 1966; Collette, Potthoff, et al. 1984; G. D. Johnson 1986; J. S. Nelson 2006:430–434), but are unvaryingly resolved within Carangiformes in molecular phylogenetic studies (e.g., Orrell et al. 2006; Little et al. 2010; Hughes et al. 2018; Ghezelayagh et al. 2022).

  • The lineages comprising Carangiformes were never grouped together in classifications based on morphology (Greenwood et al. 1966; Wiley and Johnson 2010); however, monophyly of the group is consistently supported in a wide range of molecular phylogenetic studies that include analyses of whole mtDNA genomes, Sanger-sequenced mtDNA and nuclear genes, and phylogenomic datasets (W.-J. Chen et al. 2003; Miya et al. 2003, 2005; Dettaï and Lecointre 2005, 2008; W. L. Smith and Wheeler 2006; W. L. Smith and Craig 2007; B. Li et al. 2009; C. H. Li et al. 2011; Betancur-R, Broughton, et al. 2013; Campbell, Chen, et al. 2013, 2014; Near et al. 2013; Davis et al. 2016; Harrington et al. 2016; Sanciangco et al. 2016; Smith et al. 2016; Betancur-R et al. 2017;Hughes et al. 2018; Ribeiro, Davis, et al. 2018; Shi et al. 2018; M. G. Girard et al. 2020; Ghezelayagh et al. 2022; M. G. Girard, Davis, Baldwin, et al. 2022; Mu et al. 2022). Phylogenetic analyses inferred from DNA sequences of more than 950 UCE loci and a combined dataset of 201 morphological characters and more than 450 UCE loci result in phylogenies that are strongly congruent and include three major clades within Carangiformes (M. G. Girard et al. 2020; Ghezelayagh et al. 2022): (1) Centropomus (snooks), Latidae (lates perches), Lactarius lactarius (False Trevally), and Sphyraena (barracudas); (2) Polynemidae (threadfins) and Pleuronectoidei (flatfishes); (3) and Carangoidei. The analysis of combined phenotypic and molecular characters results in the identification of morphological apomorphies for Carangiformes and several of the constituent lineages in the clade (M. G. Girard et al. 2020).

  • Composition. There are currently 1,107 living species of Carangiformes (Fricke et al. 2023) that include Lactarius lactarius, Mene maculata, Nematistius pectoralis, Xiphias gladius, and species classified in Centropomus, Leptobrama, Sphyraena, Istiophoridae, Latidae, Polynemidae, Carangoidei, and Pleuronectoidei (M. G. Girard et al. 2020; M. G. Girard, Davis, Tan, et al. 2022). Fossil lineages of Carangiformes include the pan-latid †Eolates gracilis (Sorbini 1970; Otero 2004) and several taxa in Carangoidei and Pleuronectoidei. Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, 36 new living species of Carangiformes have been described (Fricke et al. 2023), comprising 3.3% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Carangiformes include (1) presence of external process on the maxilla (M. G. Girard et al. 2020), (2) accessory gill rakers present on lateral aspect of branchial arches (M. G. Girard et al. 2020), (3) accessory gill rakers present on medial aspect of branchial arches (M. G. Girard et al. 2020), (4) presence of an epibranchial 2 toothplate that is serially associated with the second pharyngobranchial toothplate (M. G. Girard et al. 2020), (5) contact at metapterygoid-hyomandibular border ranging from a single pointed process inserting into evagination to a moderate amount of suturing between elements (M. G. Girard et al. 2020), (6) first hemal spine with a simple configuration, similar to more posterior hemal spines (M. G. Girard et al. 2020), and (7) pored lateral line scales absent from caudal fin (M. G. Girard et al. 2020).

  • Synonyms. Carangaria (Betancur-R et al. 2017:24), Carangimorphariae (Betancur-R, Broughton, et al. 2013, fig. 7; Betancur-R and Ortí 2014, fig. 1), and clade L (W.-J. Chen et al. 2003:279, tbl. 4; Dettaï and Lecointre 2005, fig. 3, tbl. 4, 2008, fig. 5, tbl. 4) are ambiguous synonyms of Carangiformes.

  • Comments. The resolution of the clade Carangiformes is not only one of several unexpected results in the molecular phylogenetics of Percomorpha (Dornburg and Near 2021), but also exemplifies the utility of molecular phylogenies in aiding with the discovery of morphological apomorphies for these newly delimited and inclusive lineages of teleost fishes (M. G. Girard et al. 2020). Over the past 10 years, the names Carangimorphariae (Betancur-R, Broughton, et al. 2013), Carangiformes (Davis et al. 2016, tbl. S3), and Carangaria (Betancur-R et al. 2017) have all been applied to this clade. We follow more recent efforts that use the name Carangiformes (Davis et al. 2016; M. G. Girard et al. 2020; Dornburg and Near 2021; Ghezelayagh et al. 2022).

  • Carangiformes includes a variety of large, usually laterally compressed and generally strong-swimming fishes, many of which exhibit a degree of phenotypic distinctiveness that motivated their classification in monotypic or monogeneric taxonomic families. Time-calibrated phylogenies indicate that Carangiformes originated in the Late Cretaceous and the major lineages diversified throughout the Paleogene (Santini and Carnevale 2015; Harrington et al. 2016; Ribeiro, Davis, et al. 2018; Ghezelayagh et al. 2022). Bayesian relaxed molecular clock analyses of Carangiformes result in an average posterior crown age estimate of 75.7 million years ago, with the credible interval ranging between 66.4 and 86.5 million years ago (Ghezelayagh et al. 2022).

  • img-z141-8_03.gif

    FIGURE 17.

    Phylogenetic relationships of the major living lineages and fossil taxa of Synbranchiformes, Synbranchoidei, Anabantoidei, Carangiformes, Carangoidei, and Pleuronectoidei. Filled circles identify the common ancestor of clades with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z140-1_03.jpg

    Pleuronectoidei P. Bleeker 1849:6

  • Definition. The least inclusive crown clade that contains Psettodes erumei (Bloch and Schneider 1801), Citharus linguatula (Linnaeus 1758), Pleuronichthys cornutus (Temminck and Schlegel 1846), Solea solea (Linnaeus 1758), and Pleuronectes platessa Linnaeus 1758. This is a minimum-crown-clade definition, but the clade is not defined using the PhyloCode.

  • Etymology. From the ancient Greek πλευρόν (pl̍ːαn) meaning flank or side and νήκτos (n̍εkto͡Ʊz) meaning swimming.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of mitochondrial and nuclear genes (Campbell et al. 2019, fig. 1). Although Pleuronectes platessa is not in the reference phylogeny, it resolves with other species of Pleuronectidae in molecular phylogenetic analyses (Kartavtsev et al. 2008, fig. 1; Ji et al. 2016, fig. 1; Vinnikov et al. 2018, fig. 1). Phylogenetic relationships of the major living lineages and fossil taxa of Pleuronectoidei are presented in Figure 17. The placements of fossil taxa in the phylogeny of Pleuronectoidei are on the basis of resolutions suggested in the literature for the pan-pleuronectoids †Amphistium paradoxum (Chanet et al. 2020), †Eobothus minimus (Chanet 1999; Friedman 2008; Campbell et al. 2019), †Heteronectes chaneti (Chanet et al. 2020), the pan-pleuronectid †Oligopleuronectes germanicus (Sakamoto et al. 2004; Harrington et al. 2016), the pan-bothid †Oligobothus pristinus (Baciu and Chanet 2002; Campbell et al. 2019), and the pan-soleid †Eobuglossus eocenicus (Chanet 1994; Campbell et al. 2019).

  • Phylogenetics. Classifications of teleosts from the late 19th through the 20th centuries grouped species of Pleuronectoidei into three separate lineages: Psettodes (spiny turbots), pleuronectoids (flounders), and soleioids (soles), often visualizing hypothesized relationships in prephylogenetic branching diagrams in which flounders and soles were depicted as closely related (Jordan and Evermann 1898:2602–2712; Regan 1910b:490; Norman 1934:43; Hubbs 1945, fig. 1; Amaoka 1969, fig. 131; Hensley and Ahlstrom 1984, fig. 358; Hensley 1997). The flounders and soles each contained two groups based on the orientation of the eyes: whether the eyes are placed on the right or left side of the head (Regan 1910b; Hubbs 1945).

  • A phylogenetic tree of Pleuronectoidei with 10 mapped morphological character state changes depicts Psettodes as the sister lineage of all other pleuronectoids and places the flounders as paraphyletic relative to the soles (Lauder and Liem 1983, fig. 63). The first explicit phylogenetic study of relationships within pleuronectoids was an analysis of 39 morphological characters and supported the monophyly of Pleuronectoidei, placed Psettodes as the sister lineage of all other pleuronectoids, and found that the traditional delimitation of Pleuronectidae (righteye flounders) is not monophyletic, prompting the removal of Samaridae (crested flounders), Poecilopsettidae (bigeye flounders), Rhombosoleidae (oblique flounders), and Paralichthodes algoensis (Peppered Flounder) (Chapleau 1993). Subsequent morphological phylogenetic analyses were aimed at resolving the relationships within Pleuronectidae (Cooper and Chapleau 1998a), the placement of Paralichthodes within Pleuronectoidei (Cooper and Chapleau 1998b), assessing the monophyly and relationships within Citharidae (largescale flounders) (Hoshino 2001), the relationships within Scophthalmidae (turbots) (Chanet 2003), and the phylogenetic resolution of pleuronectoid fossil lineages (Chanet 1999; Baciu and Chanet 2002; Friedman 2008).

  • The morphological dataset of Chapleau (1993) was modified by the rescoring of the recessus orbitalis, a fluid-filled sac behind the eyeballs that is used to elevate the eyes above the surface of the head (Bürgin 1989; Campbell et al. 2020), from present to absent in Psettodes; the addition of fossil taxa †Heteronectes and †Amphistium to the data matrix; and the inclusion of the carangiform lineages Lates and Caranx as outgroup taxa (Chanet et al. 2020). Analysis of 39 morphological characters resulted in 12 most parsimonious trees, among which six resolved Psettodes as the sister lineage of all other pleuronectoids and the other six resolved Pleuronectoidei as paraphyletic with Psettodes as the sister lineage of a clade containing Caranx and all other pleuronectoids (Chanet et al. 2020). In contrast to earlier studies that place †Heteronectes and †Amphistium as pan-pleuronectoids (Friedman 2008, 2012), phylogenetic analysis of the modified Chapleau (1993) morphological dataset resolves †Heteronectes as the sister lineage of a clade containing †Amphistium and all other pleuronectoids to the exclusion of Psettodes (Chanet et al. 2020).

  • The uncertainty regarding the monophyly of Pleuronectoidei in phylogenetic analyses of morphological characters is reflected in molecular studies. Most molecular phylogenetic studies resolve Pleuronectoidei as paraphyletic with Psettodes resolving as the sister lineage of any number of other lineages in Carangiformes, almost always with low node support and separated from other pleuronectoids by only one or a small number of nodes in the phylogeny (Dettaï and Lecointre 2005; W. L. Smith and Wheeler 2006; B. Li et al. 2009; C. H. Li et al. 2011; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Campbell, Chen, et al. 2013; Near et al. 2013; Shi et al. 2018; Lü et al. 2021). However, analysis of concatenated Sanger-sequenced mtDNA and nuclear genes and phylogenomic analyses of ultraconserved element loci resolves Psettodes as the sister lineage of all other pleuronectoids (Harrington et al. 2016; Sanciangco et al. 2016; Ghezelayagh et al. 2022), with substantial support in genome-wide concordance analysis (Harrington et al. 2016).

  • Other molecular studies had limited or no outgroup taxon sampling to test the monophyly of Pleuronectoidei (Berendzen and Dimmick 2002; Azevedo et al. 2008; Campbell et al. 2019; Atta et al. 2022), or obtained weak support for flatfish monophyly after ad hoc manipulation of DNA sequences through RY and AGY coding (Betancur-R, Li, et al. 2013). It is reasonable to consider that both random and systematic error through incomplete lineage sorting and unequal nucleotide base frequencies likely contribute to the monophyly of Pleuronectoidei as a challenging phylogenetic problem (Betancur-R, Li, et al. 2013; Betancur-R and Ortí 2014; Harrington et al. 2016). However, it is also important to consider the perspective of advocates of pleuronectoid paraphyly, who point out that higher support values in molecular phylogenies are not indicative of phylogenetic signal, the monophyly of Pleuronectoidei is not adequately demonstrated through phylogenetic analysis of morphological characters, and confidence in phylogenetic conclusions is weakened by selective reporting of results that match a researcher's expectations (Campbell et al. 2014).

  • Despite the uncertainty of pleuronectoid monophyly in molecular phylogenies, analyses with comprehensive taxon sampling resolve five major lineages of Pleuronectoidei: (1) Psettodes; (2) Citharidae; (3) Scophthalmidae, Pleuronectidae, Paralichthyidae (sand flounders), Cyclopsettidae (sand whiffs), and Bothidae (lefteye flounders) (Harrington et al. 2016; Byrne et al. 2018; Campbell et al. 2019; Ghezelayagh et al. 2022); (4) Achiridae (American soles), Paralichthodes, Oncopterus darwini (Remo Flounder), Rhombosoleidae, and Achiropsettidae (southern flounders) (Campbell et al. 2019); and (5) Samaridae, Poecilopsettidae, Soleidae (soles), and Cynoglossidae (tongue-fishes) (Betancur-R, Li, et al. 2013; Harrington et al. 2016; Byrne et al. 2018; Campbell et al. 2019; Ghezelayagh et al. 2022). Phylogenetic analysis of molecular data demonstrated the paraphyly of Paralichthyidae and Rhombosoleidae, prompting the description of the taxonomic families Cyclopsettidae and Oncopteridae (Campbell et al. 2019).

  • Composition. There are currently 818 living species of Pleuronectoidei (Munroe 2015; Fricke et al. 2023) that include Paralichthodes algoensis, Oncopterus darwini, and species classified in Achiridae, Achiropsettidae, Bothidae, Citharidae, Cyclopsettidae, Cynoglossidae, Paralichthyidae, Pleuronectidae, Poecilopsettidae, Psettodes, Rhombosoleidae, Samaridae, Scophthalmidae, and Soleidae. Fossil lineages of Pleuronectoidei include the pan-pleuronectoids †Amphistium paradoxum, †Eobothus minimus, †Heteronectes chaneti (Chanet 1999; Friedman 2008, 2012), the pan-pleuronectid †Oligopleuronectes germanicus (Sakamoto et al. 2004), the pan-bothid †Oligobothus pristinus (Baciu and Chanet 2002), and the pan-soleid †Eobuglossus eocenicus (Chanet 1994). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, there have been 26 new living species of Pleuronectoidei described (Fricke et al. 2023), comprising approximately 3.2% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Pleuronectoidei include (1) ontogeny characterized by migration of one eye across the dorsal midline (Chapleau 1993; Friedman 2008; Wiley and Johnson 2010; Chanet et al. 2020; M. G. Girard et al. 2020), (2) dorsal fin anteriorly placed, partially overlapping neurocranium (Chapleau 1993; Chanet 1995, 1997, 1999; Wiley and Johnson 2010; Chanet et al. 2020), (3) pseudomesial bar present (Harrington et al. 2016; M. G. Girard et al. 2020), (4) dorsalmost element of postcleithrum not expanded posteriorly through the margin (M. G. Girard et al. 2020), and (5) asymmetric pigmentation between eyed and blind sides (Harrington et al. 2016; M. G. Girard et al. 2020).

  • Synonyms. Heterostomata (Cope 1871a:458; Gill 1893:137; Jordan and Evermann 1898:2602; Regan 1910b:491), Zeorhombiformes (Goodrich 1909:465–474), Pleuronectiformes (Berg 1940:492–493; Greenwood et al. 1966:402; McAllister 1968:131–133; Gosline 1971:165–167; Wiley and Johnson 2010:167; J. S. Nelson et al. 2016:395–405; Betancur-R et al. 2017:25), and Pleuronectoideo (M. G. Girard et al. 2020:275) are ambiguous synonyms of Pleuronectoidei.

  • Comments. The phylogenetic relationships of Pleuronectoidei, particularly an inference of monophyly for the lineage, remains one of the most challenging problems in the phylogenetics of ray-finned fishes (Betancur-R, Li, et al. 2013; Betancur-R and Ortí 2014; Campbell et al. 2014; Harrington et al. 2016; Chanet et al. 2020). Molecular data applied to assessing pleuronectoid monophyly will likely continue to carry the burdens of random and systematic error, but recent efforts demonstrate the potential for additional discovery of morphological characters to aid in the phylogeny of Pleuronectoidei (Harrington et al. 2016; Chanet et al. 2020; M. G. Girard et al. 2020).

  • The earliest skeletal fossils of Pleuronectoidei are from the Ypresian (56.0–48.1 Ma) of Italy and include †Amphistium, †Eobothus, and †Heteronectes (Bannikov 2014b; Carnevale et al. 2014). Bayesian relaxed molecular clock analyses of Pleuronectoidei result in an average posterior crown age estimate of 67.6 million years ago, with the credible interval ranging between 59.0 and 77.5 million years ago (Ghezelayagh et al. 2022).

  • img-z144-6_03.gif

    Carangoidei P. Bleeker 1859:xxiii
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Leptobrama muelleri Steindachner 1878, Toxotes jaculatrix (Pallas 1769), Xiphias gladius Linnaeus 1758, Echeneis naucrates Linnaeus 1758, Caranx hippos (Linnaeus 1766), and Caranx melampygus Cuvier in Cuvier and Valenciennes (1833). This is a minimum-crown-clade definition.

  • Etymology. From the French carangue, referring to a Caribbean flatfish.

  • Registration number. 964.

  • Reference phylogeny. A phylogeny inferred from sequences of 1,314 ultraconserved element loci (Glass et al. 2023, fig. 2). Phylogenetic relationships of the major living lineages and fossil taxa of Carangoidei are presented in Figure 17. The placements of fossil taxa in the phylogeny of Carangoidei are on the basis of resolutions suggested in the literature for the pan-menid †Mene purdyi (Friedman and Johnson 2005), the pan-coryphaenoid †Ductor (Friedman, Johanson, et al. 2013), the pan-echeneid †Opisthomyzon (Friedman, Johanson, et al. 2013), the pan-carangid †Archaeus (Santini and Carnevale 2015), the pan-xiphioid †Palaeorhynchus (Sytchevskaya and Prokofiev 2002; Monsch and Bannikov 2011), the pan-istiophorid †Hemingwaya (Monsch and Bannikov 2011), and the pan-xiphiids †Blochius and †Xiphiorhynchus (Monsch and Bannikov 2011).

  • Phylogenetics. Carangoidei was initially delimited to include Carangidae (jacks and pompanos), Coryphaena (dolphinfishes), Echeneidae (remoras), Rachycentron canadum (Cobia), and Nematistius pectoralis (Roosterfish) on the basis of two compelling morphological apomorphies (G. D. Johnson 1984; Smith-Vaniz 1984; G. D. Johnson 1993). Within Carangoidei, morphological and molecular phylogenetic analyses support the monophyly of the echeneoids including Coryphaena, Echeneidae, and Rachycentron and resolve a clade containing Carangidae and the echeneoids (G. D. Johnson 1984; Smith-Vaniz 1984; O'Toole 2002; Reed et al. 2002; Friedman, Johanson, et al. 2013; Near et al. 2013; Santini and Carnevale 2015; Harrington et al. 2016; Ribeiro, Davis, et al. 2018; M. G. Girard et al. 2020; Ghezelayagh et al. 2022; Glass et al. 2023).

  • Where morphological and molecular phylogenies of Carangoidei differ is in the relationships of Nematistius and the monophyly of Carangidae. Molecular phylogenies suggest a more inclusive Carangoidei with Nematistius nested in a clade containing Leptobrama (beachsalmons), Toxotidae (archerfishes), Mene maculata (Moonfish), Istiophoridae (marlins), and Xiphias gladius (Swordfish) (M. G. Girard et al. 2020; Ghezelayagh et al. 2022; Glass et al. 2023). The monophyly of Carangidae is supported by the morphology of the anal fin pterygiophores and the presence of a prominent gap between the second and third anal fin spines (G. D. Johnson 1984; Smith-Vaniz 1984; Gushiken 1988); however, phylogenetic analyses of molecular data and combined molecular and morphological datasets resolve the traditionally delimited Carangidae as paraphyletic with the lineages Trachinotinae and Scomberoidinae as a monophyletic group that is the sister lineage of the echeneoids (W. L. Smith and Wheeler 2006; Near, Eytan, et al. 2012; Santini and Carnevale 2015; Harrington et al. 2016; Mirande 2017; Rabosky et al. 2018; Ribeiro, Davis, et al. 2018; M. G. Girard et al. 2020; Ghezelayagh et al. 2022; Glass et al. 2023). Consistent with a morphological phylogeny (Prokofiev 2002a), molecular phylogenies resolve Trachinotinae and Scomberoidinae as a monophyletic group, but both lineages are paraphyletic because Lichia amia (Leerfish, Trachinotinae) and Parona signata (Leatherjacket, Scomberoidinae) form a clade that is the sister lineage to a monophyletic group containing Trachinotus, Oligoplites, and Scomberoides (Rabosky et al. 2018; Glass et al. 2023). We classify species of Lichia, Parona, Trachinotus, Oligoplites, and Scomberoides in the clade Trachinotidae, which is a valid family-group name under the International Code of Zoological Nomenclature (Van der Laan et al. 2014:98).

  • Composition. There are currently 193 living species (Fricke et al. 2023) of Carangoidei that include Mene maculata, Nematistius pectoralis, Rachycentron canadum, Xiphias gladius, and species classified in Leptobrama, Toxotidae, Carangidae, Coryphaena, Echeneidae, Istiophoridae, and Trachinotidae. Fossil lineages of Carangoidei include the pan-menid †Mene purdyi (Friedman and Johnson 2005), the pan-coryphaenoid †Ductor vestenae (Friedman, Johanson, et al. 2013), the pan-istiophorid †Hemingwaya sarissa (Sytchevskaya and Prokofiev 2002; Monsch and Bannikov 2011), the pan-carangid †Archaeus oblongus (Danil'chenko 1968; Sytchevskaya and Prokofiev 2002), the pan-xiphiids †Blochius longirostris and †Xiphiorhynchus parvus (Volta 1796; Casier 1966:314–315; Bannikov 2014b; Carnevale et al. 2014), the pan-xiphioid †Palaeorhynchus senectus (Danilit'chenko 1962), and the pan-echeneid †Opisthomyzon glaronensis (Wettstein 1886; Friedman, Johanson, et al. 2013). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, there have been 10 new living species of Carangoidei described (Fricke et al. 2023), comprising 5.2% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies of Carangoidei include (1) dentition present on basihyal (M. G. Girard et al. 2020), (2) supracleithrum short (M. G. Girard et al. 2020), (3) neural spine of second preural centrum reduced, not extending posteriorly to bend in ural centrum (M. G. Girard et al. 2020), and (4) cycloid scales (M. G. Girard et al. 2020).

  • Synonyms. Carangiformes (Wiley and Johnson 2010:160; Betancur-R, Broughton, et al. 2013, fig. 7; J. S. Nelson et al. 2016:383; Betancur-R et al. 2017:24-25) is a partial synonym of Carangoidei.

  • Comments. Carangoidei was the name applied to a clade consisting of Carangidae, Coryphaena, Echeneidae, Rachycentron canadum, and Nematistius pectoralis (G. D. Johnson 1993). Molecular phylogenies result in a more inclusive Carangoidei because Nematistius is nested in a clade containing Leptobrama, Toxotidae, Mene, Istiophoridae, and Xiphias (M. G. Girard et al. 2020; Ghezelayagh et al. 2022; Glass et al. 2023).

  • The earliest fossil Carangoidei is the pan-menid †Mene purdyi from the Thanetian and Ypresian (59.2–56.0, 56.0–47.1 Ma) of Peru (Friedman and Johnson 2005). Bayesian relaxed molecular clock analyses of Carangoidei result in an average posterior crown age estimate of 69.0 million years ago, with the credible interval ranging between 61.6 and 77.9 million years ago (Ghezelayagh et al. 2022).

  • img-z146-6_03.gif

    Synbranchiformes P. H. Greenwood, D. E. Rosen, S. H. Weitzman, and G. S. Myers 1966:398
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Indostomus paradoxus Prashad and Mukerji 1929, Synbranchus marmoratus Bloch 1795, Mastacembelus mastacembelus (Banks and Solander in A. Russell 1794), Channa argus (Cantor 1842), Badis badis (Hamilton 1822), and Anabas testudineus (Bloch 1792b). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek σύν (s̍In) meaning together or with and βραγχίoν (bρ̍æki͡әn) which is Latinized as branchium meaning a fish gill. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 965.

  • Reference phylogeny. A phylogeny inferred from combined DNA sequence dataset consisting of 998 ultraconserved element loci and Sanger-sequenced mitochondrial and nuclear genes (Harrington et al. 2023, figs. 2, 3). Phylogenetic relationships among the major lineages of Synbranchiformes are presented in Figure 17.

  • Phylogenetics. The lineages that comprise Synbranchiformes were traditionally classified in groups delimited here as Synbranchoidei (sans Indostomus) and Anabantoidei (Greenwood et al. 1966; Wiley and Johnson 2010; J. S. Nelson et al. 2016:380–383, 390–395). There were suggestions based on morphology that synbranchoids and anabantoids shared common ancestry (Gosline 1971:161; Lauder and Liem 1983; Rosen and Patterson 1990; Roe 1991), but this hypothesis was dismissed in a phylogenetic study utilizing morphological characters (G. D. Johnson and Patterson 1993). Molecular phylogenetic analyses of Percomorpha consistently resolve Synbranchiformes as monophyletic (W.J. Chen et al. 2003; B. Li et al. 2009; Near, Eytan, et al. 2012; Wainwright et al. 2012; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Davis et al. 2016; Smith et al. 2016; Betancur-R et al. 2017; Hughes et al. 2018; Ghezelayagh et al. 2022; Harrington et al. 2023).

  • Composition. There are currently 414 living species (Fricke et al. 2023) of Synbranchiformes classified in Anabantoidei and Synbranchoidei. Over the past 10 years, there have been 63 new living species of Synbranchiformes described (Fricke et al. 2023), comprising 15.2% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies have not been identified for Synbranchiformes; however, most species have an accessory respiratory organ (suprabranchial organ or suprabranchial pouches), which are highly vascularized chambers located above the gill chamber that allow the fishes to breathe air (Johansen 1966; Rosen and Greenwood 1976; Lauder and Liem 1983; Tate et al. 2017).

  • Synonyms. Labyrinthici (Rosen and Patterson 1990:3), Anabantiformes (B. Li et al. 2009, tbl. 4; Near et al. 2013, fig. S1), Anabantomorphariae (Betancur-R, Broughton, et al. 2013:13), and Anabantaria (Betancur-R et al. 2017:24) are ambiguous synonyms of Synbranchiformes.

  • Comments. The name Synbranchiformes is applied to the clade containing Anabantoidei and Synbranchoidei in several recent classifications of percomorphs (Davis et al. 2016; Dornburg and Near 2021; Ghezelayagh et al. 2022).

  • Several lineages of Synbranchiformes have distributions that are disjunct between Africa and South Asia, with phylogenetic patterns consistent with vicariance due to the breakup of Gondwana and rafting of African species to Asia via the Indian subcontinent (F. Wu et al. 2019; Britz et al. 2020). However, age estimates from relaxed molecular clock analyses for Channidae and Mastacembelidae are too young to be the result of vicariance due to Gondwanan breakup and infer an Asian origin for both lineages (X. Li et al. 2006; Adamson et al. 2010; Day et al. 2017; Rüber et al. 2020; Harrington et al. 2023). Paleontological data similarly do not support the African origin or Gondwanan vicariance hypothesis for Channidae (Capobianco and Friedman 2019). Bayesian relaxed molecular clock analyses of Synbranchiformes result in an average posterior crown age estimate of 79.2 million years ago, with the credible interval ranging between 70.8 and 88.5 million years ago (Harrington et al. 2023).

  • img-z147-8_03.gif

    Synbranchoidei P. Bleeker 1859:xxxii
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Indostomus paradoxus Prashad and Mukerji 1929, Synbranchus marmoratus Bloch 1795, and Mastacembelus mastacembelus (Banks and Solander in A. Russell 1794). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek σύν (s̍In) meaning together or with and βραγχίoν (bρ̍æki͡әn) which is Latinized as branchium meaning a fish gill.

  • Registration number. 966.

  • Reference phylogeny. A phylogeny inferred from combined DNA sequence dataset consisting of 998 ultraconserved element loci and Sanger-sequenced mitochondrial and nuclear genes (Harrington et al. 2023, fig. 2). The phylogenetic relationships among the major lineages of Synbranchoidei are presented in Figure 17.

  • Phylogenetics. Several morphological studies suggest that Mastacembelidae, Chaudhuriidae, and Synbranchidae share common ancestry (McAllister 1968:156–159; Gosline 1983; Travers 1984; G. D. Johnson and Patterson 1993; Britz 1996; Britz and Kottelat 2003). Early molecular phylogenetic studies did not sample Chaudhuriidae but confirmed the monophyly of a lineage containing Mastacembelidae and Synbranchidae (e.g., W.-J. Chen et al. 2003; Dettaï and Lecointre 2005, 2008). Species of Indostomus (armored sticklebacks) were traditionally classified with seahorses and sticklebacks in the polyphyletic Gasterosteiformes or Gasterosteoidei on the basis of the presence of dermal plates along the side of the body, a reduced cranial skeleton, and small mouth size (Greenwood et al. 1966; Britz and Johnson 2002; J. S. Nelson et al. 2016). Molecular phylogenetic analyses of Percomorpha consistently resolve Indostomus in Synbranchoidei (Miya et al. 2003, 2005; Kawahara et al. 2008; B. Li et al. 2009; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Pérez-Rodríguez et al. 2016; Smith et al. 2016; Betancur-R et al. 2017; Ghezelayagh et al. 2022; Harrington et al. 2023). Morphological and molecular datasets are congruent in resolving Chaudhuriidae and Mastacembelidae as sister lineages (Travers 1984; Ghezelayagh et al. 2022).

  • Composition. There are currently 136 living species of Synbranchoidei (Britz and Kottelat 1999; Sabaj et al. 2022; Fricke et al. 2023) classified in Chaudhuriidae, Indostomus, Mastacembelidae, and Synbranchidae. Over the past 10 years, there have been 15 new living species of Synbranchoidei described, comprising 11% of the living species diversity in the clade.

  • Diagnostic apomorphies. Synbranchoidei is diagnosed by an arrangement of the upper jaw relative to the suspensorium that features a disconnect between the upper jaw elements (maxilla and premaxilla) and the palatine-ectopterygoid. The palatine is reduced or absent, the ectopterygoid is greatly enlarged, and neither element articulates with the premaxilla. The premaxilla lacks an ascending process and there is often no rostral cartilage. Instead, the premaxilla and maxilla are displaced anteriad, and the premaxilla articulates directly with the lower surface of the neurocranium (Gosline 1983; Britz and Johnson 2002; Britz and Kottelat 2003). Morphological apomorphies for Synbranchoidei that are not confirmed in Indostomus include (1) extension of dentary posteroventrally along ventral edge of anguloarticular (Travers 1984; Wiley and Johnson 2010), (2) palatine sutured along posterolateral face of vomerine shaft (Travers 1984; Wiley and Johnson 2010), (3) levator operculi inserting on dorsolateral face of opercle (Travers 1984; Wiley and Johnson 2010), (4) hyohyoidei adductores dorsolaterally expanded, sealing operculum to body wall and causing restricted opercular opening (Travers 1984; Wiley and Johnson 2010), (5) anterior surface of occipital joint of first vertebra convex, forming “pluglike” in Synbranchidae (Rosen and Greenwood 1976; Wiley and Johnson 52010) or “ball and socket” joint in Mastacembelidae and Chaudhuriidae (Travers 1984; G. D. Johnson and Patterson 1993; Wiley and Johnson 2010), and (6) anterior vertebrae with distinctive configuration (G. D. Johnson and Patterson 1993; Wiley and Johnson 2010).

  • Synonyms. Synbranchiformes is an ambiguous synonym (Gosline 1983:327; Travers 1984:141; Wiley and Johnson 2010:153; J. S. Nelson et al. 2016:380; Betancur-R et al. 2017:24) and partial synonym (Berg 1940:472) of Synbranchoidei.

  • Comments. The name Synbranchoidei is applied to this clade in several recent studies (Dornburg and Near 2021; Ghezelayagh et al. 2022; Harrington et al. 2023).

  • Bayesian relaxed molecular clock analyses of Synbranchoidei result in an average posterior crown age estimate of 69.7 million years ago, with the credible interval ranging between 58.1 and 79.7 million years ago (Harrington et al. 2023).

  • img-z148-8_03.gif

    Anabantoidei L. S. Berg 1940:485
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive clade that contains Channa argus (Cantor 1842), Badis badis (Hamilton 1822), and Anabas testudineus (Bloch 1792b). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek αναβαίνει (̍ænɐbˌe͡Iniː) meaning goes up among.

  • Registration number. 967.

  • Reference phylogeny. A phylogeny inferred from combined DNA sequence dataset consisting of 998 ultraconserved element (UCE) loci and Sanger-sequenced mitochondrial and nuclear genes (Harrington et al. 2023, figs. 2, 3). Phylogenetic relationships of the major living and fossil lineages of Anabantoidei are presented in Figure 17. The placements of the fossil taxa †Anchichanna and †Eoanabas in the phylogeny are on the basis of inferences from morphology (Murray and Thewissen 2008; F. Wu et al. 2017, 2019).

  • Phylogenetics. Anabantoidei is consistently resolved as monophyletic in molecular phylogenetic analyses and consists of five major clades (Betancur-R, Broughton, et al. 2013; Near et al. 2013; Collins et al. 2015; Betancur-R et al. 2017; F. Wu et al. 2019; Ghezelayagh et al. 2022; Harrington et al. 2023): (1) Nandidae (leaffishes and chameleonfishes), (2) Channidae (snakeheads), (3) Helostoma temminckii (Kissing Gourami), (4) Anabantidae (climbing gouramis), and (5) Osphronemidae (gouramies and fighting fishes).

  • We delimit Nandidae as containing species of Nandus, Badis, Dario, and Pristolepis. Alternatively, these four lineages are classified into three Linnaean-ranked taxonomic families, two of which contain a single genus (Rosen and Patterson 1990; Kullander and Britz 2002; Rüber et al. 2004; Britz et al. 2012; Collins et al. 2015; J. S. Nelson et al. 2016:394–395; Betancur-R et al. 2017; Kullander et al. 2019). Our delimitation of Nandidae is reflected in previous classifications (Jordan 1923:202; J. S. Nelson 2006:381–383), is consistently resolved in molecular phylogenetic analyses of Sanger-sequenced protein-coding loci and phylogenomic analysis of UCE loci (Near et al. 2013; Collins et al. 2015; Ghezelayagh et al. 2022; Harrington et al. 2023), and is supported with two distinct morphological apomorphies (Collins et al. 2015). Previous classifications involving anabantoids were complicated by the assumption that the South American and African Polycentridae, which includes Monocirrhus polyacanthus, Polycentrus, Afronandus sheljuzhkoi, and Polycentropsis abbreviata were related to lineages classified here as Nandidae (Regan 1913b; Greenwood et al. 1966:202; Liem 1970; Nelson 1994:371–373). However, molecular phylogenies resolve Polycentridae in Blenniiformes (Wainwright et al. 2012; Near et al. 2013; Collins et al. 2015; Ghezelayagh et al. 2022).

  • Morphological and molecular phylogenetic analyses are congruent in the resolution of a clade that contains lineages with a suprabranchial labyrinth organ: Helostoma, Anabantidae, and Osphronemidae (e.g., Lauder and Liem 1983; Britz 1994, 2001; Near et al. 2013; Ghezelayagh et al. 2022; Harrington et al. 2023). Most incongruence among previous phylogenetic hypotheses is because of the variable resolution of Channidae as either the sister group of the labyrinth organ clade (Springer and Johnson 2004; Near et al. 2013; Sanciangco et al. 2016; Hughes et al. 2018; F. Wu et al. 2019; Britz et al. 2020; Ghezelayagh et al. 2022; Harrington et al. 2023) or Nandidae (Betancur-R, Broughton, et al. 2013), and Helostoma as either the sister taxon of Anabantidae (Collins et al. 2015; Hughes et al. 2018; Britz et al. 2020), Osphronemidae (Rüber et al. 2006; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Sanciangco et al. 2016), or a clade containing Anabantidae and Osphronemidae (Collins et al. 2015; Ghezelayagh et al. 2022; Harrington et al. 2023).

  • Within Channidae, a phylogeny inferred from Sanger-sequenced mitochondrial and nuclear genes and a dataset of Sanger-sequenced genes and UCE loci resolves Parachanna as the sister taxon of a clade comprising Aenigmachanna and Channa (Britz et al. 2020, fig. S3; Harrington et al. 2023). In contrast, a phylogeny inferred from morphological characters places Aenigmachanna as sister lineage of a clade containing Parachanna and Channa (Britz et al. 2020, fig. 5). Several molecular analyses presented in Britz et al. (2020) were conducted using a topological constraint to reflect the results of the morphological phylogeny that prompted the description of the Linnaean-ranked family Aenigmachannidae (Britz et al. 2020). Even if Aenigmachanna was the sister taxon of a clade containing Channa and Parachanna, it would still be most effectively classified in Channidae. The description of a monogeneric Aenigmachannidae provides no information on phylogeny and only accomplishes the creation of a group name that is redundant with Aenigmachanna. We delimit Channidae as the species classified in Aenigmachanna, Channa, and Parachanna, which is supported with 10 morphological apomorphies (Britz et al. 2020, figs. S1, S2).

  • Composition. There are currently 278 living species of Anabantoidei (Fricke et al. 2023) that include Helostoma temminckii and species classified in Anabantidae, Channidae, Nandidae, and Osphronemidae. Fossil lineages of Anabantoidei include pan-anabantid †Eoanabas thibetana (F. Wu et al. 2017) and the pan-channid †Anchichanna kuldanensis (Murray and Thewissen 2008). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, 48 new living species of Anabantoidei have been described (Fricke et al. 2023), comprising 17.3% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies of Anabantoidei include (1) teeth on parasphenoid (Gosline 1968; Springer and Orrell 2004; Collins et al. 2015), and (2) presence of a cartilage-tipped uncinate process on epibranchial 1 (Springer and Orrell 2004; Collins et al. 2015).

  • Synonyms. Anabantomorpha (Springer and Johnson 2004) and Anabantiformes (Wiley and Johnson 2010:158; Betancur-R, Broughton, et al. 2013:18, fig. 6; J. S. Nelson et al. 2016:390–395; Betancur-R et al. 2017:24) are ambiguous synonyms of Anabantoidei. Labyrinthici (Müller 1845b:102, 130; Günther 1861:373–389), Labyrinthiformes (Müller 1845b:135), Ophicephaliformes (Berg 1940:470–471), Luciocephaloidei (Berg 1940:486), and Anabantiformes (Wiley and Johnson 2010:158–159) are partial synonyms of Anabantoidei.

  • Comments. The name Anabantoidei is applied to this clade in several recent studies (Dornburg and Near 2021; Ghezelayagh et al. 2022; Harrington et al. 2023).

  • The earliest fossil Anabantoidei are †Eochanna chorlakkiensis and †Anchichanna kuldanensis from the Lutetian (47.8–41.2 Ma) of Pakistan that may be conspecific (Roe 1991; Murray and Thewissen 2008). Bayesian relaxed molecular clock analyses of Anabantoidei result in an average posterior crown age estimate of 72.1 million years ago, with the credible interval ranging between 63.1 and 80.4 million years ago (Harrington et al. 2023).

  • img-z150-6_03.gif

    Eupercaria R. Betancur-R, E. O. Wiley, N. Bailly, M. Miya, G. Lecointre, and G. Ortí 2014:website
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown lineage that contains Paracanthus hepatus (Linnaeus 1766), Gerres cinereus (Walbaum 1792), Acropoma japonicum Günther 1859, Labrus bergylta Ascanius 1767, Micropterus salmoides (Lacépède 1802), and Perca fluviatilis (Linnaeus 1758). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek words εὖ (̍iːuː) meaning well or good and πέρκη (p̍ːke͡I), a name applied to many species of fishes by ancient authors (D. W. Thompson 1947:195–197).

  • Registration number. 984.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, figs. S16–S25). The placement of the fossil pan-perciform †Paleoserranus lakamhae differs from a proposed resolution as the sister lineage of Serranidae sensu lato (Cantalice et al. 2022) and is motivated by the persistent paraphyly of Serranidae (e.g., W. L. Smith and Craig 2007; Lautredou et al. 2013; Ghezelayagh et al. 2022) and the resolution of Acanthistius, Anthiadidae, and Epinephelidae as early diverging lineages within Perciformes (Ghezelayagh et al. 2022). See Figures 2 and 14 for a phylogeny of the lineages comprising Eupercaria.

  • Phylogenetics. Eupercaria is a lineage resolved entirely as a result of molecular phylogenetic analyses (W.-J. Chen et al. 2003; Miya et al. 2003, 2005; Dettaï and Lecointre 2005; B. Li et al. 2009; Near, Eytan, et al. 2012; Near, Sandel, et al. 2012; Betancur-R, Broughton, et al. 2013; Near et al. 2013, 2015; Thacker et al. 2015; Davis et al. 2016; Sanciangco et al. 2016; Smith et al. 2016; Betancur-R et al. 2017; Alfaro et al. 2018; Ghedotti et al. 2018; Hughes et al. 2018; Rabosky et al. 2018; Ghezelayagh et al. 2022; W. L. Smith et al. 2022). Despite consistent resolution as a monophyletic group, relationships among the major lineages of Eupercaria initially proved difficult to resolve (e.g., Betancur-R et al. 2017; Alfaro et al. 2018; Hughes et al. 2018; Rabosky et al. 2018). Phylogenomic analysis of UCE loci resolves Perciformes, Centrarchiformes, and Labriformes as successively branching lineages to a clade containing Acropomatiformes and Acanthuriformes (Ghezelayagh et al. 2022).

  • Composition. There are more than 7,070 living species of Eupercaria (Fricke et al. 2023) classified in Acanthuriformes, Acropomatiformes, Centrarchiformes, Labriformes, and Perciformes (Dornburg and Near 2021; Ghezelayagh et al. 2022). Over the past 10 years, 522 new living species of Eupercaria have been described (Fricke et al. 2023), comprising 7.4% of the living species in the clade.

  • Diagnostic apomorphies. There are no known morphological apomorphies for Eupercaria.

  • Synonyms. Percomorpharia is an ambiguous synonym of Eupercaria (Betancur-R, Broughton, et al. 2013, fig. 9).

  • Comments. One of the triumphs of molecular phylogenetics is the resolution of relationships among the myriad lineages of Percomorpha (Dornburg and Near 2021). Toward the end of the 20th century, percomorphs were labeled as the “bush at the top” of the teleost phylogeny (G. J. Nelson 1989:328). After the first wave of molecular phylogenetic studies, the consistent resolution of Eupercaria as a monophyletic group with a limited support for relationships within the clade led to its appropriate identification as “the new bush at the top” (Betancur-R, Broughton, et al. 2013:22). However, phylogenomic studies provide support for the monophyly of Eupercaria and elimination of the last of the percomorph “bushes” through the resolution of relationships among the Perciformes, Centrarchiformes, Labriformes, Acropomatiformes, and Acanthuriformes (Ghezelayagh et al. 2022). The name Eupercaria was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade.

  • The earliest fossil Eupercaria is the pan-perciform †Paleoserranus lakamhae dated to the Danian (66.0–61.7 Ma) in Mexico (Cantalice et al. 2022). Bayesian relaxed molecular clock analyses of Eupercaria result in an average posterior crown age estimate of 93.7 million years ago, with the credible interval ranging between 81.7 and 107.7 million years ago (Ghezelayagh et al. 2022).

  • img-z151-8_03.gif

    Perciformes A. Günther 1880:374–397
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown lineage that contains Trachypoma macracanthus Günther 1859, Cephalopholis cruentata (Lacépède 1802), Acanthistius cinctus (Günther 1859), Perca fluviatilis Linnaeus 1758, Cottus carolinae (T. N. Gill 1861b), and Sebastes norvegicus (Ascanius 1772). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek πέρκη (p̍ːke͡I), a name applied to many species of fishes by ancient authors (D. W. Thompson 1947:195–197). The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 968.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, figs. S16–S19). Phylogenetic relationships of the major lineages of Perciformes are presented in Figure 18.

  • Phylogenetics. Throughout the 20th century, most species and lineages of Percomorpha were classified in the catch-all taxon Perciformes (Goodrich 1909:472–490; Berg 1940; McAllister 1968:136–148; Gosline 1971:156–164; A. C. Gill and Mooi 2002; J. S. Nelson 2006:339–441). Any lineage of Percomorpha that was not as morphologically distinctive as flatfishes, pufferfishes, sticklebacks, or flying gurnards was relegated to Perciformes. At the turn of the century, more than half of all species of Percomorpha were classified in Perciformes, a grouping that was considered an assemblage of unrelated lineages (Figure 1; G. D. Johnson 1984, 1993; G. D. Johnson and Patterson 1993; J. S. Nelson 2006:340; J. S. Nelson et al. 2016:430). Molecular phylogenies revealed that lineages traditionally classified in Perciformes are distributed among 11 of the 13 major clades of Percomorpha (Figures 1 and 2). The two earliest branching lineages, Ophidiiformes and Batrachoididae, are the only clades of Percomorpha that do not include lineages previously classified in Perciformes (Figure 1; Dornburg and Near 2021; Ghezelayagh et al. 2022).

  • The unraveling of the traditional composition of Perciformes began with the first molecular phylogenetic studies of Percomorpha (W.-J. Chen et al. 2003; Miya et al. 2003, 2005; Dettaï and Lecointre 2004; W. L. Smith and Wheeler 2004, 2006; Dettaï and Lecointre 2005, 2008; W. L. Smith and Craig 2007; B. Li et al. 2009; Malmstrøm et al. 2016, 2017). Phylogenetic studies that sampled most of the major lineages of Percomorpha resulted in the resolution of a clade delimited as Perciformes that included Percidae (perches, walleyes, darters) and all lineages previously classified as Scorpaeniformes (e.g., sculpins, rockfishes, scorpionfishes) (J. S. Nelson 2006:318–339); Zoarcoidea (e.g., eelpouts, ronquils, pricklebacks); all lineages of Serranidae (seabasses) except for Hemilutjanus macrophthalmos (Grape-eye Seabass) and Caesioscorpis theagenes (Blowhole Perch) (P. Parenti and Randall 2020; W. L. Smith et al. 2022); Bembropidae (duckbills); Percophis brasiliensis (Brazilian Flathead); Trachinidae (weeverfishes) traditionally classified in Trachinoidei or Trachiniformes (J. S. Nelson 2006:403–409; J. S. Nelson et al. 2016:421–427); the southern cold temperate and Antarctic marine Notothenoidei (e.g., icefishes, notothens, plunderfishes); and Gasterosteidae (sticklebacks) (Matschiner et al. 2011; Betancur-R, Broughton, et al. 2013; Lautredou et al. 2013; Near et al. 2013, 2015; Thacker et al. 2015; Betancur-R et al. 2017; Hughes et al. 2018; Ghezelayagh et al. 2022; T. Tang et al. 2023).

  • Serranidae is traditionally delimited to include more than 450 species classified among 73 genera (Fricke et al. 2023) and is consistently resolved as nonmonophyletic in Perciformes (Craig and Hastings 2007; W. L. Smith and Craig 2007; Lautredou et al. 2013; Near et al. 2013; Zhuang et al. 2013; Ma et al. 2016; Ghezelayagh et al. 2022). A phylogenomic analysis of UCE loci resolves lineages traditionally classified as Serranidae among Epinephelidae (groupers), Anthiadidae (basslets and anthians), Serranidae (sensu stricto), Acanthistius (wirrahs), and Niphon spinosus (Sawedged Perch) (W. L. Smith and Craig 2007; Near et al. 2015; Ghezelayagh et al. 2022).

  • There are seven major clades of Perciformes: (1) Epinephelidae, (2) Anthiadidae, (3) Acanthistius, (4) a clade containing Bembropidae and Serranidae, (5) Percoidei, (6) Notothenioidei, and (7) Scorpaenoidei. In phylogenomic analyses of UCE loci, Epinephelidae, Anthiadidae, and Acanthistius are resolved as successive early branching lineages of Perciformes, the clade containing Serranidae and Bembropidae is resolved as the sister lineage of Percoidei, and Notothenioidei and Scorpaenoidei are resolved as sister groups (Figure 18; Ghezelayagh et al. 2022).

  • Composition. There are currently 3,200 species of Perciformes (Fricke et al. 2023) classified in Acanthistius, Anthiadidae, Bembropidae, Epinephelidae, Notothenioidei, Scorpaenoidei, and Serranidae. Over the past 10 years, there have been 209 new living species of Perciformes described (Fricke et al. 2023), comprising 6.5% of the living species diversity in the clade.

  • Diagnostic apomorphies. There are no known morphological apomorphies for Perciformes; however, a backwardly directed opercle spine is a potential apomorphy as it is present in Acanthistius, Anthiadidae, Bembrops (Bembropidae), Epinephelidae, Channichthyidae (Notothenioidei), Niphon and Trachinidae (Percoidei), Scorpaenoidei sans the traditional cottoid lineages, and Serranidae (G. D. Johnson 1983; Iwami 1985; Imamura and Yabe 2002; W. L. Smith et al. 2018).

  • Synonyms. Serraniformes (B. Li et al. 2009, tbl. 4; Lautredou et al. 2013:140–141) and Scorpaeniformes (W. L. Smith and Busby 2014:333; Sparks et al. 2014, fig. 2; Davis et al. 2016, figs. 1, 4; W. L. Smith et al. 2016, sup. fig. 1, 2018, fig. 2B) are ambiguous synonyms of Perciformes.

  • Comments. The name Perciformes is applied to this clade in several studies and was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade (Betancur-R, Broughton, et al. 2013; Near et al. 2013; Betancur-R et al. 2017; Dornburg and Near 2021; Ghezelayagh et al. 2022).

  • The earliest fossil Perciformes are two species of Scorpaenoidei dated to the Ypresian (56.0–48.1 Ma), the New Zealand platycephalid otolith taxon †Platycephalus parapercoides (Schwarzhans 2019) and †Eosynanceja brabantica from Belgium that is classified as a species of Synanceiidae (Casier 1946). Bayesian relaxed molecular clock analyses of Perciformes result in an average posterior crown age estimate of 66.9 million years ago, with the credible interval ranging between 55.2 and 78.0 million years ago (Ghezelayagh et al. 2022).

  • img-z154-5_03.gif

    FIGURE 18.

    Phylogenetic relationships of the major living lineages and fossil taxa of Perciformes, Percoidei, Notothenioidei, Scorpaenoidei, Scorpaenoidea, Cottoidea, and Zoarcoidea. Filled circles identify the common ancestor of clades with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z152-1_03.jpg

    Percoidei L. J. F. J. Fitzinger 1832:331
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive clade containing Perca fluviatilis (Linnaeus 1758), Trachinus radiatus Cuvier 1829, and Niphon spinosus Cuvier 1828 in Cuvier and Valenciennes (1828). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek πέρκη (p̍ːke͡I), a name applied to many species of fishes by ancient authors (D. W. Thompson 1947:195–197).

  • Registration number. 969.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S16). Phylogenetic relationships among the Percoidei are presented in Figure 18.

  • Phylogenetics. The first set of molecular phylogenetic studies that led to the unraveling of the traditional composition of Perciformes and Serranidae did not resolve the relationships of Niphon spinosus (Ara), Percidae (perches, walleyes, darters), or Trachinidae (weeverfishes) (W.-J. Chen et al. 2003; Dettaï and Lecointre 2004, 2005; W. L. Smith and Wheeler 2004; W. L. Smith and Wheeler 2006; W. L. Smith and Craig 2007; Dettaï and Lecointre 2008; B. Li et al. 2009). Two phylogenetic studies using Sanger-sequenced nuclear genes sampled two of the three lineages of Percoidei, one study resolving Niphon and Percidae as clade and the other resulting in Trachinidae and Percidae as a monophyletic group (Lautredou et al. 2013; Near et al. 2013). Subsequent molecular studies that sampled Niphon, Percidae, and Trachinidae resolved the lineages as a monophyletic group and placed Trachinidae as the sister lineage to a clade containing Niphon and Percidae (Figure 18; Near et al. 2015; Thacker et al. 2015; Ghezelayagh et al. 2022).

  • Composition. There are currently 254 living species of Percoidei (Fricke et al. 2023) that include Niphon spinosus and species classified in Percidae and Trachinidae. Over the past 10 years, there have been seven new living species of Percoidei described (Fricke et al. 2023), comprising 2.8% of the living species diversity in the clade.

  • Diagnostic apomorphies. There are no known morphological apomorphies for Percoidei.

  • Synonyms. There are no synonyms of Percoidei.

  • Comments. The name Percoidei is applied to this clade in several recent studies (Dornburg and Near 2021; Ghezelayagh et al. 2022).

  • The earliest fossil Percoideiis the otolith taxon †Trachinus falcatus from the Lutetian (48.1–41.0 Ma) of Germany (Schwarzhans 2007). Bayesian relaxed molecular clock analyses of Percoidei result in an average posterior crown age estimate of 52.7 million years ago, with the credible interval ranging between 38.7 and 65.4 million years ago (Ghezelayagh et al. 2022). Niphonidae is a valid family-group name under the International Code of Zoological Nomenclature (Jordan 1923:191; Van der Laan et al. 2014:72).

  • img-z155-2_03.gif

    Notothenioidei P. H. Greenwood, D. E. Rosen, S. H. Weitzman, and G. S. Myers 1966:401
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive clade containing Percophis brasiliensis Quoy and Gaimard 1825, Bovichtus diacanthus (Carmichael 1819), Eleginops maclovinus (Cuvier in Cuvier and Valenciennes 1830), and Notothenia coriiceps Richardson 1844. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek νότoς (n̍o͡Ʊt̍o͡Ʊz) meaning south and -θεν (ð̍εn), a particle placed as a suffix to nouns indicating motion from a place.

  • Registration number. 970.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S17). Phylogenetic relationships among the major lineages of Notothenioidei are presented in Figure 18. The placement of the pan-eleginopid †Proeleginops grandeastmanorum is on the basis of inferences from morphology (Balushkin 1994).

  • Phylogenetics. The traditional delimitation of Notothenioidei comprising Bovichtidae (thornfishes), Pseudaphritis urvillii (Congoli), Eleginops maclovinus (Patagonian Blennie), Nototheniidae (notothens), Harpagiferidae (plunderfishes), Bathydraconidae (Antarctic dragonfishes), and Channichthyidae (crocodile icefishes) was established in the first part of the 20th century (Dollo 1904; Regan 1913a, 1914b; Norman 1938a). Phylogenetic analyses of morphological and molecular characters resolve Percophis brasiliensis (Brazilian Flathead), Bovichtidae, and Pseudaphritis as successive branches from the lineage leading to a clade containing Eleginops and all species traditionally classified in Nototheniidae, Harpagiferidae, Bathydraconidae, and Channichthyidae (Balushkin 1992, 2000; Lecointre et al. 1997; Bargelloni et al. 2000; Dettaï and Lecointre 2004; Near, Pesavento, et al. 2004; Voskoboinikova 2004; Dettaï et al. 2012; Near, Dornburg, et al. 2012; Near et al. 2015, 2018; Ghezelayagh et al. 2022; Bista et al. 2023). Harpagiferidae, which includes species previously classified in Artedidraconidae (E. Parker and Near 2022), is the sister lineage of a clade containing Bathydraconidae and Channichthyidae (Iwami 1985; Hureau 1986; Balushkin 1992, 2000; Hastings 1993; Near et al. 2015, 2018; Ghezelayagh et al. 2022; Bista et al. 2023); however, in some morphological and molecular phylogenetic analyses Bathydraconidae is paraphyletic relative to Channichthyidae (Hastings 1993; Derome et al. 2002; Near, Dornburg, et al. 2012; Near et al. 2015; Ghezelayagh et al. 2022). The traditional delimitation of Nototheniidae is paraphyletic and comprises five disparately related lineages: Pleuragramma antarcticum (Antarctic Silver-fish), a clade containing Aethotaxis mitopteryx (Longfin Icedevil) and Dissostichus (tooth-fishes), Trematominae (notoperches), Gobionotothen (goby rockcods), and a clade containing Paranotothenia and Notothenia (Dettaï et al. 2012; Near et al. 2015, 2018).

  • Composition. There are currently 102 species of Notothenioidei (Eastman and Eakin 2021; E. Parker et al. 2022) that include Aethotaxis mitopteryx, Gvozdarus, Eleginops maclovinus, Percophis brasiliensis, Pseudaphritis urvillii, and species classified in Bovichtidae, Dissostichus, Trematominae, Gobionotothen, Nototheniidae, Harpagiferidae, Bathydraconidae, and Channichthyidae. Fossil lineages of Notothenioidei include the pan-eleginopid †Proeleginops grandeastmanorum (Appendix 1; Balushkin 1994). Over the past 10 years, there has been one new species of Notothenioidei described (Eastman and Eakin 2021), comprising 0.9% of the species diversity in the clade.

  • Diagnostic apomorphies. There are no known morphological apomorphies for Notothenioidei (Near et al. 2015).

  • Synonyms. Nototheniiformes (Regan 1913a:249–251; Jordan 1923:228; Norman 1938a:7–8) and Notothenioidae (Berg 1940:479; Gosline 1968:57–58, 1971:158) are approximate synonyms of Notothenioidei.

  • Comments. The traditional delimitation of Notothenioidei (Norman 1938a) was expanded to include Percophis brasiliensis (Near et al. 2015, 2018; Dornburg and Near 2021; Ghezelayagh et al. 2022).

  • The earliest fossil Notothenioidei is the pan-eleginopid †Proeleginops grandeastmanorum from the Ypresian (56.0–48.1 Ma) of Seymour Island, Antarctica (Balushkin 1994; Bieńkowska-Wasiluk et al. 2013). Bayesian relaxed molecular clock analyses of Notothenioidei result in an average posterior crown age estimate of 51.7 million years ago, with the credible interval ranging between 37.4 and 63.6 million years ago (Ghezelayagh et al. 2022).

  • img-z156-6_03.gif

    Scorpaenoidei P. Bleeker 1859:xxi
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive clade containing Platycephalus indicus (Linnaeus 1758), Normanichthys crockeri Clark 1937, Scorpaena porcus Linnaeus 1758, Sebastes norvegicus (Ascanius 1772), Cottus carolinae (T. N. Gill 1861b), and Zoarces elongatus Kner 1868. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek σκόρπαινα (skoːp̍e͡Ini͡ә), which is the name used by ancient authors (e.g., Aristotle and Oppian) in reference to the Mediterranean species Scorpaena porcus Linnaeus (Black Scorpionfish) and S. scrofa Linnaeus (Red Scorpionfish) (D. W. Thompson 1947:245–246).

  • Registration number. 971.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, figs. S17–S19). Although Scorpaena porcus is not included in the reference phylogeny, it resolves with other species of Scorpaena and Scorpaenidae in phylogenetic analyses of mtDNA sequences (Keskin and Atar 2013, fig. 2B; Yedier and Bostanci 2022, fig. 3). Phylogenetic relationships among the major lineages of Scorpaenoidei are presented in Figure 18.

  • Phylogenetics. Scorpaeniformes (sensu Greenwood et al. 1966), or mail-cheeked fishes, was a long-recognized taxonomic grouping of a range of lineages that included Scorpaenidae (scorpionfishes), Platycephalidae (flatheads), Hexagrammidae (greenlings), Cottidae (sculpins), Cyclopteridae (lumpfishes), Liparidae (snailfishes), Dactylopteridae (flying gurnards), and others (e.g., T. N. Gill 1889; Regan 1913c; Gregory 1933; Greenwood et al. 1966; Washington et al. 1984; Imamura and Shinohara 1998; J. S. Nelson 2006:318–339). Morphological and molecular studies confirm that the traditional delimitation of Scorpaeniformes is not monophyletic (Imamura and Yabe 2002; W.-J. Chen et al. 2003; Miya et al. 2003, 2005; W. L. Smith and Wheeler 2004, 2006; Imamura et al. 2005; W. L. Smith 2005; W. L. Smith and Craig 2007; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Lautredou et al. 2013; Near et al. 2013; Betancur-R et al. 2017). Specifically, Trichodontidae (sanddivers), traditionally classified in the polyphyletic Trachinoidei (e.g., Pietsch 1989; Pietsch and Zabetian 1990; J. S. Nelson et al. 2016:423), is closely related to the cottoids (Figure 18; Imamura et al. 2005; W. L. Smith 2005; Ghezelayagh et al. 2022); Dactylopteridae is distantly related to lineages of Scorpaenoidei and is phylogenetically nested in Syngnathiformes (Figure 15; Imamura 2000; W. L. Smith and Wheeler 2004; W. L. Smith and Craig 2007; B. Li et al. 2009; Betancur-R, Broughton, et al. 2013; Lautredou et al. 2013; Near et al. 2013; Smith et al. 2016; Betancur-R et al. 2017; Santaquiteria et al. 2021; Ghezelayagh et al. 2022); and Zoarcoidea (eelpouts), Gasterosteidae (sticklebacks), and Cottoidea (sculpins, lumpsuckers, greenlings) resolve as a monophyletic group (Figure 18; Imamura and Yabe 2002; W.-J. Chen et al. 2003; Miya et al. 2003, 2005; W. L. Smith and Wheeler 2004; W. L. Smith and Craig 2007; B. Li et al. 2009; Lautredou et al. 2013; Near et al. 2013; Davis et al. 2016; Smith et al. 2016; Betancur-R et al. 2017; Ghezelayagh et al. 2022; Maduna et al. 2022; L. Liu et al. 2023). An examination of myology and osteology led to the hypothesis that Champsodon (gapers) are related to Scorpaenoidei (Mooi and Johnson 1997); however, molecular analyses consistently resolve Champsodon as a lineage of Acropomatiformes that is distantly related to scorpaenoids (Figure 18; Near et al. 2013, 2015; Sanciangco et al. 2016; Betancur-R et al. 2017; Ghezelayagh et al. 2022).

  • Platycephalidae was long classified in Scorpaenoidei, but classifications differ in placing the group as the sole lineage in a suprafamilyranked taxon (Regan 1913c; Quast 1965; W. L. Smith 2005), with Bembridae (deepwater flatheads) and Hoplichthys (ghost flatheads) (Washington et al. 1984; Shinohara 1994), or with Plectrogeniidae (stinger flatheads), Triglidae (searobins), Bembridae, and Hoplichthys (W. L. Smith and Wheeler 2004). A morphological phylogeny and a phylogenetic analysis of combined morphological and molecular characters resolved Platycephalidae as the sister lineage of a clade containing Triglidae and Hoplichthys (Imamura 1996, 2004; W. L. Smith et al. 2018). Molecular phylogenies vary in the resolution of Platycephalidae: as the sister lineage of a clade containing Congiopodidae (horsefishes), Bembridae, and Scorpaenoidea (Lautredou et al. 2013); as the sister lineage of all other Scorpaenoidei except Hoplichthys (Near et al. 2013); nested in a clade that includes Hoplichthys and Bembridae that is the sister lineage of all other Scorpaenoidei (Betancur-R et al. 2017); or as the sister lineage of all other Scorpaenoidei (Figure 18; Ghezelayagh et al. 2022).

  • Triglidae was traditionally classified with other lineages of Scorpaenoidea (Regan 1913c; Quast 1965; Greenwood et al. 1966; W. L. Smith and Wheeler 2004). Many molecular phylogenies indicate Triglidae is more closely related to cottoids: as the sister lineage of a clade containing Gasterosteidae, Zoarcoidea, and Cottoidea (W. L. Smith and Craig 2007; B. Li et al. 2009; Lautredou et al. 2013) or as the sister lineage of a clade including Anoplopomatidae, Gasterosteidae, Zoarcoidea, and Cottoidea (Figure 18; Ghezelayagh et al. 2022). Other molecular phylogenies resolve Triglidae as the sister lineage of Bembropidae (duckbill flatheads) (Near et al. 2013; Betancur-R et al. 2017). Morphological phylogenies resolve Triglidae as paraphyletic relative to Hoplichthys (Imamura 1996, 2004), but a phylogeny inferred from a combined molecular and morphological dataset places Triglidae and Hoplichthys as sister lineages (W. L. Smith et al. 2018).

  • Anoplopomatidae (sablefishes) was placed with Hexagrammidae in several classifications (Regan 1913c; Quast 1965; Greenwood et al. 1966; Washington et al. 1984). A morphological phylogeny resolves Anoplopomatidae as the sister lineage of a clade containing all other sampled lineages of Cottoidea (Imamura et al. 2005). Molecular phylogenies resolve Anoplopomatidae as nested within Cottoidea as the sister lineage of Zaniolepididae (combfishes) (W. L. Smith and Craig 2007), as the sister lineage of all other Cottoidea (W. L. Smith and Wheeler 2004; Near et al. 2013), or as the sister lineage of a clade containing Cottoidea, Gasterosteidae, and Zoarcoidea (Figure 18; Betancur-R et al. 2017; Ghezelayagh et al. 2022; Maduna et al. 2022; L. Liu et al. 2023).

  • Composition. There are currently 2,208 living species of Scorpaenoidei (Fricke et al. 2023) classified in Anoplopomatidae, Bembridae, Cottoidea, Gasterosteidae, Platycephalidae, Scorpaenoidea, and Triglidae. Over the past 10 years, there have been 142 new living species of Scorpaenoidei described (Fricke et al. 2023), comprising 6.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Scorpaenoidei include (1) third circumorbital modified as a suborbital stay with distal end broad and strongly attached to preopercle (Cuvier 1829:158; Boulenger 1904a, 1904b:692; Greenwood et al. 1966; Lauder and Liem 1983; Bowne 1994; Imamura and Yabe 2002; Imamura 2004; W. L. Smith et al. 2018), (2) presence of an extrinsic gas bladder muscle connected anteriorly to neurocranium and posteriorly to vertebrae (Imamura and Yabe 2002; Imamura 2004; W. L. Smith et al. 2018), (3) absence of supraneurals (W. L. Smith et al. 2018), and (4) hypurals 3 and 4 fused (W. L. Smith et al. 2018).

  • Synonyms. Cottoidea (T. N. Gill 1872:6), Scleroparei (Boulenger 1904a:184–185, 1904b:692–702, fig. 399; Regan 1913c), Cataphracti (Jordan 1923:208–215), Cottoidei (Berg 1940:487–490; McAllister 1968:148), and Cottida (Matsubara 1955:1040–1048) are partial synonyms of Scorpaenoidei. Scorpaeniformes is both a partial (Goodrich 1909:449–454; Greenwood et al. 1966:399; Gosline 1971:167–168; Washington et al. 1984:438) and an approximate (J. S. Nelson et al. 2016:467–495) synonym of Scorpaenoidei.

  • Comments. Scorpaenoidei was the name applied to a paraphyletic group that contained Platycephalidae, Scorpaenoidea, Bembridae, and Triglidae, but excluded Anoplopomatidae, Cottoidea, Gasterosteidae, Normanichthys, and Zoarcoidea (W. L. Smith et al. 2018).

  • The earliest fossil Scorpaenoidei is the pan-synanceiid †Eosynanceja brabantica from the Ypresian (56.0–48.1 Ma) of Belgium (Casier 1946). Bayesian relaxed molecular clock analyses of Scorpaenoidei result in an average posterior crown age estimate of 59.3 million years ago, with the credible interval ranging between 50.2 and 67.8 million years ago (Ghezelayagh et al. 2022).

  • img-z158-6_03.gif

    Scorpaenoidea T. N. Gill 1889:579
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive clade containing Hoplichthys langsdorfii Cuvier 1829 in Cuvier and Valenciennes (1829b), Congiopodus leucopaecilus (Richardson 1846), Inimicus didactylus (Pallas 1769), Scorpaena porcus Linnaeus 1758, and Sebastes norvegicus (Ascanius 1772), but not Platycephalus indicus (Linnaeus 1758). This is a minimum-crown-clade definition with an external specifier.

  • Etymology. From the ancient Greek σκόρπαινα (skoː͡ɹp̍e͡Ini͡ә), which is the name used by ancient authors (e.g., Aristotle and Oppian) in reference to the Mediterranean species Scorpaena porcus Linnaeus (Black Scorpionfish) and S. scrofa Linnaeus (Red Scorpionfish) (D. W. Thompson 1947:245–246).

  • Registration number. 972.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, fig. S17). Although Scorpaena porcus is not included in the reference phylogeny, it resolves with other species of Scorpaena and Scorpaenidae in phylogenetic analyses of mtDNA sequences (e.g., Yedier and Bostanci 2022, fig. 3). The phylogenetic relationships of the major lineages of Scorpaenoidea are presented in Figure 18. The placement of Plectrogeniidae (stinger flatheads) follows a phylogenetic analysis of a combined dataset of morphological and molecular characters (W. L. Smith et al. 2018).

  • Phylogenetics. The monophyly of Scorpaenoidea is supported in phylogenomic analyses of UCE loci (Ghezelayagh et al. 2022). Morphological studies result in phylogenies that do not resolve Congiopodidae (horsefishes), Hoplichthys (ghost flatheads), or Normanichthys crockeri (Mote Sculpin) within Scorpaenoidea (Ishida 1994; Imamura 2004) and place Triglidae (searobins), Bembridae (deepwater flatheads), and Platycephalidae (flatheads) as phylogenetically nested in Scorpaenoidea (Imamura 2004). Molecular phylogenies inferred from Sanger-sequenced mitochondrial and nuclear genes do not resolve Scorpaenoidea as monophyletic; specifically, Bembridae, Percoidei, Triglidae, Cottoidea, Platycephalidae, and Serranidae are nested within Scorpaenoidea (W. L. Smith and Craig 2007); Bembridae is nested within Scorpaenoidea and resolved as the sister lineage of Synanceiidae (stonefishes) (Lautredou et al. 2013), Hoplichthys is placed outside of Scorpaenoidea as the sister lineage of a clade containing Bembridae and Platycephalidae (Betancur-R et al. 2017), and Hoplichthys and Congiopodidae are placed outside of Scorpaenoidea, respectively as the sister lineages of Triglidae and Cottoidea (W. L. Smith et al. 2018).

  • The phylogenetic relationships of Normanichthys crockeri were unresolved from the time the species was described (Clark 1937) to the application of phylogenomic datasets to investigate relationships of Acanthomorpha (Ghezelayagh et al. 2022). On the basis of morphology, Normanichthys was hypothesized to be closely related to lineages of Cottoidea (Norman 1938b; Berg 1940:489; Fowler 1951; Greenwood et al. 1966) or a distinct lineage within Scorpaenoidei (Washington et al. 1984; Yabe and Uyeno 1996; J. S. Nelson et al. 2016:478). Larval morphology was the basis to suggest the phylogenetic placement of Normanichthys outside Scorpaenoidei (Velez et al. 2003). A maximum parsimony analysis of DNA sequences from mitochondrial and nuclear genes resolved Normanichthys as the sister lineage of an unlikely clade containing Hoplichthys, Synanceiidae, and the ovalentarians Gramma, Menidia, Labrisomus, and Salarias (W. L. Smith and Wheeler 2004). Phylogenomic analyses of UCE loci resolve Normanichthys and Hoplichthys as a clade that is the sister lineage of all other Scorpaenoidea (Figure 18; Ghezelayagh et al. 2022).

  • Composition. There are currently 579 species of Scorpaenoidea (Fricke et al. 2023) that include Normanichthys crockeri and species classified in Congiopodidae, Hoplichthys, Neosebastidae (gurnard perches), Plectrogeniidae (stinger flatheads), Scorpaenidae (scorpionfishes), and Synanceiidae. Over the past 10 years, there have been 44 new living species of Scorpaenoidei described (Fricke et al. 2023), comprising 7.6% of the living species diversity in the clade. Diagnostic apomorphies. There are no known morphologicalapomorphiesforScorpaenoidea.

  • Synonyms. Scorpaeniformes (Bleeker 1859:xxi; Regan 1913c:170–171; W. L. Smith 2005:153), Scorpaenoidae (Berg 1940:488; Quast 1965:587–589), Scorpaenicae (Matsubara 1955:1040–1048), and Scorpaenoidei (Greenwood et al. 1966:399; Washington et al. 1984:439; J. S. Nelson et al. 2016:468–475) are all partial synonyms of Scorpaenoidea.

  • Comments. The group name Scorpaenoidea has been applied to several paraphyletic groups of species classified in Scorpaenoidei. The first use of Scorpaenoidea was for a group containing Synanceiidae, Scorpaenidae, Hexagrammidae, and Anoplopomatidae (T. N. Gill 1889). By the early 20th century, Platycephalidae, Hoplichthys, Plectrogeniidae, Neosebastidae, Scorpaenidae, Synanceiidae, Bembridae, and Triglidae were included in Scorpaenoidea (Imamura 2004).

  • The earliest fossil Scorpaenoidea is the pan-synanceiid †Eosynanceja brabantica from the Ypresian (56.0–48.1 Ma) of Belgium (Casier 1946). Bayesian relaxed molecular clock analyses of Scorpaenoidea result in an average posterior crown age estimate of 52.3 million years ago, with the credible interval ranging between 41.3 and 62.7 million years ago (Ghezelayagh et al. 2022). Scorpaenoidea is a valid family-group name under the International Code of Zoological Nomenclature (Van der Laan et al. 2014:83).

  • img-z159-8_03.gif

    Cottoidea T. N. Gill 1872:6
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive clade that contains Cottus gobio Linnaeus, Cottus carolinae (T. N. Gill 1861b), Zaniolepis latipinnis C. Girard 1858b, Hexagrammos decagrammus (Pallas 1810), and Eumicrotremus orbis (Günther 1861). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek κόττoς (k̍αːt̍ʌs), which is the name used by ancient authors (e.g., Aristotle) in reference to Cottus gobio Linnaeus (European Bullhead) (D. W. Thompson 1947:128–129).

  • Registration number. 973.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, fig. S18). Although Cottus gobio is not in the reference phylogeny, several phylogenetic studies based on mtDNA, DNA sequences from nuclear genes, and morphology nest C. gobio in a clade with other species of Cottus (Kontula et al. 2003, fig. 2; Yokoyama and Goto 2005, fig. 1; Lautredou et al. 2013, fig. 3; W. L. Smith and Busby 2014, fig. 3). Phylogenetic relationships of the major living and fossil lineages of Cottoidea are presented in Figure 18. The placement of Jordaniidae (longfin sculpins) follows a phylogenetic analysis of a combined dataset of morphological and molecular characters (W. L. Smith and Busby 2014). The placement of the fossil pan-hexagrammids †Sakhalinia and †Paraophiodon are on the basis of inferences from morphology (Nazarkin 1997; Nazarkin et al. 2013).

  • Phylogenetics. The cottoids were traditionally delimited to include Scorpaenichthys marmoratus (Cabezon) and species now classified in Rhamphocottidae (grunt sculpins), Agonidae (poachers), Cottidae (sculpins), Jordaniidae, and Psychrolutidae (fathead sculpins) (Greenwood et al. 1966; Yabe 1985; Jackson 2003; W. L. Smith 2005; W. L. Smith and Busby 2014). Morphological and molecular analyses consistently resolve Hexagrammidae (greenlings), Zaniolepididae (combfishes), Trichodontidae (sandfishes), Cyclopteridae (lumpfishes), Liparidae (snailfishes), and Anoplopomatidae as closely related to the cottoids (Washington et al. 1984; Yabe 1985; Shinohara 1994; W. L. Smith and Wheeler 2004; Imamura et al. 2005; W. L. Smith and Craig 2007; B. Li et al. 2009; Lautredou et al. 2013; Near et al. 2013; W. L. Smith and Busby 2014; Betancur-R et al. 2017). However, phylogenetic analysis of whole mtDNA genomes and UCE loci resolves Anoplopomatidae as the sister lineage of a more inclusive clade containing Cottoidea, Gasterosteidae (sticklebacks), and Zoarcoidea (Figure 18; Ghezelayagh et al. 2022; Maduna et al. 2022; L. Liu et al. 2023). Within Cottoidea, Zaniolepididae, Hexagrammidae, and a clade containing Trichodontidae, Cyclopteridae, and Liparidae are successive branching lineages leading to a core cottoid clade containing Rhamphocottidae, S. marmoratus, Agonidae, Cottidae, and Psychrolutidae (Figure 18). Phylogenetic analysis of a combined morphological and molecular dataset resolves Jordaniidae as the sister lineage to all other core cottoids (W. L. Smith and Busby 2014).

  • Composition. There are currently 897 species of Cottoidea (Fricke et al. 2023) that include Scorpaenichthys marmoratus and species classified in Agonidae, Cottidae, Cyclopteridae, Hexagrammidae, Jordaniidae, Liparidae, Psychrolutidae, Rhamphocottidae, Trichodontidae, and Zaniolepididae. Fossil lineages of Cottoidea include the pan-hexagrammids †Sakhalinia multispinata and †Paraophiodon nessovi from the Serravallian (13.82–11.63 Ma) of Russia (Nazarkin 1997; Nazarkin et al. 2013). Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, 61 new living species of Scorpaenoidei have been described (Fricke et al. 2023), comprising 6.8% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological and reproductive apomorphies for Cottoidea include (1) presence of a lachryopalatine articulation (Yabe 1985; Shinohara 1994; Imamura et al. 2005), (2) parhypural and lower hypural plate fused (Yabe 1985; Shinohara 1994; Imamura et al. 2005), (3) third and fourth hypurals fused (Shinohara 1994; Imamura et al. 2005), (4) spawn adhesive demersal eggs (Watson et al. 1984; Shinohara 1994; Imamura et al. 2005; Muñoz 2010), and (5) absence of a connection between preopercle and the temporal sensory canals (Imamura et al. 2005).

  • Synonyms. Cottiformes (Regan 1913c:171–172; Jordan 1923:211–215), Cottoidae (Berg 1940:489–490; Quast 1965:595–597), Cotticae (Matsubara 1955:1040–1048), Cottoidei (Greenwood et al. 1966:399; Washington et al. 1984:444–445; Shinohara 1994:80; Imamura et al. 2005:274; W. L. Smith 2005:153; J. S. Nelson et al. 2016:485–494), and Cottales (Betancur-R et al. 2017:31) are all partial synonyms of Cottoidea.

  • Comments. The name Cottoidea was applied to a less inclusive group that included Agonidae, Cottidae, Jordaniidae, Psychrolutidae, Rhamphocottidae, and Scorpaenichthys (W. L. Smith and Busby 2014).

  • TheearliestfossilCottoideais†Cottusotiakensis, an otolith taxon from the Late Oligocene (27.30–23.04 Ma) of New Zealand (Frost 1928; PBDB 2023). Bayesian relaxed molecular clock analyses of Cottoidea result in an average posterior crown age estimate of 40.5 million years ago, with the credible interval ranging between 33.8 and 49.0 million years ago (Ghezelayagh et al. 2022). Cottoidea is a valid family-group name under the International Code of Zoological Nomenclature (Van der Laan et al. 2014:87).

  • img-z161-4_03.gif

    Gasterosteidae C. L. Bonaparte 1831:156, 169
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown lineage that contains Hypoptychus dybowskii Steindachner 1880, Aulichthys japonicus Brevoort in T. N. Gill 1862, Aulorhynchus flavidus T. N. Gill 1861c, Gasterosteus aculeatus Linnaeus 1758, and Apeltes quadracus (Mitchill 1815). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek γαστήρ (̍æstɚ) meaning belly and ὀστέoν (̍αːstIәn) meaning bone.

  • Registration number. 974.

  • Reference phylogeny. A phylogeny of 10 species of Gasterosteidae inferred from a supermatrix of 27 nuclear and mitochondrial genes (Rabosky et al. 2018; J. Chang et al. 2019). The phylogeny is available on the Dryad data repository (Rabosky et al. 2019). The phylogenetic resolution of Gasterosteidae within Perciformes is shown in Figure 18.

  • Phylogenetics. The identification of a natural group containing species traditionally classified in Gasterosteidae (sticklebacks) and Aulorhynchidae (tubesnouts) is reflected in precladistic classifications from the second half of the 19th century (T. N. Gill 1862, 1872; Jordan and Evermann 1896:742–753). Long classified with species of Ammodytidae (sand lances) (Jordan 1923:230; Berg 1940:481; Gosline 1963b; Greenwood et al. 1966), a study citing osteology and reproductive traits proposed Hypoptychus dybowskii (Korean Sand-lance) as most closely related to Gasterosteidae (Ida 1976). While not universally accepted initially (J. S. Nelson 1978, 1984:245), morphological and molecular phylogenetic analyses resolve a clade containing Hypoptychus and all other sampled species of Gasterosteidae (Pietsch 1978; Kawahara et al. 2008; Near et al. 2013; Betancur-R et al. 2017; Ghezelayagh et al. 2022). A study of the osteology of the oral jaws concluded that the traditional delimitation of Aulorhynchidae is paraphyletic because Aulichthys japonicus (Tubenose) and Hypoptychus are more closely related relative to Aulorhynchus flavidus (Tubesnout) (G. D. Johnson and Patterson 1993), a result supported in molecular phylogenetic analyses of Sanger-sequenced mtDNA and nuclear genes (Betancur-R et al. 2017; Rabosky et al. 2018). However, phylogenetic analyses of Sanger-sequenced nuclear genes and a phylogenomic analysis of ultraconserved element loci do not resolve a clade containing Hypoptychus and Aulichthys, but rather place Hypoptychus as the sister lineage of all other Gasterosteidae (Near et al. 2013; Ghezelayagh et al. 2022). Two phylogenetic studies of Gasterosteidae did not test the monophyly of Aulorhynchidae: an analysis of morphological characters did not sample Hypoptychus but resolved Aulichthys and Aulorhynchus as sister lineages (Orr 1995), and analysis of Sanger-sequenced whole mtDNA genomes and nuclear genes did not test the monophyly of Aulorhynchidae because the inferred phylogenies were rooted with Hypoptychus (Kawahara et al. 2009).

  • Phylogenetic relationships inferred among the core gasterosteiids, Apeltes quadracus (Four-spine Stickleback), Culaea inconstans (Brook Stickleback), Gasterosteus (threespine sticklebacks), Pungitius (ninespined sticklebacks), and Spinachia spinachia (Sea Stickleback) vary among studies and types of data analyzed. Phylogenies inferred from morphology, behavior, and mtDNA gene trees resolve Spinachia as the sister lineage of all other core gasterosteiids (McLennan 1993; Bowne 1994; McLennan and Mattern 2001; Keivany and Nelson 2004; Mattern 2004, 2007; Mattern and McLennan 2004). However, molecular phylogenies inferred from combinations of mtDNA and nuclear genes resolve either Gasterosteus (Kawahara et al. 2009; Rabosky et al. 2018) or a clade containing Gasterosteus and Pungitius as the sister lineage of all other core gasterosteiids (Betancur-R et al. 2017). Regardless of the type of character data, most phylogenetic analyses are consistent in resolving Culaea and Pungitius as sister lineages (McLennan 1993; Bowne 1994; McLennan and Mattern 2001; Keivany and Nelson 2004; Mattern 2004; Mattern and McLennan 2004; Kawahara et al. 2009; Rabosky et al. 2018).

  • Composition. There are currently 23 species of Gasterosteidae (Fricke et al. 2023) that include Apeltes quadracus, Aulichthys japonicus, Aulorhynchus flavidus, Culaea inconstans, Hypoptychus dybowskii, Spinachia spinachia, and species classified in Gasterosteus and Pungitius. Over the past 10 years, there have been two new living species of Gasterosteidae described (Fricke et al. 2023), comprising 8.7% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Gasterosteidae include (1) loss of both upper circumorbital bones (W. L. Smith 2005), (2) nasal and neurocranium fused (W. L. Smith 2005), (3) medial extrascapulars absent (W. L. Smith 2005), (4) articular and ascending process of premaxilla continuous (W. L. Smith 2005), (5) palatine teeth absent (W. L. Smith 2005), (6) endopterygoid absent (W. L. Smith 2005), (7) four branchiostegal rays (W. L. Smith 2005), (8) basihyal uniform in width rostrally (W. L. Smith 2005), (9) loss of gill rakers from second, third, and fourth epibranchials (W. L. Smith 2005), (10) absence of Baudelot's ligament, (11) four pelvic fin rays (W. L. Smith 2005), (12) absence of anterior pelvic fin ray processes (W. L. Smith 2005), (13) dorsal spines absent from first two dorsal pterygiophores (W. L. Smith 2005), (14) second preural centrum with elongate neural spines (W. L. Smith 2005), (15) caudal hypurapophysis absent (W. L. Smith 2005), and (16) fused upper and lower hypural plates (W. L. Smith 2005).

  • Synonyms. Gasterosteoidei is a partial (Bleeker 1859:xxiii; Goodrich 1909:411–412) and approximate synonym (Greenwood et al. 1966:398; J. S. Nelson et al. 2016:482–485) of Gasterosteidae. Hemibranchii (Jordan 1923:173–174) and Gasterosteiformes (Berg 1940:458–460) are approximate synonyms of Gasterosteidae. Gasterosteales is an ambiguous synonym of Gasterosteidae (Betancur-R et al. 2017:31).

  • Comments. The species Aulichthys japonicus, Aulorhynchus flavidus, and Hypoptychus dybowskii are classified in Gasterosteidae, in contrast to the traditional classification of these species in Aulorhynchidae and Hypoptychidae (e.g., J. S. Nelson et al. 2016:482–483). This is motivated by the consistent resolution of Gasterosteidae as a monophyletic group (Pietsch 1978; Kawahara et al. 2008; Near et al. 2013; Betancur-R et al. 2017; Ghezelayagh et al. 2022). If Aulichthys japonicus is classified in Aulorhynchidae or Hypoptychidae, a rank-based classification would include two family group names for three species. Classifying all three of these species in Gasterosteidae reflects the most robust inferences of their phylogenetic relationships and reduces the number of redundant group names among ray-finned fishes. The name Gasterosteidae was selected as the clade name over its synonyms because it seems to be the name most frequently applied to a taxon approximating the named clade. Gasterosteidae is a valid family-group name under the International Code of Zoological Nomenclature (Van der Laan et al. 2014:64).

  • The earliest fossil Gasterosteidae is †Gasterosteus cf. aculeatus from the Serravallian (13.82–11.63 Ma) of California, USA (Bell et al. 2009). Bayesian relaxed molecular clock analyses of Gasterosteidae result in an average posterior crown age estimate of 30.9 million years ago, with the credible interval ranging between 23.1 and 39.6 million years ago (Ghezelayagh et al. 2022).

  • img-z163-3_03.gif

    Zoarcoidea O. A. Radchenko, I. A. Chereshnev, and A. V. Petrovskaya 2014:473
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive clade containing Azygopterus corallinus Andriashev and Makushok 1955, Bathymaster signatus Cope 1873, Stichaeus punctatus (Fabricus 1780), and Zoarces viviparus (Linnaeus 1758). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek ζωαρκής (z̍oː͡ɹkәz) meaning life supporting.

  • Registration number. 975.

  • Reference phylogeny. A phylogeny inferred from sequences of DNA sequences of Sanger-sequenced mitochondrial and nuclear genes (Hotaling et al. 2021, fig. S1). Phylogenetic relationships of the major lineages of living and fossil lineages of Zoarcoidea are presented in Figure 18. The placements of the pan-pholids †Agnevichthys gretchinae and †Palaeopholis laevis in the phylogeny are on the basis of an analysis of morphological characters (Nazarkin 2002).

  • Phylogenetics. Zoarcoidea as delimited here was first presented in mid-20th century prephylogenetic morphological studies (Makushok 1958; Gosline 1968). The monophyly of Zoarcoidea is supported in several morphological (Anderson 1984, 1994; Kiernan 1990; Imamura and Yabe 2002; Kwun 2013; Clardy 2014) and molecular (Near et al. 2013; Betancur-R et al. 2017; Ghezelayagh et al. 2022) phylogenetic studies. Two molecular phylogenetic analyses of Zoarcoidea provide dense taxon sampling but do not test monophyly of the clade because they each utilize a single outgroup taxon (Kwun and Kim 2013; Hotaling et al. 2021). Morphological phylogenies differ in resolving Ulvaria subbifurcata (Radiated Shanny) (Clardy 2014) or Bathymasteridae (ronquils) (Anderson 1984, 1994; Hilton et al. 2019) as the sister lineage of all other Zoarcoidea, and molecular studies differ in placing Bathymasteridae (Turanov et al. 2012; Radchenko, Chereshnev, Petrovskaya, et al. 2014; Radchenko 2015, 2016, 2017; Turanov et al. 2017; Hotaling et al. 2021) or Eulophiidae (spinous eelpouts) (Kwun 2013; Kwun and Kim 2013) as the earliest diverging lineages of Zoarcoidea.

  • Morphological (Clardy 2014) and molecular (Radchenko et al. 2010; Turanov et al. 2012; Chereshnev et al. 2013; Radchenko, Chereshnev, Petrovskaya, et al. 2014; Radchenko 2015, 2016, 2017; Betancur-R et al. 2017; Rutenko et al. 2019; Hotaling et al. 2021) phylogenetic analyses resolve the traditional delimitation of Stichaeidae (pricklebacks) (Makushok 1958; Mecklenburg and Sheiko 2004; Zemnukhov 2012; J. S. Nelson et al. 2016:480) as nonmonophyletic. Specifically, the enigmatic Graveldiver, Scytalina cerdale, long classified in the monotypic Scytalinidae (Mecklenburg 2003e; Hilton 2009), is nested in Stichaeidae as the sister lineage of a clade containing Phytichthys chirus (Ribbon Prickleback) and Xiphister in a phylogenomic analysis of ultraconserved element (UCE) loci (Ghezelayagh et al. 2022). Morphological studies resolve a clade containing Scytalina and Xiphister or place Scytalina as the sister lineage of a clade containing Phytichthys, Xiphister, Ptilichthys goodei (Quillfish), and Pholidae (gunnels) (Hilton 2009; Clardy 2014). At least five other lineages traditionally classified as Stichaeidae are more closely related to other lineages of Zoarcoidea in molecular and morphological phylogenetic analyses: Opisthocentridae (rearspined pricklebacks) is the sister lineage of a clade containing Ptilichthys and Pholidae (Figure 18; Radchenko et al. 2012; Chereshnev et al. 2013; Kwun 2013; Kwun and Kim 2013; Radchenko 2015, 2016; Rutenko et al. 2019; Hotaling et al. 2021; Ghezelayagh et al. 2022); Lumpenidae (eel pricklebacks) is the sister lineage of Cryptacanthodes (wrymouths) (Figure 18; Kwun and Kim 2013; Radchenko 2015, 2016; Hotaling et al. 2021; Ghezelayagh et al. 2022); Cebidichthyidae (monkeyfaces), including Cebidichthys, Dictyosoma, Esselenia, and Esselenichthys, was formerly classified in the stichaeid subclade Xiphisterinae but is now resolved as the sister lineage of all other Zoarcoidea to the exclusion of Bathymasteridae and possibly Eulophiidae (Figure 18; Radchenko et al. 2012; Turanov et al. 2012; Chereshnev et al. 2013; Radchenko, Chereshnev, Petrovskaya, et al. 2014; Radchenko 2015, 2016; Hotaling et al. 2021; Ghezelayagh et al. 2022); Neozoarcidae (kissing eelpouts) is resolved as sister to a clade containing Anarhichadidae (wolffishes) and Zoarcidae (eelpouts) (Figure 18; Radchenko 2015, 2016; Turanov et al. 2017; Hotaling et al. 2021; Ghezelayagh et al. 2022) or as the sister lineage of Zoarcidae (Clardy 2014); and Kasatkia, thought to be closely related to Opisthocentridae (Posner and Lavenberg 1999), is resolved as the sister lineage of Ptilichthys (Hotaling et al. 2021).

  • Composition. There are currently 431 species of Zoarcoidea (Mecklenburg 2003a, 2003b, 2003c, 2003d, 2003e, 2003f; Mecklenburg and Sheiko 2004; Fricke et al. 2023) that include Ptilichthys goodei, Zaprora silenus (Prowfish), and species classified in Anarhichadidae, Bathymasteridae, Cebidichthyidae, Cryptacanthodes, Eulophiidae, Lumpenidae, Neozoarcidae, Opisthocentridae, Pholidae, Stichaeidae, and Zoarcidae. Fossil taxa include the pan-pholids †Agnevichthys gretchinae and †Palaeopholis. Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, there have been 19 new living species of Zoarcoidea described (Fricke et al. 2023), comprising approximately 4.4% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Zoarcoidea include (1) basisphenoid absent (Anderson 1984, 1994; Imamura and Yabe 2002; Wiley and Johnson 2010), (2) a single pair of nostrils present due to loss of the posterior nostrils (Anderson 1984, 1994; Imamura and Yabe 2002; Wiley and Johnson 2010), (3) mesial portion of A2,3 section of adductor mandibulae extends posterior to levator arcus palatini (Anderson 1994), and (4) dorsal and anal fin stays absent (Imamura and Yabe 2002; Wiley and Johnson 2010).

  • Synonyms. Zoarceoidea (Gill 1893:136; Makushok 1958:34; Gosline 1971:158), Zoarcicae (Hubbs 1952:51, fig. 1), Zoarcoidei (Greenwood et al. 1966:397; Springer and Johnson 2004:209; Wiley and Johnson 2010:161; J. S. Nelson et al. 2016:478–482), and Zoarcales (Betancur-R et al. 2017:31) are ambiguous synonyms of Zoarcoidea.

  • Comments. Makushok (1958) is attributed as using the group name Zoarcoidea to delimit Zoarcidae and other closely related lineages (Radchenko, Chereshnev, and Petrovskaya 2014); however, Makushok (1958:34) used the name Zoarceoidea. The group name Zoarcoidea is used in reference to a taxonomic suborder without attribution (Fletcher et al. 1988). The application of molecular and morphological data to the phylogenetics of Zoarcoidea led to the discovery that the traditional delimitation of Stichaeidae was not monophyletic (e.g., Clardy 2014; Radchenko 2015; Betancur-R et al. 2017; Hotaling et al. 2021; Ghezelayagh et al. 2022), necessitating the addition of five families to Zoarcoidea in rank-based classifications (Fricke et al. 2023). The Graveldiver Scytalina cerdale is classified in Stichaeidae on the basis of the results of phylogenomic analyses of UCE loci (Ghezelayagh et al. 2022).

  • The earliest fossil Zoarcoidea include †Zaprora koreana from the Middle Miocene (16.0–11.6 Ma) of Korea (Nam and Nazarkin 2018) and the pan-pholids †Agnevichthys gretchinae and †Palaeopholis laevis from the Serravallian of Russia (Nazarkin 2002). Bayesian relaxed molecular clock analyses of Zoarcoidea result in an average posterior crown age estimate of 29.9 million years ago, with the credible interval ranging between 23.0 and 37.2 million years ago (Ghezelayagh et al. 2022). Zoarcoidea is a valid family-group name under the International Code of Zoological Nomenclature (Van der Laan et al. 2014:114).

  • img-z165-2_03.gif

    Centrarchiformes P. Bleeker 1859:xix
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive clade that contains Centrarchus macropterus (Lacépède 1801), Micropterus salmoides (Lacépède 1802), Percalates colonorum (Günther 1863b), and Kuhlia marginata (Valenciennes in Cuvier and Valenciennes 1829a). This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek κέντρoν (k̍εntɹαːn), which can refer to any sharp point such as the tip of a spear, and άρχός (̍α͡ɹko͡Ʊz) meaning anus. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 977.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, fig. S20). Although Centrarchus macropterus is not included in the reference phylogeny, the species resolves with other species of Centrarchidae in phylogenetic analyses of DNA sequences of Sanger-sequenced mitochondrial and nuclear genes (Near, Bolnick, et al. 2004, fig. 1; Near and Kim 2021, fig. 2A). Phylogenetic relationships among the major lineages of Centrarchiformes are presented in Figure 19. The phylogenetic placement of Caesioscorpis theagenes (Blowhole Perch) is on the basis of preliminary analyses of 10 Sanger-sequenced nuclear genes used in other studies of centrarchiform and acanthomorph phylogeny (e.g., Near et al. 2012c, 2013).

  • Phylogenetics. Centrarchiformes as delimited here was first resolved as a monophyletic group in phylogenetic analyses of Sanger-sequenced mtDNA and nuclear genes (Near et al. 2012c, 2013; Betancur-R, Broughton, et al. 2013; W.-J. Chen, Lavoué, et al. 2014; Sanciangco et al. 2016). Molecular phylogenetic analyses consistently resolve three major lineages within Centrarchiformes: (1) Percalates (estuary perches) as the sister lineage of all other Centrarchiformes (Figure 19; Near et al. 2012c; W.-J. Chen, Lavoué, et al. 2014; Lavoué, Nakayama, et al. 2014; Ghezelayagh et al. 2022); (2) Terapontoidei including Girellidae (nibblers), Scorpididae (halfmoons), Parascorpis typus (Jutjaw), Dichistius (galjoen fishes), Microcanthidae (stripeys), Caesioscorpis theagenes (Blowhole Perch), Oplegnathus (knifejaws), Kyphosidae (sea chubs), Kuhlia (flagtails), and Terapontidae (grunters) (Figure 19; Yagishita et al. 2002, 2009; Knudsen and Clements 2016; Sanciangco et al. 2016; Betancur-R et al. 2017; Knudsen et al. 2019; Ghezelayagh et al. 2022); and (3) Centrarchoidei including Enoplosus armatus (Oldwife), Percichthyidae (temperate perches), Centrarchidae (sunfishes, blackbasses, and pygmy sunfishes), Sinipercidae (Chinese perches), Cirrhitidae (hawkfishes), Latridae (trumpeters), Chironemus (kelpfishes), Cheilodactylus (fingerfins), and Aplodactylus (marblefishes) (Figure 19; C. H. Li, Ortí, et al. 2010; Near et al. 2012c, 2013; Sanciangco et al. 2016; Betancur-R et al. 2017; Song et al. 2017; Ghezelayagh et al. 2022).

  • The classification of Centrarchiformes is dynamic and unsettled, reflected in part by a high proportion of families in rank-based classifications that contain a single genus (Regan 1913b; J. B. L. Smith 1935; G. D. Johnson 1984; Gosline 1985; Betancur-R et al. 2017; Fricke et al. 2023). Molecular phylogenies consistently resolve two sets of traditionally delimited centrarchiform families as nonmonophyletic. First, the two species of Percalates were classified as Percichthyidae (G. D. Johnson 1984), but resolve as the sister lineage of all other centrarchiforms and there is no described rank-based taxonomic family to accommodate the classification of Percalates (Figure 19; Near et al. 2012c, 2013; Betancur-R, Broughton, et al. 2013; Sanciangco et al. 2016; Betancur-R et al. 2017; Rabosky et al. 2018; Ghezelayagh et al. 2022). Second, the classification of families within cirrhitoids was dramatically realigned because of molecular phylogenetic analyses. Traditionally, Cheilodactylidae (morwongs) contained three to five genera and approximately 22 species (Greenwood 1995; J. S. Nelson 2006:386). Phylogenetic analyses of mtDNA gene sequences, morphology, and a phylogenomic UCE dataset resolved Cheilodactylidae as polyphyletic, with all but two of the species traditionally classified as cheilodactylids nested within a paraphyletic Latridae (Figure 19; Burridge and Smolenski 2004; Kimura et al. 2018; Ludt et al. 2019). The results of these phylogenetic analyses resulted in a transfer of these species to Latridae from Cheilodactylidae. Phylogenomic analyses of UCE loci differ in resolving Cheilodactylus and Chironemus rather than Cheilodactylus and Aplodactylus as sister lineages (Ludt et al. 2019; Ghezelayagh et al. 2022).

  • The two species of Percilia (southern basses) were traditionally classified in the monogeneric family Perciliidae (J. S. Nelson et al. 2016:433–434). They are classified here as species of Percichthyidae, reflecting the results of several molecular phylogenetic analyses (Near et al. 2013; W.-J. Chen, Lavoué, et al. 2014; Lavoué, Nakayama, et al. 2014; Betancur-R et al. 2017; Ghezelayagh et al. 2022).

  • The relationships of Elassoma (pygmy sunfishes) were a long-standing problem in the systematics and taxonomy of percomorph fishes (Boulenger 1895:34–35; Branson and Moore 1962; G. D. Johnson and Patterson 1993; Jones and Quattro 1999; Near et al. 2012c). The consistent resolution of Elassoma and all other centrarchids as sister lineages in molecular phylogenetic analyses (Near et al. 2012c; W.-J. Chen, Lavoué, et al. 2014; Ghezelayagh et al. 2022) motivated the classification of Elassoma as a lineage of Centrarchidae (Near et al. 2012c:391).

  • Composition. There are currently 304 species of Centrarchiformes (Fricke et al. 2023) that include Enoplosus armatus, Parascorpis typus, and species classified in Aplodactylus, Caesioscorpis, Centrarchidae, Cheilodactylus, Chironemus, Cirrhitidae, Dichistius, Girellidae, Kuhlia, Kyphosidae, Latridae, Microcanthidae, Oplegnathus, Percalates, Percichthyidae, Scorpididae, Sinipercidae, and Terapontidae. Over the past 10 years, there have been 11 new living species of Centrarchiformes described (Fricke et al. 2023), comprising 3.6% of the living species diversity in the clade.

  • Diagnostic apomorphies. There are no known morphological apomorphies for Centrarchiformes.

  • Synonyms. There are no synonyms of Centrarchiformes.

  • Comments. The name Centrarchiformes was applied to a clade containing Girellidae, Oplegnathus, Kuhlia, Kyphosidae, Terapontidae, Percalates, Enoplosus, Percichthyidae, Cheilodactylus, Cirrhitidae, Sinipercidae, and Centrarchidae resolved in phylogenetic analyses of Sanger-sequenced nuclear genes (Near et al. 2013, fig. S1).

  • The earliest Centrarchiformes fossils include a premaxilla attributed to an undetermined species of Oplegnathidae from the Ypresian (56.0–48.1 Ma) of Seymour Island, Antarctica (Cione et al. 1995), and the centrarchid †Plioplarchus whitei from the Priabonian (37.7–33.9 Ma) of North Dakota, USA (Cope 1883; Near and Kim 2021). Bayesian relaxed molecular clock analyses of Centrarchiformes result in an average posterior crown age estimate of 53.3 million years ago, with the credible interval ranging between 35.7 and 78.1 million years ago (Ghezelayagh et al. 2022).

  • img-z167-10_03.gif

    FIGURE 19.

    Phylogenetic relationships of the major living lineages and fossil taxa of Centrarchiformes, Labriformes, and Acropomatiformes. Filled circles identify the common ancestor of clades with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z166-1_03.jpg

    Labriformes P. Bleeker 1862:416
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Labrus mixtus Linnaeus 1758, Labrus bergylta Ascanius 1767, Bodianus rufus (Linnaeus 1758), Parapercis hexophtalma (Cuvier in Cuvier and Valenciennes 1829a), Astroscopus y-graecum (Cuvier in Cuvier and Valenciennes 1829a), and Centrogenys vaigiensis Quoy and Gaimard 1824. This is a minimum-crown-clade definition.

  • Etymology. Derived from the Latin labrum meaning lip. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 978.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, fig. S21). Although not included in the reference phylogeny, Labrus mixtus is nested in Labridae with other species of Labrus in a phylogeny resulting from analysis of Sanger-sequenced mitochondrial and nuclear genes (Aiello et al. 2017, fig. S2). Phylogenetic relationships among major lineages of Labriformes are presented in Figure 19. The placements of fossils in the phylogeny of Labriformes are on the basis of resolutions suggested from morphological inferences for the pan-labrids †Bellwoodilabrus (Bannikov and Carnevale 2010) and †Labrobolcus (Bannikov and Bellwood 2015).

  • Phylogenetics. Phylogenomic analysis of UCE loci and analysis of Sanger-sequenced mitochondrial and nuclear genes resolves Labriformes as a monophyletic group that contains two major lineages: (1) a clade containing Uranoscopidae (stargazers), Ammodytidae (sand lances), Pinguipedidae (sandperches), Leptoscopidae (southern sandfishes), and Cheimarrichthys fosteri (Torrentfish) (Figure 19; Betancur-R et al. 2017; Ghezelayagh et al. 2022); and (2) Labridae (wrasses and parrotfishes) and Centrogenys (false scorpionfishes) (Figure 19; Ghezelayagh et al. 2022; Hughes et al. 2023; Matsunuma and Johnson 2023). Molecular phylogenetic studies are the basis for considerable adjustments to the delimitation of Labridae, specifically the inclusion of species formerly classified as Scaridae (parrotfishes) and Odacidae (cales) (Westneat and Alfaro 2005; Alfaro, Brock, et al. 2009; Cowman et al. 2009; Baliga and Law 2016; Hughes et al. 2023).

  • Morphological studies place Cheimarrichthys as closely related to Pinguipedidae or as the sister lineage of all “trachinioids” including Uranoscopidae, Ammodytidae, Pinguipedidae, and Leptoscopidae (McDowall 1973; Pietsch 1989; Pietsch and Zabetian 1990). The hypothesis that Cheimarrichthys and Pinguipedidae share common ancestry was rejected through the discovery that Cheimarrichthys shares more derived character states with Leptoscopidae than any other “trachinioid” lineage (Imamura and Matsuura 2003). Reflective of the shared common ancestry inferred from morphology and molecular phylogenetic studies (Imamura and Matsuura 2003; Thacker et al. 2015; Ghezelayagh et al. 2022), Cheimarrichthys and Leptoscopidae have a similar geographic distribution: Cheimarrichthys is an anadromous species widely distributed among the rivers of New Zealand, and leptoscopids are distributed along the Pacific and Indian coasts of Australia and New Zealand (McDowall 2000; Last 2001).

  • Labridae and Centrogenys vaigiensis are resolved as sister lineages (Betancur-R et al. 2017; Ghezelayagh et al. 2022; Hughes et al. 2023). The species-rich Labridae and the blenniiform lineages Cichlidae, Embiotocidae, and Pomacentridae were hypothesized to be closely related in a clade based on the morphology of the modified “labroid” pharyngeal jaw apparatus (Liem and Greenwood 1981; Kaufman and Liem 1982; Stiassny and Jensen 1987; Springer and Orrell 2004). Previous molecular phylogenetic analyses using Sanger-sequenced nuclear genes demonstrated that the lineages with the modified “labroid” pharyngeal jaw apparatus do not resolve as closely related, but these early molecular studies resulted in an ambiguous and poorly supported resolution of the species-rich Labridae (Streelman and Karl 1997; W. L. Smith and Wheeler 2004, 2006; Sparks and Smith 2004a; Mabuchi et al. 2007; Wainwright et al. 2012; Near et al. 2013). The resolution of Labridae and Centrogenys vaigiensis as sister lineages is interesting as both lineages have all three components of the modified labroid pharyngeal jaw apparatus, a set of traits that has originated multiple times in Percomorpha (Wainwright et al. 2012).

  • Composition. There are currently 887 living species of Labriformes (Fricke et al. 2023; Matsunuma and Johnson 2023) that include Centrogenys, Cheimarrichthys fosteri, and species classified in Ammodytidae, Labridae, Leptoscopidae, Pinguipedidae, and Uranoscopidae. Fossil lineages of Labriformes include the pan-labrids †Bellwoodilabrus landinii and †Labrobolcus giorgioi from the Ypresian (56.0–48.1 Ma) of Italy (Appendix 2; Bannikov and Carnevale 2010; Bannikov and Bellwood 2015). Over the past 10 years, there have been 77 new living species of Labriformes described (Fricke et al. 2023), comprising 8.7% of the living species diversity in the clade.

  • Diagnostic apomorphies. There are no known morphological apomorphies for Labriformes.

  • Synonyms. There are no synonyms of Labriformes.

  • Comments. Labriformes was applied as a name to a polyphyletic group containing Labridae, Embiotocidae, Cichlidae, and Pomacentridae (Wiley and Johnson 2010). In a recent classification of ray-finned fishes, Labriformes was limited to Labridae (Betancur-R et al. 2017). The delimitation of Labriformes presented here is consistent with relationships inferred in phylogenomic studies (Ghezelayagh et al. 2022; Hughes et al. 2023).

  • The earliest fossil Labriformes are species classified as Labridae or pan-labrids that include †Bellwoodilabrus landinii (Bannikov and Carnevale 2010), †Labrobolcus giorgioi (Bannikov and Bellwood 2015), †Eocoris bloti (Bannikov and Sorbini 1990), †Phyllopharyngodon longipinnis (Bellwood 1990), †Zorzinilabrus furcatus (Bannikov and Bellwood 2017), and †Paralabrus rossiae (Bannikov and Zorzin 2019) from the Ypresian (56.0–48.1 Ma) of Italy. Bayesian relaxed molecular clock analyses of Labriformes result in an average posterior crown age estimate of 76.1 million years ago, with the credible interval ranging between 66.2 and 87.0 million years ago (Ghezelayagh et al. 2022).

  • img-z169-8_03.gif

    Acropomatiformes M. P. Davis, J. S. Sparks, and W. L. Smith 2016:fig. 1
    [C. E. Thacker and T. J. Near], converted clade name

  • Definition. The least inclusive crown clade that contains Acropoma japonicum Günther 1859, Pempheris schomburgkii Müller and Troschel in Schomburgk 1848, Stereolepis gigas Ayres 1859, and Pteropsaron evolans Jordan and Snyder 1902. This is a minimum-crown-clade definition.

  • Etymology. Derived from the ancient Greek ἄκρoς (̍ækςo͡Ʊz) meaning at the end or at the top and ππµα (p̍o͡Ʊmә) meaning lid or cover. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 979.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, fig. S22). Phylogenetic relationships of the major lineages of Acropomatiformes are presented in Figure 19. The placements of Dinolestes lewini, Malakichthyidae, Schuettea, and Synagropidae in the phylogeny are on the basis of analysis of DNA sequences from nine Sanger-sequenced mtDNA and nuclear genes, and 457 UCE loci (W. L. Smith et al. 2022).

  • Phylogenetics. Acropomatiformes is a lineage resolved entirely as a result of molecular phylogenetic analyses (W. L. Smith and Craig 2007; B. Li et al. 2009; Betancur-R, Broughton, et al. 2013; Near et al. 2013, 2015; Thacker et al. 2015; Davis et al. 2016; Sanciangco et al. 2016; Betancur-R et al. 2017; Mirande 2017; Ghedotti et al. 2018; Rabosky et al. 2018; Satoh 2018; Oh et al. 2021; Ghezelayagh et al. 2022; W. L. Smith et al. 2022). Phylogenomic studies delimit three major clades of Acropomatiformes (Figure 19): (1) Ostracoberyx (shellskin alfonsinos), Acropomatidae (lanternbellies), Scombrops (gnomefishes), Symphysanodontidae (slopefishes), Epigonidae (deepwater cardinalfishes), and Howellidae (oceanic basslets); (2) Polyprion (wreckfishes), Lateolabrax (Asian seabasses), Glaucosoma (pearl perches), and Pempheridae (sweepers); and (3) Stereolepis (giant seabasses), Banjos (banjofishes), Pentacerotidae (armorheads), Dinolestes lewini (Long-finned Pike), Malakichthyidae (temperate seabasses), Bathyclupeidae (deepsea herrings), Synagropidae (splitfin seabasses), Champsodon (gapers), Schuettea (moony pomfrets), Creediidae (sandburrowers), and Hemerocoetidae (signalfishes).

  • The lineages comprising Acropomatiformes were previously classified in the defunct Trachinoidei or the historic taxonomic wastebasket Percoidei (G. D. Johnson 1984; Imamura and Odani 2013; J. S. Nelson et al. 2016:431–463). Less inclusive groups were classified with distantly related species in Linnean-ranked taxonomic families: Scombrops was classified in Pomatomidae (Nelson 1994:350–351), Lateolabrax was placed in Percichthyidae (J. S. Nelson 2006:344), Hemilutjanus macrophthalmos was considered a species of Serranidae (J. S. Nelson et al. 2016:446–448), and Schuettea was long classified in Monodactylidae (Regan 1913b; Jordan 1923:205; J. S. Nelson et al. 2016:452–453) despite the recognition of appreciable morphological differences with Monodactylus (Tominaga 1968). The traditional delimitation of Acropomatidae (J. S. Nelson et al. 2016:434) is not monophyletic (W. L. Smith and Craig 2007; Betancur-R, Broughton, et al. 2013; Near et al. 2013, 2015; Thacker et al. 2015; Davis et al. 2016; Sanciangco et al. 2016; Mirande 2017; Ghedotti et al. 2018; Rabosky et al. 2018; Oh et al. 2021; Ghezelayagh et al. 2022; W. L. Smith et al. 2022), necessitating the elevation of Malakichthyidae to include Hemilutjanus macrophthalmos, Malakichthys, and Verilus; Synagropidae to include Caraibops trispinosus, Kaperangus microlepis, Parascombrops, and Synagrops; and limiting Acropomatidae to Acropoma and Doederleinia berycoides (W. L. Smith et al. 2022).

  • Composition. There are currently 306 living species of Acropomatiformes (Fricke et al. 2023) that include Dinolestes lewini and species classified in Acropomatidae, Banjos, Bathyclupeidae, Champsodon, Creediidae, Epigonidae, Glaucosoma, Hemerocoetidae, Howellidae, Lateolabrax, Malakichthyidae, Ostracoberyx, Pempheridae, Pentacerotidae, Polyprion, Scombrops, Schuettea, Synagropidae, and Symphysanodontidae (W. L. Smith et al. 2022). Over the past 10 years, there have been 84 new living species of Acropomatiformes described (Fricke et al. 2023), comprising 27.5% of the living species diversity in the clade. Most of these new taxa are species of Pempheris (e.g., J. E. Randall and Victor 2015).

  • Diagnostic apomorphies. There are no known morphological apomorphies for Acropomatiformes.

  • Synonyms. Pempheriformes is an ambiguous (Sanciangco et al. 2016, fig. 5; Betancur-R et al. 2017:29) and partial synonym (Betancur-R, Broughton, et al. 2013, app. 2) of Acropomatiformes. Trachiniformes (J. S. Nelson et al. 2016:421) and Clade R (B. Li et al. 2009:358) are partial synonyms of Acropomatiformes.

  • Comments. Smith et al. (2022:9) provide a discussion and justification for the use of the group name Acropomatiformes for this clade.

  • Relative to other lineages of Percomorpha, Acropomatiformes includes a large proportion of species that exhibit bioluminescence and occupy deepwater oceanic habitats, traits that seem to have multiple origins in the clade (Davis et al. 2016; Ghedotti et al. 2018; W. L. Smith et al. 2022).

  • The earliest fossil Acropomatiformes are otoliths identified as Pempheridae or †Pempheris huddlestoni from the Maastrichtian (72.2–66.0 Ma) in the Cretaceous of Maryland, USA (Huddleston and Savoie 1983; Stringer and Schwarzhans 2021), †Acropoma sp. from the Selandian (61.7–59.2 Ma) of Denmark (Schwarzhans 2003), and †Epigonidarum tyassminensis from the Selandian (61.7–59.2 Ma) to Thanetian (59.2–56.0 Ma) of Ukraine (Schwarzhans and Bratishko 2011). Bayesian relaxed molecular clock analyses of Acropomatiformes result in an average posterior crown age estimate of 46.5 million years ago, with the credible interval ranging between 32.6 and 61.0 million years ago (Ghezelayagh et al. 2022).

  • img-z171-2_03.gif

    Acanthuriformes D. S. Jordan 1923:207
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The most inclusive crown clade that contains Acanthurus lineatus (Linnaeus 1758), but not Perca fluviatilis (Linnaeus 1758), Centrarchus macropterus (Lacépède 1801), Labrus mixtus Linnaeus 1758, and Acropoma japonicum Günther 1859. This is a maximum-crown-clade definition.

  • Etymology. From the ancient Greek ἄκανθα (ɐk̍'ænθә) meaning thorn or spine. The suffix is from the Latin forma meaning form, figure, or appearance.

  • Registration number. 980.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, figs. S22–S25). Although Centrarchus macropterus is not included in the reference phylogeny, the species resolves with other species of Centrarchidae in phylogenetic analyses of DNA sequences of Sanger-sequenced mitochondrial and nuclear genes (Near, Bolnick, et al. 2004, fig. 1; Near and Kim 2021, fig. 2A), and Labrus mixtus is nested in Labridae with other species of Labrus in a phylogeny resulting from analysis of Sanger-sequenced mitochondrial and nuclear genes (Aiello et al. 2017, fig. S2). The phylogenetic relationships of the major living lineages and fossil taxa of Acanthuriformes are presented in Figure 20. The placements of †Eoscatophagus, †Oligoscatophagus, †Ruffoichthys, †Eosiganus, †Siganopygaeus, †Protosiganus, and †Caprosimilis in the phylogeny are on the basis of analyses of morphological characters (Tyler and Bannikov 1997; Tyler and Sorbini 1999; Bannikov and Tyler 2001; Bieńkowska-Wasiluk and Bonde 2015; Siqueira et al. 2019).

  • Phylogenetics. Acanthuriformes as delimited here is resolved as a monophyletic group in phylogenomic analyses of UCE loci (Ghezelayagh et al. 2022). Other phylogenomic analyses and analyses of DNA sequences from Sanger-sequenced mtDNA and nuclear genes place Gerreidae (mojarras) outside of Acanthuriformes in variable resolutions that include: as the sister lineage of an inclusive clade that contains all sampled Acanthuriformes and Labriformes (W.-J. Chen et al. 2003), the sister lineage of Labridae (Wainwright et al. 2012; Mu et al. 2022); a deeply branching lineage among Labriformes, Centrarchiformes, Perciformes, and Acanthuriformes (Near et al. 2013; Betancur-R et al. 2017); the sister lineage of a clade containing Centrogenys and Labridae (Smith et al. 2016); or as the sister lineage of Labriformes with (Hughes et al. 2018, 2023) and without Labridae (Rabosky et al. 2018).

  • The discovery that Acanthuriformes is a monophyletic group including more than 2,365 species classified into 59 taxonomic families in rank-based classifications is an important advance in the systematics of percomorph fishes, adding clarity to the relationships of several lineages that long evaded phylogenetic resolution. Moronidae (temperate basses) were traditionally classified in the catchall “percoid” taxonomic wastebasket (e.g., G. D. Johnson 1984); however, molecular phylogenies resolved Moronidae as a deeply nested lineage within Acanthuriformes (Wainwright et al. 2012; Near et al. 2013; Smith et al. 2016; Hughes et al. 2018) or in a clade with Sillaginidae (whitings) as the sister lineage of all other Acanthuriformes (Figure 20; Betancur-R et al. 2017; Ghezelayagh et al. 2022).

  • Lineages of Acanthuriformes were among those that comprised Squamipinnes, which are characterized by scales on the bases of the second dorsal and anal fins and are among the earliest proposed inclusive groups of percomorphs (Cuvier 1816; Matsubara 1955; Mok and Shen 1983). Hypotheses for the composition of Squamipinnes in the 20th century ranged from inclusion of lineages of Carangiformes (Toxotidae), Acropomatiformes (Pentacerotidae), and Centrarchiformes (Scorpididae) (Mok and Shen 1983) to a delimitation that included only lineages of Acanthuriformes: Acanthuroidei, Caproidae (boarfishes), Chaetodontidae (butterflyfishes), Drepane (sicklefishes), Ephippidae (spadefishes), Pomacanthidae (angelfishes), and Scatophagidae (scats) (Blum 1988; Tyler et al. 1989; Rosen and Patterson 1990). Morphological studies identified synapomorphies supporting the monophyly of a lineage containing Acanthuroidei, Chaetodontidae, Ephippidae, Pomacanthidae, Scatophagidae, and Siganus (rabbitfishes) (Tyler et al. 1989); however, molecular phylogenetic analyses of Percomorpha consistently fail to resolve lineages traditionally classified in Squamipinnes as a monophyletic group (Holcroft and Wiley 2008; Near et al. 2013; Smith et al. 2016; Betancur-R et al. 2017; Rabosky et al. 2018; Ghezelayagh et al. 2022). A phylogeny inferred from morphology limited Acanthuriformes to Acanthuroidei, Caproidae, Siganus, Scatophagidae, Leiognathidae (ponyfishes), Ephippidae, Chaetodontidae, Drepane, Pomacanthidae, and Lobotidae (tripletails and barbled grunters: Datnioides, Hapalogenys, and Lobotes) (A. C. Gill and Leis 2019). A delimitation of Acanthuriformes that excludes Lophioidei and Tetraodontoidei is not resolved in molecular phylogenetic analyses (e.g., Miya et al. 2005; Holcroft and Wiley 2008; Near et al. 2013; Smith et al. 2016; Betancur-R et al. 2017; Ghezelayagh et al. 2022; Mu et al. 2022).

  • Chaetodontidae and Pomacanthidae were long considered as closely related, which was reflected in classifications that treated pomacanthidsasasubfamilyofChaetodontidae(Berg 1940:245–246; Greenwood et al. 1966; J. S. Nelson 1976). Other investigators noted differences between the two lineages, leading to their classification as two separate Linnaean-ranked taxonomic families (J. B. L. Smith 1955b; Burgess 1974; J. S. Nelson et al. 2016:454–456). Morphological phylogenetic analyses resolve Chaetodontidae and Pomacanthidae as sister lineages (Mok and Shen 1983; Blum 1988; Tyler et al. 1989); however, the two lineages do not resolve as a monophyletic group in molecular phylogenies (Bellwood et al. 2004; Fessler and Westneat 2007; Wainwright et al. 2012; Near et al. 2013; Smith et al. 2016; Betancur-R et al. 2017; Rabosky et al. 2018; Ghezelayagh et al. 2022). Molecular phylogenetic analyses resolve Chaetodontidae and Leiognathidae as sister lineages (Wainwright et al. 2012; Near et al. 2013; Betancur-R et al. 2017; Ghezelayagh et al. 2022). Pomacanthidae is resolved as a deeply nested lineage among Acanthuriformes with poor node support (Betancur-R, Broughton, et al. 2013; Near et al. 2013; Rabosky et al. 2018), as the sister lineage of the clade containing Chaetodontidae and Leiognathidae (Smith et al. 2016; Ghezelayagh et al. 2022), or as the sister lineage of Scatophagidae (Bellwood et al. 2004; Fessler and Westneat 2007).

  • Prephylogenetic systematic studies identified two “percoid” lineages (G. D. Johnson 1980), Lutjanidae (snappers) and Caesionidae (fusiliers), and the sparoids that includes Sparidae (porgies), Nemipteridae (threadfin breams), and Lethrinidae (emperors) that were each resolved as acanthuriform clades in molecular phylogenies (Near et al. 2013; Sanciangco et al. 2016; Betancur-R et al. 2017; Ghezelayagh et al. 2022). Callanthiidae (splendid perches) is resolved as the sister lineage of the sparoids (Figure 20; Ghezelayagh et al. 2022). Phylogenies inferred from Sanger-sequenced mtDNA and nuclear genes resolve Haemulidae (grunts) as the sister lineage of a clade containing Lutjanidae and Caesionidae with low node support (Wainwright et al. 2012; Near et al. 2013; Betancur-R et al. 2017; Rabosky et al. 2018); however, this clade is not resolved in phylogenomic analyses (Hughes et al. 2018; Ghezelayagh et al. 2022). Phylogenomic analysis of UCE loci resolves Haemulidae and Dinopercidae (cavebasses) as sister lineages (Figure 20; Ghezelayagh et al. 2022). In contrast, a survey of morphological features associated with the skull and caudal skeleton identified the acropomatiform Glaucosoma and the perciform Serranidae (sensu lato) as the possible relatives of Dinopercidae (Heemstra and Hecht 1986). Molecular phylogenies support the monophyly of Dinopercidae (W. L. Smith and Craig 2007).

  • One of the most surprising findings from molecular studies of teleost phylogeny is the resolution of Lophioidei (anglerfishes, formerly Lophiiformes) and Tetraodontoidei (puffers and molas, formerly Tetraodontiformes) as sister lineages within Acanthuriformes (Miya et al. 2003, 2005; Yamanoue et al. 2007; Holcroft and Wiley 2008; Yagishita et al. 2009; Near, Eytan, et al. 2012; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Smith et al. 2016; Alfaro et al. 2018; Hughes et al. 2018; Ghezelayagh et al. 2022; Mu et al. 2022). In morphology-based classifications, the lophioids were placed in Paracanthopterygii (Patterson and Rosen 1989; J. S. Nelson 2006:250–260), phylogenetically distantly related to other lineages of Percomorpha. The migration of Lophioidei as paracanthopterygians into a derived clade of percomorphs is among the most significant changes in 21st-century vertebrate phylogenetics (Dornburg and Near 2021), akin to moving a morphologically placed lineage from within marsupials to the sister lineage of primates. While the discovery that Lophioidei and Tetraodontoidei are closely related is on the basis of phylogenetic analysis of molecular data, subsequent investigation of morphology identifies shared traits in the lateral line system (Nakae and Sasaki 2010), several soft tissue characters that are likely synapomorphies of a lophioid-tetraodontoid clade (Chanet et al. 2013), and unique morphology and pigmentation in larvae shared by lophioids and tetraodontoids (Baldwin 2013). Other traits thought to be unique to both lophioids and tetraodontoids may turn out to be morphological synapomorphies for this clade, including: absence of anal-fin spines (Pietsch 1981, 1984; Tyler and Sorbini 1996), absence of ribs (Pietsch 1981; Tyler and Sorbini 1996), reduced number of caudal-fin rays (Pietsch 1981, 1984; Tyler and Sorbini 1996), reduced number of vertebrae (Pietsch 1984; Tyler and Sorbini 1996), and a restricted opercular opening (Pietsch 1981; Tyler and Sorbini 1996).

  • Molecular phylogenetic analyses resolve an inclusive clade within Acanthuriformes that contains Siganus, Scatophagidae, Priacanthidae (bigeyes), Cepolidae (bandfishes), Caproidae, Lophioidei, and Tetraodontoidei (Near et al. 2013; Smith et al. 2016; Betancur-R et al. 2017; Ghezelayagh et al. 2022). A close relationship between Caproidae and Tetraodontoidei was proposed on the basis of analysis of morphology (Zehren 1987).

  • Composition. There are currently 2,376 living species of Acanthuriformes (Allen 1985; Carpenter 1988; Carpenter and Allen 1989; B. C. Russell 1990; McKay 1992, 1997; Fricke et al. 2023) classified in Acanthuroidei, Callanthiidae, Caproidae, Cepolidae, Chaetodontidae, Dinopercidae, Drepane, Emmelichthyidae (rovers), Ephippidae, Gerreidae, Haemulidae, Leiognathidae, Lethrinidae, Lobotidae, Lophioidei, Lutjanidae, Malacanthidae (tilefishes), Monodactylus (moonfishes), Moronidae, Nemipteridae, Pomacanthidae, Priacanthidae, Scatophagidae, Sciaenidae (drums), Siganus, Sillaginidae, Sparidae, and Tetraodontoidei. Fossil lineages of Acanthuriformes include the pan-scatophagids †Eoscatophagus and †Oligoscatophagus (Tyler and Sorbini 1999); the pan-siganids †Ruffoichthys, †Eosiganus, †Siganopygaeus, and †Protosiganus (Tyler and Bannikov 1997); the pan-caproid †Caprosimilis (Bieńkowska-Wasiluk and Bonde 2015); and several taxa in Lophioidei and Tetraodontoidei. Details of the ages and locations of the fossil taxa are presented in Appendix 1. Over the past 10 years, 141 new living species of Acanthuriformes have been described (Fricke et al. 2023), comprising 5.9% of the living species diversity in the clade.

  • Diagnostic apomorphies. There are no known synapomorphies for Acanthuriformes. The presence of posterolateral tooth replacement was hypothesized as a synapomorphy for a delimitation of Acanthuriformes limited to Acanthuroidei, Caproidae, Siganus, Scatophagidae, Leiognathidae, Ephippidae, Chaetodontidae, Drepane, Pomacanthidae, and Lobotidae (A. C. Gill and Leis 2019).

  • Synonyms. There are no synonyms of Acanthuriformes.

  • Comments. The first use of Acanthuriformes in post-Hennigian systematics was as the name for a group containing Siganus, Luvarus, Zanclus, and Acanthuridae (Wiley et al. 2000), which is a synonym of Acanthuroidei (Tyler et al. 1989). The delimitation of Acanthuriformes presented here follows on Davis et al. (2016), but we include Gerreidae.

  • The monophyly of the major lineages of Percomorpha delimited in this classification was discovered primarily as a result of molecular phylogenetic analyses (Dornburg and Near 2021). The earliest molecular phylogenetic studies of Percomorpha revealed the challenge of resolving relationships among Perciformes, Centrarchiformes, Acropomatiformes, and what Davis et al. (2016) first delimited as Acanthuriformes (Miya et al. 2005; Dettaï and Lecointre 2008; B. Li et al. 2009; Chanet et al. 2013; Near et al. 2013). The limits of the phylogenetic resolution offered in the first wave of molecular studies was particularly acute for the lineages classified as Acanthuriformes; however, the application of phylogenomic methods provide an important incremental step toward a strongly supported hypothesis and an inclusive classification. As the result of phylogenomic analyses, 10 of 11 lineages classified by Betancur-R et al. (2017) as incertae sedis in Eupercaria find resolution in Acanthuriformes (Dornburg and Near 2021; Ghezelayagh et al. 2022).

  • The earliest fossil Acanthuriformes include several lineages of pan-scatophagids and pan-siganids from the Ypresian (56.0–48.1 Ma). Details on the phylogenetic placement and location of these fossil species are given in Appendix 1. Bayesian relaxed molecular clock analyses of Acanthuriformes result in an average posterior crown age estimate of 78.5 million years ago, with the credible interval ranging between 72.0 and 86.6 million years ago (Ghezelayagh et al. 2022).

  • img-z175-4_03.gifimg-AIsjR_03.gif

    FIGURE 20.

    Phylogenetic relationships of the major living lineages and fossil taxa of Acanthuriformes and Acanthuroidei. Filled circles identify the common ancestor of clades with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z172-1_03.jpg

    Acanthuroidei P. Bleeker 1859:xxii
    [T. J. Near and C. E. Thacker], converted clade name

  • Definition. The least inclusive crown clade containing Paracanthurus hepatus (Linnaeus 1766), Acanthurus lineatus (Linnaeus 1758), Zanclus cornutus (Linnaeus 1758), and Luvarus imperialis Rafinesque 1810a. This is a minimum-crown-clade definition.

  • Etymology. From the ancient Greek ἄκανθα (ɐk̍'ænθә) meaning thorn or spine.

  • Registration number. 981.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element loci (Ghezelayagh et al. 2022, fig. S23). Phylogenetic relationships of the major living lineages and fossil taxa of Acanthuroidei are presented in Figure 20. The placements of pan-luvarids †Avitoluvarus and †Kushlukia, the pan-acanthurids †Padovathurus and †Gazolaichthys, and the pan-zanclid †Massalongius in the phylogeny are on the basis of inferences from morphological characters (Tyler 2005a, 2005b; Tyler and Bannikov 2005; Siqueira et al. 2019).

  • Phylogenetics. Acanthuroidei was traditionally delimited as including Acanthuridae (surgeonfishes), Zanclus cornutus (Moorish Idol), and Siganus (rabbitfishes) (Greenwood et al. 1966; Gosline 1968, 1971:158; Mok and Shen 1983). Morphological phylogenetic studies resulted in an expansion of Acanthuroidei to include Luvarus imperialis (Louvar), Ephippidae (spadefishes), and Scatophagidae (scats) (Tyler et al. 1989; Winterbottom 1993a; Winterbottom and McLennan 1993). Analysis of larval morphology grouped Zanclus and Acanthuridae as sister lineages (G. D. Johnson and Washington 1987), a relationship resolved in both morphological and molecular phylogenetic analyses (Tyler et al. 1989; Winterbottom 1993a; K. L. Tang et al. 1999; Holcroft and Wiley 2008; Near et al. 2013; Betancur-R et al. 2017; A. C. Gill and Leis 2019; Ghezelayagh et al. 2022). Molecular phylogenetic analyses are consistent with a delimitation of Acanthuroidei that includes Luvarus, Zanclus, and Acanthuridae; Siganus, Ephippidae, and Scatophagidae are more closely related to other lineages of Acanthuriformes (Figure 20; Holcroft and Wiley 2008; Near et al. 2013; Z. Liu et al. 2016; J. S. Nelson et al. 2016:499–500; Betancur-R et al. 2017; Ghezelayagh et al. 2022).

  • Composition. There are currently 87 living species of Acanthuroidei (Fricke et al. 2023) that include Luvarus imperialis, Zanclus cornutus, and species classified in Acanthuridae. Fossil taxa of Acanthuroidei include the pan-acanthurids †Padovathurus and †Gazolaichthys (Tyler 2005a, 2005b), and the pan-zanclid †Massalongius (Tyler and Bannikov 2005). Details of the phylogenetic placement and location of these fossil species are given in Appendix 1. Over the past 10 years, there has been one new living species of Acanthuroidei described (Fricke et al. 2023), comprising 1.1% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies of Acanthuroidei include (1) presence of 9 precaudal vertebrae and 13 caudal vertebrae (Tyler et al. 1989; Winterbottom 1993a; Tyler and Sorbini 1999), (2) first dorsal pterygiophore fully inserts in the space of the first interneural and its tip extends into the dorsal area of the foramen magnum (Tyler et al. 1989; Winterbottom 1993a; Tyler and Sorbini 1999; Tyler and Bannikov 2005), (3) infraorbital series changes direction anteriorly below the lateral ethmoid, continuing along the side of the snout (Tyler et al. 1989; Winterbottom 1993a; Tyler and Bannikov 2005), (4) palatine is forward of the lateral ethmoid and there is no

  • articulation between the two bones (Tyler et al. 1989; Winterbottom 1993a; Tyler and Bannikov 2005), (5) absence of the spinna occipitalis, specifically the epiotics contact along the posterior midline of the neurocranium, separating the supraoccipital from the exoccipitals and foramen magnum (Tyler et al. 1989; Winterbottom 1993a; Tyler and Bannikov 2005), (6) presence of small spicules laterally along most or all of the length of the soft rays of the dorsal, anal, caudal, pectoral, and pelvic fins (Tyler et al. 1989; Winterbottom 1993a; Tyler and Bannikov 2005), (7) scales of adults are circular to ovoid plates with upright spicules projecting from their surface (Tyler et al. 1989; Winterbottom 1993a; Tyler and Bannikov 2005), (8) presence of a single postcleithrum in adults (Tyler et al. 1989; Winterbottom 1993a; Tyler and Bannikov 2005), (9) extremely compressed and kite-shaped body (Tyler et al. 1989; Winterbottom 1993a), (10) dome-shaped midbrain that is deeper than it is long, housed in an elongate cranial cavity (Tyler et al. 1989; Winterbottom 1993a), (11) early forming scales bear lamina that project upright from basal plane (Tyler et al. 1989; Winterbottom 1993a), (12) presence of spines on the ascending process of the premaxilla (Tyler et al. 1989; Winterbottom 1993a), (13) lateral surface of lachrymal with two or three serrate ridges (Tyler et al. 1989; Winterbottom 1993a), (14) dentary with two serrate longitudinal ridges (Tyler et al. 1989; Winterbottom 1993a), (15) presence of a locking mechanism for the elongate second or third dorsal spine (Tyler et al. 1989; Winterbottom 1993a), and (16) absence of insertion of abductor superficialis pelvicus on the pelvic spine (Winterbottom 1993a).

  • Synonyms. Acanthuriformes (Betancur-R et al. 2017:27) is an ambiguous synonym of Acanthuroidei.

  • Comments. Acanthuroidei is used as a group as delimited here in recent studies (Dornburg and Near 2021; Ghezelayagh et al. 2022).

  • The earliest fossil taxa of Acanthuroidei include several lineages from the Ypresian (56.0–48.1 Ma) that include the pan-luvarids †Avitoluvarus and †Kushlukia (Bannikov and Tyler 1995), the pan-acanthurids †Padovathurus and †Gazolaichthys (Tyler 2005a, 2005b), and the pan-zanclid †Massalongius (Tyler and Bannikov 2005). Bayesian relaxed molecular clock analyses of Acanthuroidei result in an average posterior crown age estimate of 59.1 million years ago, with the credible interval ranging between 56.3 and 63.0 million years ago (Ghezelayagh et al. 2022).

  • img-z177-2_03.gif

    Lophioidei P. Bleeker 1859:xvi

  • Definition. The least inclusive crown clade that contains Lophius piscatorius Linnaeus 1758, Lophius gastrophysus Miranda Ribeiro 1915, Ogcocephalus radiatus (Mitchill 1818b), and Cryptopsaras couesii T. N. Gill 1883. This is a minimum-crown-clade definition, but the clade is not defined using the PhyloCode.

  • Etymology. From the ancient Greek λόϕoς (l̍o͡Ʊfo͡Ʊz) meaning the back of the neck or the crest of a helmet.

  • Reference phylogeny. A phylogeny inferred from sequences of 989 ultraconserved element (UCE) loci (Ghezelayagh et al. 2022, fig. S25). Although Lophius piscatorius is not in the reference phylogeny, the species resolves with other species of Lophius in phylogenetic analyses of morphology and mtDNA (Leslie and Grant 1994, fig. 4; Landi et al. 2014, fig. S1). Phylogenetic relationships of the major living lineages and fossil taxa of Lophioidei are presented in Figure 21. The placements of the fossil taxa †Sharfia and †Tarkus in the phylogeny are on the basis of inferences from morphological characters (Carnevale and Pietsch 2011, 2012).

  • Phylogenetics. The delimitation of Lophioidei presented here is similar or identical to several pre-Hennigian classifications (Jordan and Sindo 1902; Regan 1912c; Jordan 1923:242–243; Berg 1940:498–500; Greenwood et al. 1966; McAllister 1968; Gosline 1971:173–174). By the early 20th century, species of Lophioidei were classified into three main lineages (Regan 1912c): (1) Lophiidae (goosefishes); (2) the antennarioids that included Antennariidae (frogfishes), Tetrabrachiidae (four-armed frogfishes), Brachionichthyidae (handfishes), Chaunacidae (coffinfishes), Ogcocephalidae (batfishes), and Lophichthys boschmai (Boschma's Frogfish) (Boeseman 1964; Pietsch 1981, 1984); and (3) the ceratioids that included Centrophryne spinulosa (Horned Lanternfish), Neoceratias spinifer (Spiny Seadevil), Caulophrynidae (fanfins), Ceratiidae (warty seadevils), Diceratiidae (double anglers), Gigantactinidae (whipnose anglers), Himantolophus (footballfishes), Linophrynidae (leftvents), Melanocetus (black seadevils), Oneirodidae (dreamers), and Thaumatichthyidae (wolftrap anglers).

  • The first phylogenetic analyses of Lophioidei utilized morphological characters to test the monophyly of Regan's (1912c) delimitation of the antennarioids (Pietsch 1981, 1984). In the morphological phylogeny Lophiidae is resolved as the sister lineage of all other Lophioidei, which consists of two major clades: a modified antennarioid group that includes Antennariidae, Tetrabrachiidae, Lophichthys, and Brachionichthyidae, and a lineage comprising Chaunacidae that is the sister lineage of a clade containing Ogcocephalidae and the ceratioids (Pietsch 1984). Subsequent morphological phylogenetic analyses were aimed at resolving relationships among ceratioids and resulted in differing topologies and degrees of resolution resulting from the separate phylogenetic analyses of characters scored from metamorphosed females, metamorphosed males, and larvae (Pietsch and Orr 2007; Pietsch 2009, fig. 203), characters scored only from metamorphosed females (Pietsch 2009, fig. 202), and characters scored from metamorphosed females with the exclusion of characters that show reductive or simplified states (Miya et al. 2010). Congruence across these morphological phylogenetic analyses includes support for the monophyly of the ceratioids, the resolution of Centrophryne and Ceratiidae as a clade that is the sister lineage to all other ceratioids, and the resolution of Neoceratias and Gigantactinidae as sister lineages (Pietsch and Orr 2007; Pietsch 2009; Miya et al. 2010).

  • Molecular phylogenetic studies of Lophioidei include analyses of datasets of mtDNA (Shedlock et al. 2004; Miya et al. 2010; Poulsen 2019), nuclear genes (Near et al. 2013; Arnold 2014), combinations of mtDNA and nuclear genes (Lundsten et al. 2012; Arnold 2014; Derouen et al. 2015; Betancur-R et al. 2017; Rabosky et al. 2018), and phylogenomic datasets of UCE loci (Ghezelayagh et al. 2022; Hart et al. 2022). Consistent results across molecular studies include the placement of Lophiidae as the sister lineage of all other Lophioidei (Figure 21; Miya et al. 2010; Near et al. 2013; Arnold 2014; Derouen et al. 2015; Betancur-R et al. 2017; Rabosky et al. 2018; Poulsen 2019; Ghezelayagh et al. 2022; Hart et al. 2022), resolution of Chaunacidae and the ceratioids as sister lineages (Figure 21; Shedlock et al. 2004; Miya et al. 2010; Lundsten et al. 2012; Near et al. 2013; Arnold 2014; Derouen et al. 2015; Betancur-R et al. 2017; Rabosky et al. 2018; Ghezelayagh et al. 2022; Hart et al. 2022), resolution of a clade containing Ogcocephalidae, Antennariidae, Brachionichthyidae, and Tetrabrachiidae (Lundsten et al. 2012; Arnold 2014; Ghezelayagh et al. 2022; Hart et al. 2022), Caulophrynidae placed as the sister lineage to all other ceratioids (Figure 21; Miya et al. 2010; Ghezelayagh et al. 2022), and the resolution of Oneirodidae, Himantolophus, Diceratiidae, and Melanocetus as a monophyletic group within the ceratioids (Figure 21; Shedlock et al. 2004; Miya et al. 2010; Lundsten et al. 2012; Arnold 2014; Betancur-R et al. 2017; Rabosky et al. 2018; Ghezelayagh et al. 2022; Hart et al. 2022).

  • Within the ceratioids, phylogenomic analysis of UCE loci resolves Neoceratias spinifer, Linophrynidae, and Ceratiidae as a monophyletic group (Figure 21; Ghezelayagh et al. 2022). These three lineages all exhibit male obligate sexual parasitism and dramatically altered immunity through loss of the capacity for somatic diversification of antigen receptor genes (Regan 1925; Pietsch 2005; Swann et al. 2020). All previous morphological and molecular phylogenetic analyses of ceratioids result in nonmonophyly of the lineages exhibiting obligate male sexual parasitism, implying multiple origins of this reproductive mode (Pietsch and Orr 2007; Pietsch 2009; Miya et al. 2010; Lundsten et al. 2012; Arnold 2014; Swann et al. 2020; Hart et al. 2022). The phylogeny of Lophioidei shown in Figure 21 implies a single evolutionary origin of this unique trait.

  • Two molecular phylogenetic analyses with dense taxon sampling resolve Tetrabrachiidae and Brachionichthyidae nested within a paraphyletic Antennariidae (Arnold 2014; Hart et al. 2022). One proposed solution is to classify the lineages that comprise Tetrabrachiidae and Brachionichthyidae in Antennariidae (Arnold 2014:70–71). An alternative response to the paraphyly of Antennariidae is the description of three new Linnaean-ranked taxonomic families: Histiophrynidae, Rhycheridae, and Tathicarpidae (Hart et al. 2022).

  • Composition. There are currently 408 living species of Lophioidei (Fricke et al. 2023) that include Centrophryne spinulosa, Lophichthys boschmai, Neoceratias spinifer, and species classified in Antennariidae, Caulophrynidae, Ceratiidae, Diceratiidae, Gigantactinidae, Himantolophus, Linophrynidae, Melanocetus, Oneirodidae, and Thaumatichthyidae. Fossil lophioid taxa include the pan-lophiid †Sharfia (Pietsch and Carnevale 2011) and the pan-ogcocephalid †Tarkus (Carnevale and Pietsch 2011). Details of the phylogenetic placement and location of these fossil species are given in Appendix 1. Over the past 10 years, 43 new living species of Lophioidei have been described (Fricke et al. 2023), comprising 10.5% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies for Lophioidei include (1) eggs spawned in a double scroll-shaped sheath (Rasquin 1958; Pietsch 1981, 1984, 2009:177; Pietsch and Grobecker 1987:269; Carnevale and Pietsch 2009, 2011; Wiley and Johnson 2010; Pietsch and Arnold 2020:371), (2) a single hypural plate formed by fusion of the second ural centrum with first preural centra that emanates from a single half centrum (Rosen and Patterson 1969; Pietsch 1981, 1984, 2009:177; Pietsch and Grobecker 1987:269; Carnevale and Pietsch 2009, 2011; Wiley and Johnson 2010; Pietsch and Arnold 2020:371), (3) spinous dorsal fin modified as a luring apparatus (Pietsch 1981, 1984, 2009:177; Pietsch and Grobecker 1987:268; Carnevale and Pietsch 2009, 2011; Wiley and Johnson 2010; Pietsch and Arnold 2020:369), (4) epiotics separated from parietals and meet on the midline posterior of the supraoccipital (Pietsch 1981, 1984, 2009:177; Pietsch and Grobecker 1987:268; Carnevale and Pietsch 2009, 2011; Wiley and Johnson 2010; Pietsch and Arnold 2020:369), (5) gill openings restricted to a small and elongate tubelike opening positioned near the base of the pectoral fin (Pietsch 1981, 1984, 2009:177; Pietsch and Grobecker 1987:268–269; Carnevale and Pietsch 2009, 2011; Wiley and Johnson 2010; Pietsch and Arnold 2020:371), (6) pectoral radials elongate and narrow, ventralmost radial expanded distally (Pietsch 1981, 1984, 2009:177; Pietsch and Grobecker 1987:269; Carnevale and Pietsch 2009, 2011; Wiley and Johnson 2010; Pietsch and Arnold 2020:371), (7) pterygiophores of the spinous dorsal fin develop from a single condensation of tissue that later divides into separate pterygiophores (Everly 2002), (8) the first pterygiophore supports both the first and second dorsal-fin spines (Everly 2002), and (9) urohyal absent, rectus communis muscle originating from the dorsal hypohyal (Datovo et al. 2014; Pietsch and Arnold 2020:371).

  • Synonyms. Pediculati (Günther 1861:178–205; T. N. Gill 1872:2; Boulenger 1904a:188–189, 1904b:717–720; Goodrich 1909:461–462; Jordan 1923:242–243; Regan 1929:326–327) and Lophiiformes (Regan 1926:3; Berg 1940:498–500; Greenwood et al. 1966:397; McAllister 1968:159–163; J. S. Nelson et al. 2016:508–518; Betancur-R et al. 2017:28) are ambiguous synonyms of Lophioidei.

  • Comments. More than 27% of the recognized taxonomic families in Linnaean-based classifications of Percomorpha contain only a single genus or a single species (J. S. Nelson et al. 2016; Fricke et al. 2023). The description of most of these monotypic and monogeneric lineages dates to a time before the introduction of phylogenetic systematics or the application of molecular data to the resolution of relationships among fishes (G. D. Johnson 1984, 1993). The description of the taxonomic families Rhycheridae and Tathicarpidae in 2022 to classify three species does not contribute to a classification of Lophioidei that reflects phylogeny but is anachronistic and only adds redundant group names. We follow the proposal presented by Arnold (2014) to classify the lineages traditionally placed in Brachionichthyidae (14 species) and Tetrabrachiidae (2 species) in Antennariidae, and treat Histiophrynidae (17 species), Rhycheridae (2 species), and Tathicarpidae (1 species) (Hart et al. 2022) as partial synonyms of Antennariidae (68 species).

  • The oldest lophioid fossils date to the Ypresian (56.0–48.1 Ma) of Italy and include the pan-lophiid †Sharfia (Pietsch and Carnevale 2011), the pan-ogcocephalid †Tarkus (Carnevale and Pietsch 2011), and the antennariids †Eophryne, †Histionotophorus, †Orrichthys, and †Neilpeartia (Carnevale and Pietsch 2009, 2010; Carnevale et al. 2020). Bayesian relaxed molecular clock analyses of Lophioidei result in an average posterior crown age estimate of 59.4 million years ago, with the credible interval ranging between 55.3 and 63.6 million years ago (Ghezelayagh et al. 2022).

  • img-z180-6_03.gif

    FIGURE 21.

    Phylogenetic relationships of the major living lineages and fossil taxa of Lophioidei and Tetraodontoidei. Filled circles identify the common ancestor of clades with formal names defined in the clade accounts. Open circles highlight clades with informal group names. Fossil lineages are indicated with a dagger (†). Details of the fossil taxa are presented in Appendix 1.

    img-z178-1_03.jpg

    Tetraodontoidei P. Bleeker 1866:19

  • Definition. The least inclusive crown clade that contains Tetraodon lineatus Linnaeus 1758, Mola mola (Linnaeus 1758), Takifugu rubripes (Temminck and Schlegel 1850), Ostracion cubicus Linnaeus 1758, and Balistes vetula Linnaeus 1758. This is a minimum-crown-clade definition, but the clade is not defined using the PhyloCode.

  • Etymology. From the ancient Greek τετρς(t̍εtɹә) meaning four in compound words and όδoύϛ (h̍o͡Ʊduːz) meaning tooth.

  • Reference phylogeny. A phylogeny inferred from DNA sequences of 1,103 exons (Troyer et al. 2022, fig. S2). Although Tetraodon lineatus is not included in the reference phylogeny, the species resolves with other species of Tetraodontidae in phylogenetic analysis of Sanger-sequenced mitochondrial and nuclear genes (Santini, Nguyen, et al. 2013, fig. 1; Mar'ie and Allam 2019, figs. 1, 4). Phylogenetic relationships of the major lineages of Tetraodontoidei are presented in Figure 21. The placements of the fossil taxa †Balkaria, †Bolcabalistes, †Ctenoplectus, †Eomola, †Eoplectus, †Eospinus, †Iraniplectus, †Moclaybalistes, †Proaracana, †Protobalistum, †Spinacanthus, and †Zignoichthys in the phylogeny are on the basis of inferences from morphological characters (Santini and Tyler 2003, 2004; Tyler et al. 2006; Close et al. 2016; Arcila and Tyler 2017; Bannikov et al. 2017; Carnevale et al. 2021; Troyer et al. 2022).

  • Phylogenetics. Most of the lineages classified in Tetraodontoidei were grouped together in the early 19th century in one of the first comprehensive classifications of teleosts (Cuvier 1816). Essentially all post-Darwinian classifications of teleosts that predate Hennigan phylogenetic systematics recognize the tetraodontoids as a lineage sharing common ancestry (e.g., Cope 1871a; T. N. Gill 1872, 1884a; Regan 1903; Boulenger 1904a:189–190; Jordan 1923:239–241; Berg 1940:495–497; Greenwood et al. 1966).

  • Some of the earliest phylogenetic analyses of ray-finned fishes focused on relationships within Tetraodontoidei, and there are several phylogenetic analyses based on morphological and molecular datasets (Winterbottom 1974; Leis 1984; Rosen 1984; Santini and Tyler 2003, 2004; Holcroft 2005; Alfaro et al. 2007; Yamanoue et al. 2007; Betancur-R, Broughton, et al. 2013; Santini, Sorenson, et al. 2013; Arcila et al. 2015; Close et al. 2016; Arcila and Tyler 2017; Bannikov et al. 2017; Ghezelayagh et al. 2022; Troyer et al. 2022). While there are important differences among nearly all the hypothesized phylogenies of tetraodontoids, most analyses consistently resolve three to four sets of sister lineages that include Triacanthodidae (spikefishes) and Triacanthidae (triplespines); Diodontidae (porcupinefishes) and Tetraodontidae (puffers); Balistidae (triggerfishes) and Monacanthidae (filefishes); and Aracanidae (deepwater boxfishes) and Ostraciidae (boxfishes) (Winterbottom 1974; Santini and Tyler 2003; Alfaro et al. 2007; Betancur-R, Broughton, et al. 2013; Near et al. 2013; Santini, Sorenson, et al. 2013; Arcila et al. 2015; Matsuura 2015; Arcila and Tyler 2017; Bannikov et al. 2017; Ghezelayagh et al. 2022; Troyer et al. 2022).

  • Most phylogenetic analyses differ on how the four sets of sister lineages are related to one another and are also incongruent regarding the relationships of Triodon macropterus (Threetooth Puffer) and Molidae (molas and ocean sunfishes). Phylogenies inferred from morphology or combinations of morphological and molecular characters resolve Triodon as the sister lineage of a clade containing Molidae, Diodontidae, and Tetraodontidae (Winterbottom 1974; Santini and Tyler 2003; Arcila et al. 2015; Arcila and Tyler 2017). The monophyly of this group was inferred, in part, on the basis of the upper and lower jaws with beak-like teeth and a nonprotractile upper jaw (Santini and Tyler 2003). Early molecular studies are incongruent in the relationships of Triodon and Molidae (Holcroft 2005; Alfaro et al. 2007; Yamanoue et al. 2007; Santini, Sorenson, et al. 2013). Phylogenomic analyses are congruent with one another in resolving three major lineages of Tetraodontoidei: (1) Aracanidae and Ostraciidae; (2) Triacanthodidae, Triacanthidae, Balistidae, and Monacanthidae; and (3) Molidae, Diodontidae, and Tetraodontidae (Figure 21; Ghezelayagh et al. 2022; Troyer et al. 2022). Triodon resolves as the sister lineage of a clade containing Aracanidae and Ostraciidae (Figure 21; Ghezelayagh et al. 2022).

  • Plectocretacicoidei is a lineage of four fossil tiny armored acanthomorph taxa from the Cretaceous of Italy (†Cretatriacanthus), Slovenia (†Protriacanthus and †Slovenitriacanthus), and Lebanon (†Plectocretacius), ranging in age from the Cenomanian (100.5–93.9 Ma) to the Campanian (83.2–72.2 Ma) and hypothesized to be the sister lineage of Tetraodontoidei (Tyler and Sorbini 1996; Santini and Tyler 2003, 2004; Tyler and Santini 2005; Tyler and KriŽnar 2013; Close et al. 2016; Arcila and Tyler 2017; Bannikov et al. 2017). When the hypothesis was introduced by Tyler and Sorbini (1996), it was assumed the paracanthopterygian Zeiformes and Tetraodontoidei were sister lineages (Rosen 1984), limiting the sampling of outgroups and not ensuring a robust test of the monophyly in the morphological analyses (Santini and Tyler 2003; Arcila et al. 2015; Close et al. 2016; Arcila and Tyler 2017; Bannikov et al. 2017; Troyer et al. 2022). Regardless, phylogenetic analyses using the limited set of outgroups do not consistently resolve plectocretacicoids and tetraodontoids as sister lineages (Arcila and Tyler 2017). Many of the morphological features presented as synapomorphies in support of plectocretacicoid-tetraodontoid monophyly (e.g., absence of anal-fin spines, absence of ribs, reduced number of vertebrae, reduced number of caudal-fin rays, restricted opercular opening) are found in many or all species of Lophioidei (Regan 1912c; Pietsch 1981, 1984; Chanet et al. 2013), suggesting these features are synapomorphies for a more inclusive clade within acanthuriforms (Benton et al. 2015; A. C. Gill and Leis 2019). Tetraodontoidei and Lophioidei share several derived morphological traits not present in †Plectocretacicoidei, which include: absence of infraorbital bones (Pietsch 1981; Carnevale and Pietsch 2012); sutural relationship between the posttemporal and cranium (Pietsch 1981; Carnevale and Pietsch 2012); six or fewer branchiostegal rays, a trait shared with several other lineages of Acanthuriformes (McAllister 1968; Benton et al. 2015; A. C. Gill and Leis 2019); lateral line unenclosed by bony canals (Nakae and Sasaki 2010); and absence of procurrent caudal rays (Pietsch 1981, 1984).

  • Composition. There are currently 433 living species of Tetraodontoidei (Nyegaard et al. 2018; Fricke et al. 2023) that include Triodon macropterus and species classified in Aracanidae, Balistidae, Diodontidae, Molidae, Monacanthidae, Ostraciidae, Tetraodontidae, Triacanthidae, and Triacanthodidae. Fossil tetraodontoid taxa include †Balkaria, †Bolcabalistes, †Ctenoplectus, †Eomola, †Eoplectus, †Eospinus, †Iraniplectus, †Moclaybalistes, †Proaracana, †Protobalistum, †Spinacanthus, and †Zignoichthys. Details of the ages and locations of fossil taxa are given in Appendix 1. Over the past 10 years, 10 new living species of Tetraodontoidei have been described (Fricke et al. 2023), comprising 2.3% of the living species diversity in the clade.

  • Diagnostic apomorphies. Morphological apomorphies of Tetraodontoidei include (1) anal spines absent (Rosen 1984; Tyler and Sorbini 1996; Wiley and Johnson 2010), (2) caudal fin with 12 or fewer principal rays (Rosen 1984; Tyler and Sorbini 1996; Wiley and Johnson 2010), (3) infraorbitals absent (Rosen 1984; Tyler and Sorbini 1996), (4) parietals absent (Rosen 1984; Tyler and Sorbini 1996; Wiley and Johnson 2010), (5) a small gill opening slightly anterior to the pectoral fin base (Rosen 1984; Tyler and Sorbini 1996; Wiley and Johnson 2010), (6) posterior process of pelvic basipterygia fused or sutured medially (Rosen 1984; Tyler and Sorbini 1996; Wiley and Johnson 2010), (7) pelvic fin with no more than one spine and two rays (Rosen 1984; Tyler and Sorbini 1996; Wiley and Johnson 2010), (8) nasal bones absent (Tyler and Sorbini 1996; Wiley and Johnson 2010), and (9) sensory canal in dentary absent (Tyler and Sorbini 1996).

  • Synonyms. Plectognathes (Cuvier 1816:144–155), Plectognathi (Haeckel 1866:cxxviii; Günther 1870:207–320; Cope 1871a:456, 458; T. N. Gill 1872:1; Regan 1903:285–286, 1929:325–326; Boulenger 1904a:189–190, 1904b:721–727; Jordan 1923:239–241, Gymnodontes (Cuvier 1816:145), and Tetraodontiformes (Berg 1940:495–497; Greenwood et al. 1966:403; Gosline 1971:169–170) are ambiguous synonyms of Tetraodontoidei.

  • Comments. Tetraodontoidei has a rich fossil record with the earliest crown lineage taxa from the Ypresian (56.0–48.1 Ma) in localities that include Denmark, Italy, Russia, United Kingdom, and Turkmenistan (Figure 21; Appendix 2; e.g., Tyler and Bannikov 1992; Tyler and Santini 2002; Close et al. 2016; Bannikov et al. 2017). Phylogenetic analyses integrating morphological and molecular datasets to resolve relationships among extinct and living lineages of Tetraodontoidei advanced the practice of tip-dating, in which fossil taxa in phylogenies provide time calibration in relaxed clock analyses (Arcila et al. 2015), show the effect of changes in paleoclimate on extinction dynamics (Arcila and Tyler 2017), and reveal the relationship between changes in paleoclimate and the evolution of body size (Troyer et al. 2022). Bayesian relaxed molecular clock analyses of Tetraodontoidei result in an average posterior crown age estimate of 62.5 million years ago, with the credible interval ranging between 60.5 and 87.3 million years ago (Troyer et al. 2022).

  • img-z183-2_03.gif

    Acknowledgments

    We thank Rosemary Volpe and Patrick Sweeney for outstanding editorial support throughout the production of this monograph. Julie Johnson of Life Science Studios, LLC, created the paintings of fish species used in the phylogeny figures. We are grateful for all of our colleagues who read and reviewed portions of this monograph: Gloria Arratia (Albulidae, Elopiformes, Elopomorpha, Oseanacephala, and Teleostei), Kevin Conway (Cobitoidei and Cyprinoidei), Jessica Glass (Carangiformes, Carangoidei, and Pleuronectoidei), Richard Harrington (Anabantoidei, Carangiformes, Carangoidei, Pleuronectoidei, and Synbranchiformes), Bruno Melo (Characiformes, Cithariniformes, Gymnotiformes, Loricarioidei, Ostariophysi, Otophysi, Siluriformes, and Siluroidei), Peter Rask Møller and Jørgen Nielsen (Ophidiiformes and Bythitoidei), Rene Martin (Ctenosquamata and Myctophiformes), Timo Moritz (Argentiniformes, Euteleostei, Polypteridae, and Otophysi), Theodore Pietsch (Lophioidei), Kyle Piller (Atheriniformes, Atherinoidei, Belonoidei, and Cyprinodontoidei), João Paulo C. B. da Silva (Anguilliformes, Anguilloidei, Congroidei, Muraenoidei, and Synaphobranchoidei), Melanie Stiassny (Acanthomorpha, Acanthopterygii, Blenniiformes, and Percomorpha), Peter Wainwright (Blenniiformes, Labriformes, and Perciformes), Mark Westneat (Labriformes), Mark Wilson (Osteoglossomorpha, Osteoglossiformes, Osteoglossidae, Salmoniformes, Esocidae, and Percopsiformes), and Diego Vaz (Batrachoididae). Prosanta Chakrabarty, Bruno Melo, Jon Moore, and Kevin de Queiroz provided peer reviews for the journal. Leo Smith and John Sullivan identified errors in preprint versions of the monograph. Alex Dornburg, Kevin de Queiroz, and Larry Page provided advice and encouragement through the course of our work on this monograph. The Bingham Oceanographic fund maintained by the Yale Peabody Museum, Yale University, provided support for this research. The resources and staff at the Yale University Library were fundamental to the completion of this project. Thacker thanks Bill Fink and Bill Gosline for their insights into actinopterygian phylogeny and evolution. Near thanks Allison Near, Alice Near, and Rebecca Near for their unwavering support and humor though the course of preparing this monograph.

    © 2024 Yale Peabody Museum, Yale University. All rights reserved.  https://peabody.yale.edu

    Literature Cited

    1.

    Aarn, and W. Ivantsoff. 1997. Descriptive anatomy of Cairnsichthys rhombosomoides and Iriatherina werneri (Teleostei: Atheriniformes), and a phylogenetic analysis of Melanotaeniidae. Ichthyological Exploration of Freshwaters 8(2):107–150. Google Scholar

    2.

    Aarn, W. Ivantsoff, and M. Kottelat. 1998. Phylogenetic analysis of Telmatherinidae (Teleostei: Atherinomorpha), with description of Marosatherina, a new genus from Sulawesi. Ichthyological Exploration of Freshwaters 9(3):311–323. Google Scholar

    3.

    Adamson, E. A. S., D. A. Hurwood, and P. B. Mather. 2010. A reappraisal of the evolution of Asian snakehead fishes (Pisces, Channidae) using molecular data from multiple genes and fossil calibration. Molecular Phylogenetics and Evolution 56(2):707–717. Google Scholar

    4.

    Afonso, G. V. F., F. Di Dario, L. N. Eduardo, F. Lucenafrédou, A. Bertrand, and M. M. Mincarone. 2021. Taxonomy and distribution of deep-sea bigscales and whalefishes (Teleostei: Stephanoberycoidei) collected off northeastern Brazil, including seamounts and oceanic islands. Ichthyology and Herpetology 109(2):467–488. Google Scholar

    5.

    Afsari, S., M. Yazdi, A. Bahrami, and G. Carnevale. 2014. A new deep-sea hatchetfish (Teleostei: Stomiiformes: Sternoptychidae) from the Eocene of Ilam, Zagros Basin, Iran. Bollettino della Società Paleontologica Italiana 53(1):27–37. Google Scholar

    6.

    Agassiz, L. 1835. Recherches sur les Poissons Fossiles. Tome 4. Neuchatel, Switzerland: Petitpierre. 318 pp. Google Scholar

    7.

    Agassiz, L. 1853. Extraordinary fishes from California, constituting a new family, described by L. Agassiz. American Journal of Science and Arts, Series 2, 16:380–390. Google Scholar

    8.

    Agassiz, L. 1854. Appendix–Additional notes on the Holconoti. American Journal of Science and Arts, Series 2, 17:365–369. Google Scholar

    9.

    Ahl, E. 1924. Neue afrikanische Zahnkarpfen aus dem Zoologischen Museum Berlin. Zoologischer Anzeiger 61:135–145. Google Scholar

    10.

    Ahlstrom, E. H., H. G. Moser, and D. M. Cohen. 1984. Argentinoidei: development and relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 155–169. (Special Publication 1.) Google Scholar

    11.

    Aiello, B. R., M. W. Westneat, and M. E. Hale. 2017. Mechanosensation is evolutionarily tuned to locomotor mechanics. Proceedings of the National Academy of Sciences 114(17):4459–4464. Google Scholar

    12.

    Albert, J. S. 2001. Species Diversity and Phylogenetic Systematics of American Knifefishes (Gymnotiformes, Teleostei). Ann Arbor: University of Michigan Museum of Zoology. 127 pp. (Miscellaneous Publications 190.) Google Scholar

    13.

    Albert, J. S., and W. G. R. Crampton. 2005. Diversity and phylogeny of Neotropical electric fishes (Gymnotiformes). In: T. H. Bullock, C. D. Hopkins, A. N. Popper, and R. R. Fay, eds. Electroreception. Berlin: Springer. pp. 360–409. Google Scholar

    14.

    Albert, J. S., and W. L. Fink. 2007. Phylogenetic relationships of fossil Neotropical electric fishes (Osteichthyes: Gymnotiformes) from the Upper Miocene of Bolivia. Journal of Vertebrate Paleontology 27(1):17–25. Google Scholar

    15.

    Albert, J. S., M. J. Lannoo, and T. Yuri. 1998. Testing hypotheses of neural evolution in gymnotiform electric fishes using phylogenetic character data. Evolution 52(6):1760–1780. Google Scholar

    16.

    Alcock, A. 1890. Natural history notes from H. M. Indian marine survey steamer ‘Investigator,’ Commander R. F. Hoskyn, R. N., commanding.—No. 18. On the bathybial fishes of the Arabian Sea, obtained during the sea-son 1889–90. Annals and Magazine of Natural History, Series 6, 6(34):295–311. Google Scholar

    17.

    Alda, F., V. A. Tagliacollo, M. J. Bernt, B. T. Waltz, W. B. Ludt, B. C. Faircloth, M. E. Alfaro, J. S. Albert, and P. Chakrabarty. 2019. Resolving deep nodes in an ancient radiation of Neotropical fishes in the presence of conflicting signals from incomplete lineage sorting. Systematic Biology 68(4):573–593. Google Scholar

    18.

    Alfaro, M. E., C. D. Brock, B. L. Banbury, and P. C. Wainwright. 2009. Does evolutionary innovation in pharyngeal jaws lead to rapid lineage diversification in labrid fishes? BMC Evolutionary Biology 9:255.  https://doi.org/10.1186/1471-2148-9-255 Google Scholar

    19.

    Alfaro, M. E., B. C. Faircloth, R. C. Harrington, L. Sorenson, M. Friedman, C. E. Thacker, C. H. Oliveros, D. Černý, and T. J. Near. 2018. Explosive diversification of marine fishes at the Cretaceous–Palaeogene boundary. Nature Ecology and Evolution 2:688–696.  https://doi.org/10.1038/s41559-018-0494-6 Google Scholar

    20.

    Alfaro, M. E., F. Santini, and C. D. Brock. 2007. Do reefs drive diversification in marine teleosts? Evidence from the pufferfishes and their allies (order Tetraodontiformes). Evolution 61(9):2104–2126. Google Scholar

    21.

    Alfaro, M. E., F. Santini, C. Brock, H. Alamillo, A. Dornburg, D. L. Rabosky, G. Carnevale, and L. J. Harmon. 2009. Nine exceptional radiations plus high turnover explain species diversity in jawed vertebrates. Proceedings of the National Academy of Sciences of the United States of America 106(32):13410–13414. Google Scholar

    22.

    Allen, G. R. 1985. Snappers of the World. Vol. 6 of FAO Species Catalogue. Rome: Food and Agriculture Organization of the United Nations. 208 pp. (FAO Fisheries Synopsis 125.) Google Scholar

    23.

    Almada, F., V. C. Almada, T. Guillemaud, and P. Wirtz. 2005. Phylogenetic relationships of the north-eastern Atlantic and Mediterranean blenniids. Biological Journal of the Linnean Society 86(3):283–295. Google Scholar

    24.

    Altner, M., and B. Reichenbacher. 2015. †Kenyaichthyidae fam. nov. and †Kenyaichthys gen. nov.—first record of a fossil aplocheiloid killifish (Teleostei, Cyprinodontiformes). PLOS ONE 10(4):e0123056.  https://doi.org/10.1371/journal.pone.0123056 Google Scholar

    25.

    Alvarado-Ortega, J., E. Ovalles-Damián, and G. Arratia. 2008. A review of the interrelationships of the order Ellimmichthyiformes (Teleostei: Clupeomorpha). In: G. Arratia, H. P. Schultze, and M. V. H. Wilson, eds. Mesozoic Fishes 4: Homology and Phylogeny. Munich: Verlag Dr. Friedrich Pfeil. pp. 257–278. Google Scholar

    26.

    Alvarado-Ortega, J., and B. A. Than-Marchese. 2012. A Cenomanian aipichthyoid fish (Teleostei, Acanthomorpha) from America, Zoqueichthys carolinae gen. and sp. nov. from El Chango quarry (Cintalapa Member, Sierra Madre Formation), Chiapas, Mexico. Revista Mexicana de Ciencias Geológicas 29(3):735–748. Google Scholar

    27.

    Alvarado-Ortega, J.2013. The first record of a North American Cenomanian Trachichthyidae fish (Acanthomorpha, Acanthopterygii), Pepemkay maya, gen. et sp. nov., from El Chango quarry (Sierra Madre Formation), Chiapas, Mexico. Journal of Vertebrate Paleontology 33(1):48–57. Google Scholar

    28.

    Alvarado-Ortega, J., B. A. Than-Marchese, and M. P. Melgarejo-Damián. 2020. On the Albian occurrence of Armigatus (Teleostei, Clupeomorpha) in America, a new species from the Tlayúa Lagerstätte, Mexico. Palaeontologia Electronica 23(3):a52.  https://doi.org/10.26879/1107 Google Scholar

    29.

    Alves, Y. M., J. Alvarado-Ortega, and P. M. Brito. 2020. †Epaelops martinezi gen. and sp. nov. from the Albian limestone deposits of the Tlayúa quarry, Mexico—a new late Mesozoic record of Elopiformes of the western Tethys. Cretaceous Research 110:104260.  https://doi.org/10.1016/j.cretres.2019.104260 Google Scholar

    30.

    Alves-Gomes, J. A., G. Ortí, M. Haygood, W. Heiligenberg, and A. Meyer. 1995. Phylogenetic analysis of the South American electric fishes (order Gymnotiformes) and the evolution of their electrogenic system: a synthesis based on morphology, electrophysiology, and mitochondrial sequence data. Molecular Biology and Evolution 12(2):298–318. Google Scholar

    31.

    Amaoka, K. 1969. Studies on the sinistral flounders found in the waters around Japan: taxonomy, anatomy and phylogeny. Journal of the Shimonoseki University of Fisheries 18:65–340. Google Scholar

    32.

    Amaral, C. R. L., J. Avarado-Ortega, and P. Brito. 2013. Sapperichthys gen. nov., a new gonorynchid from the Cenomanian of Chiapas, Mexico. In: G. Arratia, H.-P. Schultze, and M. V. H. Wilson, eds. Mesozoic Fishes 5: Global Diversity and Evolution. Munich: Verlag Dr. Friedrich Pfeil. pp. 305–323. Google Scholar

    33.

    Amaral, C. R. L., and P. M. Brito. 2012. A new Chanidae (Ostariophysii: Gonorynchiformes) from the Cretaceous of Brazil with affinities to Laurasian gonorynchiforms from Spain. PLOS ONE 7(5):e37247.  https://doi.org/10.1371/journal.pone.0037247 Google Scholar

    34.

    Amorim, P. F., and W. J. E. M. Costa. 2018. Multigene phylogeny supports diversification of four-eyed fishes and one-sided livebearers (Cyprinodontiformes: Anablepidae) related to major South American geological events. PLOS ONE 13(6):e0199201.  https://doi.org/10.1371/journal.pone.0199201 Google Scholar

    35.

    Anðelković, J. S. 1989. Tertiary Fishes of Yugoslavia: A Stratigraphic–Paleontologic–Paleoecological Study = Tercijarne Ribe Jugoslavije: Stratigrafsko–PaleontoloŠko–PaleoekoloŠka Studija. Zagreb: Jugoslavenska akademija znanosti i umjetnosti. 121 pp. (Palaeontologia Jugoslavica 38.) Google Scholar

    36.

    Anderson, M. E. 1984. On the anatomy and phylogeny of Zoarcidae (Teleostei: Perciformes) [dissertation]. College of William and Mary. 254. pp. Dissertations, Theses, and Masters Projects, Paper 1539616550.  https://scholarworks.wm.edu/etd/1539616550/ Google Scholar

    37.

    Anderson, M. E. 1994. Systematics and Osteology of the Zoarcidae (Teleostei: Perciformes). Grahamstown, South Africa: Rhodes University. 120 pp. (Ichthyological Bulletin of the J. L. B. Smith Institute of Ichthyology 60.) Google Scholar

    38.

    Anderson, M. E. 1998. A late Cretaceous (Maastrichtian) Galaxiid Fish from South Africa. Grahamstown, South Africa: Rhodes University. 12 pp. (Special Publication of the J. L. B. Smith Institue of Ichthyology 60.) Google Scholar

    39.

    Andrews, A. C. 1955. Greek and Latin terms for salmon and trout. Transactions and Proceedings of the American Philological Association 86(1955):308–318. Google Scholar

    40.

    Andrews, J. V., J. P. Schein, and M. Friedman. 2023. An earliest Paleocene squirrelfish (Teleostei: Beryciformes: Holocentroidea) and its bearing on the timescale of holocentroid evolution. Journal of Systematic Palaeontology 21(1):2168571.  https://doi.org/10.1080/14772019.2023.2168571 Google Scholar

    41.

    Andriashev, A. P., and V. M. Makushok. 1955. Azygopterus corallinus (Pisces, Blennioidei): a new fish without paired fins. Voprosy Ikhtiologii 3:50–53. [In Russian.] Google Scholar

    42.

    Arambourg, C. 1966. Résultats scientifiques de la mission C. Arambourg en Syrie et en Iran (1938–1939). II. Les poissons oligocènes de l'Iran. Paris: Muséum national d'histoire naturelle. 210 pp. (Notes et Mémoires sur le Moyen-Orient 8.) Google Scholar

    43.

    Arbour, V. M., M. E. Burns, and R. L. Sissons. 2009. A redescription of the ankylosaurid dinosaur Dyoplosaurus acutosquameus Parks, 1924 (Ornithischia: Ankylosauria) and a revision of the genus. Journal of Vertebrate Paleontology 29(4):1117–1135. Google Scholar

    44.

    Arcila, D., L. C. Hughes, B. Melendez-Vazquez, C. C. Baldwin, W. T. White, K. E. Carpenter, J. T. Williams, M. D. Santos, J. J. Pogonoski, M. Miya, G. Ortí, and R. Betancur-R. 2021. Testing the utility of alternative metrics of branch support to address the ancient evolutionary radiation of tunas, stromateoids, and allies (Teleostei: Pelagiaria). Systematic Biology 70(6):1123–1144. Google Scholar

    45.

    Arcila, D., G. Ortí, R. Vari, J. W. Armbruster, M. L. J. Stiassny, K. D. Ko, M. H. Sabaj, J. Lundberg, L. J. Rev-Ell, and R. Betancur-R. 2017. Genome-wide interrogation advances resolution of recalcitrant groups in the tree of life. Nature Ecology and Evolution 1:0020.  https://doi.org/10.1038/s41559-016-0020 Google Scholar

    46.

    Arcila, D., P. Petry, and G. Ortí. 2018. Phylogenetic relationships of the family Tarumaniidae (Characiformes) based on nuclear and mitochondrial data. Neotropical Ichthyology 16(3):e180016.  https://doi.org/10.1590/1982-0224-20180016 Google Scholar

    47.

    Arcila, D., R. A. Pyron, J. C. Tyler, G. Ortí, and R. Betancur-R. 2015. An evaluation of fossil tip-dating versus node-age calibrations in tetraodontiform fishes (Teleostei: Percomorphaceae). Molecular Phylogenetics and Evolution 82(Part A):131–145. Google Scholar

    48.

    Arcila, D., and J. C. Tyler. 2017. Mass extinction in tetraodontiform fishes linked to the Palaeocene–Eocene thermal maximum. Proceedings of the Royal Society B, Biological Sciences 284(1866):20171771. Google Scholar

    49.

    Argyriou, T., S. Giles, and M. Friedman. 2022. A Permian fish reveals widespread distribution of neopterygian-like jaw suspension. eLife 11:e58433.  https://doi.org/10.7554/eLife.58433 Google Scholar

    50.

    Argyriou, T., S. Giles, M. Friedman, C. Romano, I. Kogan, and M. R. Sánchez-Villagra. 2018. Internal cranial anatomy of Early Triassic species of †Saurichthys (Actinopterygii: †Saurichthyiformes): implications for the phylogenetic placement of †saurichthyiforms. BMC Evolutionary Biology 18:161.  https://doi.org/10.1186/s12862-018-1264-4 Google Scholar

    51.

    Armbruster, J. W., M. L. Niemiller, and P. B. Hart. 2016. Morphological evolution of the cave-, spring-, and swampfishes of the Amblyopsidae (Percopsiformes). Copeia 104(3):763–777. Google Scholar

    52.

    Arnold, R. J. 2014. Evolutionary relationships of the enigmatic anglerfishes (Teleostei: Lophiiformes): can nuclear DNA provide resolution for conflicting morphological and mitochondrial phylogenies? [dissertation]. University of Washington. 92 pp. Google Scholar

    53.

    Arratia, G. 1981. Varasichthys ariasi n. gen. et sp. from the Upper Jurassic of Chile (Pisces, Teleostei, Varasichthyidae n. fam.). Palaeontographica Abteilung A Palaeozoologie-Stratigraphie 175:107–139. Google Scholar

    54.

    Arratia, G. 1987a. Anaethalion and similar teleosts (Actinopterygii, Pisces) from the Late Jurassic (Tithonian) of southern Germany and their relationships. Palaeontographica Abteilung A 200:1–44. Google Scholar

    55.

    Arratia, G. 1987b. Description of the Primitive Family Diplomystidae (Siluriformes, Teleostei, Pisces): Morphology, Taxonomy and Phylogenetic Implications. Bonn: Zoologisches Forschungsinstitut und Museum Alexander Koenig. 120 pp. (Bonner Zoologische Monographien 24.) Google Scholar

    56.

    Arratia, G. 1991. The caudal skeleton of Jurassic teleosts: a phylogenetic analysis. In: M.-M. Chang, Y.-H. Liu, and G.-R. Zhang, eds. Early Vertebrates and Related Problems in Evolutionary Biology. Beijing: Science Press. pp. 249–340. Google Scholar

    57.

    Arratia, G. 1992. Development and Variation of the Suspensorium of Primitive Catfishes (Teleostei: Ostariophysi) and Their Phylogenetic Relationships. Bonn: Zoologisches Forschungsinstitut und Museum Alexander Koenig. 140 pp. (Bonner Zoologische Monographien 32.) Google Scholar

    58.

    Arratia, G. 1996a. The Jurassic and the early history of teleosts. In: G. Arratia and G. Viohl, eds. Mesozoic Fishes: Systematics and Paleoecology. Munich: Verlag Dr. Friedrich Pfeil. pp. 243–259. Google Scholar

    59.

    Arratia, G. 1996b. Reassessment of the phylogenetic relationships of certain Jurassic teleosts and their implications on teleost phylogeny. In: G. Arratia and G. Viohl, eds. Mesozoic Fishes: Systematics and Paleoecology. Munich: Verlag Dr. Friedrich Pfeil. pp. 219–242. Google Scholar

    60.

    Arratia, G. 1997. Basal Teleosts and Teleostean Phylogeny. Munich: Verlag Dr. Friedrich Pfeil. 168 pp. (Palaeo Ichthyologica 7.) Google Scholar

    61.

    Arratia, G. 1999. The monophyly of Teleostei and stem-group teleosts. Consensus and disagreements. In: G. Arratia and H.-P. Schultze, eds. Mesozoic Fishes 2: Systematics and Fossil Record. Munich: Verlag Dr. Friedrich Pfeil. pp. 265–334. Google Scholar

    62.

    Arratia, G. 2000a. New teleostean fishes from the Jurassic of southern Germany and the systematic problems concerning the ‘pholidophoriforms.’ Paläontologische Zeitschrift 74:113–143. Google Scholar

    63.

    Arratia, G. 2000b. Phylogenetic relationships of Teleostei: past and present. Estudios Oceanologicos 19:19–51. Google Scholar

    64.

    Arratia, G. 2000c. Remarkable teleostean fishes from the Late Jurassic of southern Germany and their phylogenetic relationships. Fossil Record 3(1):137–179. Google Scholar

    65.

    Arratia, G. 2001. The sister-group of Teleostei: consensus and disagreements. Journal of Vertebrate Paleontology 21(4):767–773. Google Scholar

    66.

    Arratia, G. 2003a. Catfish head skeleton—an overview. In: G. Arratia, B. G. Kapoor, M. Chardon, and R. Diogo, eds. Catfishes, Volume 1. Enfield, NH: Science Publishers. pp. 3–46. Google Scholar

    67.

    Arratia, G. 2003b. The siluriform postcranial skeleton—an overview. In: G. Arratia, B. G. Kapoor, M. Chardon, and R. Diogo, eds. Catfishes, Volume 1. Enfield, NH: Science Publishers. pp. 121–157. Google Scholar

    68.

    Arratia, G. 2008. The varasichthyid and other crossognathiform fishes, and the break-up of Pangaea. In: L. Cavin, A. Longbottom, and M. Richter, eds. Fishes and the Break-up of Pangaea. London: Geological Society of London. pp. 71–92. Google Scholar

    69.

    Arratia, G. 2010a. The Clupeocephala re-visited: analysis of characters and homologies. Revista de Biología Marina y Oceanografía 45(S1):635–657. Google Scholar

    70.

    Arratia, G. 2010b. Critical analysis of the impact of fossils on teleostean phylogenies, especially that of basal teleosts. In: D. K. Elliot, J. G. Maisey, X. Yu, and D. Miao, eds. Morphology, Phylogeny, and Paleobiogeography of Fossil Fishes. Munich: Verlag Dr. Friedrich Pfeil. pp. 247–274. Google Scholar

    71.

    Arratia, G. 2013. Morphology, taxonomy, and phylogeny of Triassic pholidophorid fishes (Actinopterygii, Teleostei). Journal of Vertebrate Paleontology 33(Suppl. 1):1–138. Google Scholar

    72.

    Arratia, G. 2016. New remarkable Late Jurassic teleosts from southern Germany: Ascalaboidae n. fam., its content, morphology, and phylogenetic relationships. Fossil Record 19(1):31–59. Google Scholar

    73.

    Arratia, G. 2017. New Triassic teleosts (Actinopterygii, Teleosteomorpha) from northern Italy and their phylogenetic relationships among the most basal teleosts. Journal of Vertebrate Paleontology 37(2):e1312690.  https://doi.org/10.1080/02724634.2017.1312690 Google Scholar

    74.

    Arratia, G. 2018.Otomorphs(=otocephalansorostarioclupeomorphs) revisited. Neotropical Ichthyology 16(18):e180079.  https://doi.org/10.1590/1982-0224-20180079 Google Scholar

    75.

    Arratia, G., and A. Cione. 1996. The record of fossil fishes of southern South America. In: G. Arratia, ed. Münchner Geowissenshaftliche Abhandlungen. Munich: Verlag Dr. Friedrich Pfeil. pp. 9–72. Google Scholar

    76.

    Arratia, G., and M. Gayet. 1995. Sensory canals and related bones of Tertiary siluriform crania from Bolivia and North America and comparison with recent forms. Journal of Vertebrate Paleontology 15(3):482–505. Google Scholar

    77.

    Arratia, G., and H.-P. Schultze. 2024. The oldest teleosts (Teleosteomorpha): their early taxonomic, phenotypic, and ecological diversification during the Triassic. Fossil Record 27(1):29–53. Google Scholar

    78.

    Arratia, G., H.-P. Schultze, S. Gouiric-Cavalli, and C. Quezada-Romegialli. 2021. The intriguing †Atacamichthys fish from the Middle Jurassic of Chile—an amiiform or teleosteomorph? In: A. Pradel, J. S. S. Denton, and P. Janvier, eds. Ancient Fishes and Their Living Relatives: A Tribute to John G. Maisey. Munich: Verlag Dr. Friedrich Pfeil. pp. 19–36. Google Scholar

    79.

    Arratia, G., and D. Thies. 2001. A new teleost (Osteichthyes, Actinopterygii) from the Early Jurassic Posidonia shale of northern Germany. Fossil Record 4:167–187. Google Scholar

    80.

    Arratia, G., and H. Tischlinger. 2010. The first record of Late Jurassic crossognathiform fishes from Europe and their phylogenetic importance for teleostean phylogeny. Fossil Record 13(2):317–341. Google Scholar

    81.

    Arroyave, J., J. S. S. Denton, M. L. J. Stiassny, and W. J. Murphy. 2013. Are characiform fishes Gondwanan in origin? Insights from a time-scaled molecular phylogeny of the Citharinoidei (Ostariophysi: Characiformes). PLOS ONE 8(10):e77269.  https://doi.org/10.1371/journal.pone.0077269 Google Scholar

    82.

    Arroyave, J., A. F. Mar-Silva, and P. Díaz-Jaimes. 2022. The complete mitochondrial genome of the Mexican blind brotula Typhlias pearsei (Ophidiiformes: Dinematichthydae): an endemic and troglomorphic cavefish from the Yucatán Peninsula karst aquifer. Mitochondrial DNA Part B 7:1151–1153. Google Scholar

    83.

    Arroyave, J., and M. L. J. Stiassny. 2011. Phylogenetic relationships and the temporal context for the diversification of African characins of the family Alestidae (Ostariophysi: Characiformes): evidence from DNA sequence data. Molecular Phylogenetics and Evolution 60(3):385–397. Google Scholar

    84.

    Artedi, P. 1738. Genera piscium. In: Quibus Systema Totum Ichthyologiæ Proponitus cum Classibus, Ordinibus, Generum Characteribus, Specierum Differentiis, Observationibus Plurimis. Redactis Speciebus 242 ad Genera 52. Ichthyologiæ Pars 3. Leiden: Lugduni Batavorum apud Conradum Wishoff. 88 pp. Google Scholar

    85.

    Artyukhin, E. N. 2006. Morphological phylogeny of the order Acipenseriformes. Journal of Applied Ichthyology 22(Suppl. 1):66–69. Google Scholar

    86.

    Ascanius, P. 1767. Icones rerum naturalium, ou figures enluminées d'histoire naturelle du Nord, premier cahier. Copenhagen: E. A. H. Möller. 22 pp. Google Scholar

    87.

    Ascanius, P. 1772. Icones rerum naturalium, ou figures enluminées d'histoire naturelle du nord, second cahier. Copenhagen: E. A. H. Möller. 8 pp. Google Scholar

    88.

    Aschliman, N. C., I. R. Tibbetts, and B. B. Collette. 2005. Relationships of sauries and needlefishes (Teleostei: Scomberesocoidea) to the internally fertilizing halfbeaks (Zenarchopteridae) based on the pharyngeal jaw apparatus. Proceedings of the Biological Society of Washington 118(2):416–427. Google Scholar

    89.

    Astolfi, L., I. Dupanloup, R. Rossi, P. M. Bisol, E. Faure, and L. Congiu. 2005. Mitochondrial variability of sand smelt Atherina boyeri populations from north Mediterranean coastal lagoons. Marine Ecology Progress Series 297:233–243. Google Scholar

    90.

    Astudillo-Clavijo, V., M. L. J. Stiassny, K. L. Ilves, Z. Musilova, W. Salzburger, and H. López-Fernández. 2023. Exon-based phylogenomics and the relationships of African cichlid fishes: tackling the challenges of reconstructing phylogenies with repeated rapid radiations. Systematic Biology 72(1):134–149. Google Scholar

    91.

    Atta, C. J., H. Yuan, C. Li, D. Arcila, R. Betancur-R, L. C. Hughes, G. Ortí, and L. Tornabene. 2022. Exon-capture data and locus screening provide new insights into the phylogeny of flatfishes (Pleuronectoidei).MolecularPhylogeneticsandEvolution166:107315.  https://doi.org/10.1016/j.ympev.2021.107315 Google Scholar

    92.

    Austin, C. M., M. H. Tan, L. J. Croft, M. P. Hammer, and H. M. Gan. 2015. Whole genome sequencing of the Asian Arowana (Scleropages formosus) provides insights into the evolution of ray-finned fishes. Genome Biology and Evolution 7(10):2885–2895. Google Scholar

    93.

    Ayers, W. O. 1859 October 17. [Descriptions of fishes]. Proceedings of the California Academy of Sciences, Series 1, 2:25–32. Google Scholar

    94.

    Azevedo, M. F. C., C. Oliveira, B. G. Pardo, P. Martinez, and F. Foresti. 2008. Phylogenetic analysis of the order Pleuronectiformes (Teleostei) based on sequences of 12S and 16S mitochondrial genes. Genetics and Molecular Biology 31(Suppl. 1):284–292. Google Scholar

    95.

    Azpelicueta, M. D. L. M., and A. L. Cione. 2011. Redescription of the Eocene catfish Bachmannia chubutensis (Teleostei: Bachmanniidae) of southern South America. Journal of Vertebrate Paleontology 31(2):258–269. Google Scholar

    96.

    Baciu, D. S., and B. Chanet. 2002. Les poissons plats fossils (Teleostei: Pleuronectiformes) de l'Oligocène de Piatra Neamt (Roumanie). Oryctos 4:17–38. Google Scholar

    97.

    Baird, S. F., and C. Girard. 1853. Descriptions of new species of fishes collected by Mr. John H. Clark, on the U.S. and Mexican Boundary Survey, under Lt. Col. Jas. D. Graham. Proceedings of the Academy of Natural Science Philadelphia 6:387–390. Google Scholar

    98.

    Bakke, I., and S. D. Johansen. 2005. Molecular phylogenetics of Gadidae and related Gadiformes based on mitochondrial DNA sequences. Marine Biotechnology 7:61–69. Google Scholar

    99.

    Baldwin, C. C. 2013. The phylogenetic significance of colour patterns in marine teleost larvae. Zoological Journal of the Linnean Society 168(3):496–563. Google Scholar

    100.

    Baldwin, C. C., and G. D. Johnson. 1995. A larva of the Atlantic flashlight fish, Kryptophanaron alfredi (Beryciformes, Anomalopidae), with a comparison of beryciform and stephanoberyciform larvae. Bulletin of Marine Science 56:1–24. Google Scholar

    101.

    Baldwin, C. C.1996. Interrelationships of Aulopiformes. In: M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson, eds. Interrelationships of Fishes. San Diego: Academic Press. pp. 355–404. Google Scholar

    102.

    Baliga, V. B., and C. J. Law. 2016. Cleaners among wrasses: phylogenetics and evolutionary patterns of cleaning behavior within Labridae. Molecular Phylogenetics and Evolution 94(Part A):424–435. Google Scholar

    103.

    Balushkin, A. V. 1992. Classification, phylogenetic relationships, and origins of the families of the suborder Notothenioidei (Perciformes). Journal of Ichthyology 32(7):90–110. Google Scholar

    104.

    Balushkin, A. V. 1994. Proeleginops grandeastmanorum gen. et sp. nov. (Perciformes, Notothenioidei, Eleginopsidae) from the Late Eocene of Seymour Island (Antarctica) is a fossil notothenioid, not a gadiform. Journal of Ichthyology 34(8):10–23. Google Scholar

    105.

    Balushkin, A. V. 2000. Morphology, classification, and evolution of notothenioid fishes of the Southern Ocean (Notothenioidei, Perciformes). Journal of Ichthyology 40(S1):S74–S109. Google Scholar

    106.

    Bannikov, A. F. 1987. On the taxonomy, composition and origin of the family Carangidae. Journal of Ichthyology 27(1):1–8. Google Scholar

    107.

    Bannikov, A. F. 1988. A new species of stromateoid fishes (Perciformes) the Lower Oligocene of the Caucasus. Paleontological Journal 22(4):107–112. Google Scholar

    108.

    Bannikov, A. F. 1990. An Eocene veliferoid (Teleostei, Lampridiformes) from Bolca. Studi e Ricerche sui Giacimenti Terziari di Bolca 6:161–174. Google Scholar

    109.

    Bannikov, A. F. 1998. New blennioid fishes of the families Blenniidae and Clinidae (Perciformes) from the Miocene of the Caucasus and Moldova. Paleonotological Journal 32(4):385–389. Google Scholar

    110.

    Bannikov, A. F. 1999. Review of fossil Lampridiformes (Teleostei) finds with a description of a new Lophotidae genus and species from the Oligocene of the northern Caucasus. Paleontological Journal 33(1):68–76. Google Scholar

    111.

    Bannikov, A. F. 2008. Revision of the atheriniform fish genera Rhamphognathus Agassiz and Mesogaster Agassiz (Teleostei) from the Eocene of Bolca, northern Italy. Studi e Ricerche sui Giacimenti Terziari di Bolca 9:65–76. Google Scholar

    112.

    Bannikov, A. F. 2014a. A new genus of the family Palaeocentrotidae (Teleostei, Lampridiformes) from the Oligocene of the northern Caucasus and comments on other fossil Veliferoidei. Paleontological Journal 48(6):624–632. Google Scholar

    113.

    Bannikov, A. F. 2014b. The systematic composition of the Eocene actinopterygian fish fauna from Monte Bolca, northern Italy, as known to date. Studi e Ricerche sui Giacimenti Terziari di Bolca 15:23–34. Google Scholar

    114.

    Bannikov, A. F., and F. Bacchia. 2005. New species of the Cenomanian Eurypterygii (Pisces, Teleostei) from Lebanon. Paleontological Journal 39(5):514–522. Google Scholar

    115.

    Bannikov, A. F., and D. Bellwood. 2015. A new genus and species of labrid fish (Perciformes) from the Eocene of Bolca in northern Italy. Studi e Ricerche sui Giacimenti Terziari di Bolca 16(13):5–16. Google Scholar

    116.

    Bannikov, A. F.2017. Zorzinilabrus furcatus, a new genus and species of labrid fish (Perciformes) from the Eocene of Bolca in northern Italy. Studi e Ricerche sui Giacimenti Terziari di Bolca 18(15):5–14. Google Scholar

    117.

    Bannikov, A. F., and G. Carnevale. 2010. Bellwoodilabrus landinii n. gen., n. sp., a new genus and species of labrid fish (Teleostei, Perciformes) from the Eocene of Monte Bolca. Geodiversitas 32(2):201–220. Google Scholar

    118.

    Bannikov, A. F.. 2016. †Carlomonnius quasigobius gen. et sp. nov.: the first gobioid fish from the Eocene of Monte Bolca, Italy. Bulletin of Geosciences 91(1):13–22. Google Scholar

    119.

    Bannikov, A. F.. 2017. Eocene ghost pipefishes (Teleostei, Solenostomidae) from Monte Bolca, Italy. Bollettino della Società Paleontologica Italiana 56(3):319–331. Google Scholar

    120.

    Bannikov, A. F., and T. H. Fraser. 2016. A new genus and species of cardinalfish (Percomorpha, Apogonidae) from the Eocene of Bolca, northern Italy (Monte Postale site). Studi e Ricerche sui Giacimenti Terziari di Bolca 17(14):13–23. Google Scholar

    121.

    Bannikov, A. F., N. V. Parin, and J. Pinna. 1985. Rhamphexocoetus volans, gen. et sp. nov. a new beloniform fish (Beloniformes, Exocoetoidei) from the Lower Eocene of Italy. Journal of Ichthyology 25(2):150–155. Google Scholar

    122.

    Bannikov, A. F., and L. Sorbini. 1990. Eocoris bloti, a new genus and species of labrid fish (Perciformes, Labroidei) from the Eocene of Monte Bolca, Italy. Studie Ricerche sui Giacimenti Terziari di Bolca 6:133–148. Google Scholar

    123.

    Bannikov, A. F.. 2000. Preliminary note on a Lower Paleocene fish fauna from Trebiciano (Trieste–north-eastern Italy). Atti del Museo Civico di Storia Naturale di Trieste 48:15–30. Google Scholar

    124.

    Bannikov, A. F., and J. C. Tyler. 1995. Phylogenetic revision of the fish families Luvaridae and †Kushlukiidae (Acanthuroidei), with a new genus and two new species of Eocene luvarids. Washington, DC: Smithsonian Institution Press. 45 pp. (Smithsonian Contributions to Paleobiology 81.) Google Scholar

    125.

    Bannikov, A. F.. 2001. A new species of the luvarid fish genus †Avitoluvarus (Acanthuroidei: Perciformes) from the Eocene of the Caucasus in southwest Russia. Proceedings of the Biological Society of Washington 114(3):579–588. Google Scholar

    126.

    Bannikov, A. F., J. C. Tyler, D. Arcila, and G. Carnevale. 2017. A new family of gymnodont fish (Tetraodontiformes) from the earliest Eocene of the Peri-Tethys (Kabardino-Balkaria, northern Caucasus, Russia). Journal of Systematic Palaeontology 15(2):129–146. Google Scholar

    127.

    Bannikov, A., and R. Zorzin. 2019. Paralabrus rossiae, a new genus and species of putative labroid fish (Perciformes) from the Eocene of Bolca in northern Italy. Studi e Ricerche sui Giacimenti Terziari di Bolca 19:39–47. Google Scholar

    128.

    Bargelloni, L., S. Marcato, L. Zane, and T. Patarnello. 2000. Mitochondrial phylogeny of notothenioids: a molecular approach to Antarctic fish evolution and biogeography. Systematic Biology 49(1):114–129. Google Scholar

    129.

    Barros-García, D., E. Froufe, R. Bañón, J. C. Arronte, and A. De Carlos. 2018. Phylogenetic analysis shows the general diversification pattern of deep-sea notacanthiforms (Teleostei: Elopomorpha). Molecular Phylogenetics and Evolution 124:192–198. Google Scholar

    130.

    Baskin, J. N. 1973. Structure and relationships of the Trichomycteridae [dissertation]. City University of New York. pp. 389. Google Scholar

    131.

    Bean, L. 2021. Revision of the Mesozoic freshwater fish clade Archaeomaenidae. Alcheringa: An Australasian Journal of Palaeontology 45(2):217–259. Google Scholar

    132.

    Bean, L. B., and G. Arratia. 2020. Anatomical revision of the Australian teleosts Cavenderichthys talbragarensis and Waldmanichthys koonwarri impacting on previous phylogenetic interpretations of teleostean relationships. Alcheringa: An Australasian Journal of Palaeontology 44(1):121–159. Google Scholar

    133.

    Bean, T. H., and G. B. Goode. 1879. A catalogue of the fishes of Essex County, Massachusetts, including the fauna of Massachusetts Bay and the contiguous deep waters. Bulletin of the Essex Institute 11:1–38. Google Scholar

    134.

    Beaulieu, J. M., R. H. Ree, J. Cavender-Bares, G. D. Weiblen, and M. J. Donoghue. 2012. Synthesizing phylogenetic knowledge for ecological research. Ecology 93(8):S4–S13. Google Scholar

    135.

    Beckett, H. T., S. Giles, and M. Friedman. 2018. Comparative anatomy of the gill skeleton of fossil Aulopiformes (Teleostei: Eurypterygii). Journal of Systematic Palaeontology 16(14):1221–1245. Google Scholar

    136.

    Beckett, H. T., S. Giles, Z. Johanson, and M. Friedman. 2018. Morphology and phylogenetic relationships of fossil snake mackerels and cutlassfishes (Trichiuroidea) from the Eocene (Ypresian) London Clay Formation. Papers in Palaeontology 4(4):577–603. Google Scholar

    137.

    Beebe, W., and J. Crane. 1939. Deep-sea fishes of the Bermuda Oceanographic Expeditions. Family Melanostomiatidae. Zoologica 24:65–228. Google Scholar

    138.

    Begle, D. P. 1991. Relationships of the osmeroid fishes and the use of reductive characters in phylogenetic analysis. Systematic Zoology 40(1):33–53. Google Scholar

    139.

    Begle, D. P. 1992. Monophyly and relationships of the argentinoid fishes. Copeia 1992(2):350–366. Google Scholar

    140.

    Bell, M. A., J. D. Stewart, and P. J. Park. 2009. The world's oldest fossil threespine stickleback fish. Copeia 2009(2):256–265. Google Scholar

    141.

    Bellwood, D. R. 1990. A new fossil fish Phyllopharyngodon longipinnis gen. et sp. nov. (family Labridae) from the Eocene, Monte Bolca, Italy. Studi e Ricerche sui Giacimenti Terziari di Bolca 6:149–160. Google Scholar

    142.

    Bellwood, D. R., L. Van Herwerden, and N. Konow. 2004. Evolution and biogeography of marine angelfishes (Pisces: Pomacanthidae). Molecular Phylogenetics and Evolution 33(1):140–155. Google Scholar

    143.

    Belouze, A. 2002. Compréhension morphologique et phylogénétique des taxons actuels et fossiles rapportés aux Anguilliformes (“poissons,” téléostéens). Lyon: U.F.R. des Sciences de la Terre, Université Claude-Bernard. 401 pp. (Documents des Laboratoires de Géologie, Lyon 158.) Google Scholar

    144.

    Belouze, A., M. Gayet, and C. Atallah. 2003. Les premiers Anguilliformes: II. Paraphylie du genre Urenchelys Woodward, 1900 et relations phylogénétiques = The first Anguilliformes: II. Paraphyly of the genus Urenchelys Woodward, 1900 and phylogenetic relationships. Géobios 36(4):351–378. Google Scholar

    145.

    Bemis, K. E., and B. B. Collette. 2019. Family Scomberesocidae. In: B. B. Collette, K. E. Bemis, N. V. Parin, and I. B. Shakhovskoy. Order Beloniformes: Needlefishes, Sauries, Halfbeaks, and Flyingfishes. B. B. Collette and T. J. Near, eds. New Haven, CT: Peabody Museum of Natural History, Yale University. pp. 79–88. (Memoir 1, Fishes of the Western North Atlantic 10.) Google Scholar

    146.

    Bemis, W. E., E. K. Findeis, and L. Grande. 1997. An overview of Acipenseriformes. Environmental Biology of Fishes 48:25–72. Google Scholar

    147.

    Benton, M. J., P. C. J. Donoghue, R. J. Asher, M. Friedman, T. J. Near, and J. Vinther. 2015. Constraints on the timescale of animal evolutionary history. Palaeontologia Electronica 18.1.1FC:1–106.  https://doi.org/10.26879/424 Google Scholar

    148.

    Berendzen, P. B., and W. W. Dimmick. 2002. Phylogenetic relationships of pleuronectiformes based on molecular evidence. Copeia 2002(3):642–652. Google Scholar

    149.

    Berg, L. S. 1937. A classification of fish-like vertebrates. Bulletin de l'Acadèmie des Sciences de l'URSS, Classe des Sciences Mathématiques et Naturelles 4:1277–1280. Google Scholar

    150.

    Berg, L. S. 1940. Classification of fishes both Recent and fossil. Travaux de l'Institute de l'Academie des Sciences de l'URSS 5:87–517. Google Scholar

    151.

    Berra, T. M. 2003. Nurseryfsh, Kurtus gulliveri (Perciformes: Kurtidae), from northern Australia: redescription, distribution, egg mass, and comparison with K. indicus from southeast Asia. Ichthyological Exploration of Freshwaters 14(4):295–306. Google Scholar

    152.

    Berra, T. M., and J. D. Humphrey. 2002. Gross anatomy and histology of the hook and skin of forehead brooding male nurseryfish, Kurtus gulliveri, from northern Australia. Environmental Biology of Fishes 65:263–270. Google Scholar

    153.

    Bertelsen, E., J. G. Nielsen, and D. G. Smith. 1989. Suborder Saccopharyngoidei. In: E. B. Böhlke, ed. Orders Anguilliformes and Saccopharyngiformes, Volume 1. New Haven, CT: Sears Foundation for Marine Research, Yale University. pp. 636–655. (Memoir 1, Fishes of the Western North Atlantic 9.) Google Scholar

    154.

    Bertin, L., and C. Arambourg. 1958. Super-ordre des téléostéens (Teleostei). In: P. Grassé, ed. Traité de zoologie, tome 13, fascicule 3. Paris: Libraires de L'académi de médecine. pp. 2204–2500. Google Scholar

    155.

    Bertrand, S., and H. Escrivá. 2014. Chordates. In: P. Vargas and R. Zardoya, eds. The Tree of Life. Sunderland, MA: Sinauer Associates. pp. 458–468. Google Scholar

    156.

    Betancur-R., R., D. Arcila, R. P. Vari, L. C. Hughes, C. Oliveira, M. H. Sabaj, and G. Ortí. 2019. Phylogenomic incongruence, hypothesis testing, and taxonomic sampling: the monophyly of characiform fishes. Evolution 73(2):329–345. Google Scholar

    157.

    Betancur-R, R., R. E. Broughton, E. O. Wiley, K. Carpenter, J. A. López, C. Li, N. I. Holcroft, D. Arcila, M. Sanciangco, J. C. Cureton II , ET AL. 2013. The tree of life and a new classification of bony fishes. PLOS Currents 5.  https://doi.org/10.1371/currents.tol.53ba26640df0ccaee75bb165c8c26288Google Scholar

    158.

    Betancur-R, R., C. Li, T. A. Munroe, J. A. Ballesteros, and G. Ortí. 2013. Addressing gene tree discordance and non-stationarity to resolve a multi-locus phylogeny of the flatfishes (Teleostei: Pleuronectiformes). Systematic Biology 62(5):763–785. Google Scholar

    159.

    Betancur-R, R., and G. Ortí. 2014. Molecular evidence for the monophyly of flatfishes (Carangimorpharia: Pleuronectiformes). Molecular Phylogenetics and Evolution 73:18–22. Google Scholar

    160.

    Betancur-R, R., E. O. Wiley, G. Arratia, A. Acero, N. Bailly, M. Miya, G. Lecointre, and G. Ortí. 2017. Phylogenetic classification of bony fishes. BMC Evolutionary Biology 17:162.  https://doi.org/10.1186/s12862-017-0958-3 Google Scholar

    161.

    Betancur-R, R., E. O. Wiley, M. Miya, G. Lecointre, N. Bailly, and G. Ortí. 2014. New and revised classification of bony fishes. Version 2. Accessed 24 June 2014.  http://www.deepfin.org/Classification_v2.htmGoogle Scholar

    162.

    Bi, X., K. Wang, L. Yang, H. Pan, H. Jiang, Q. Wei, M. Fang, H. Yu, C. Zhu, Y. Cai, ET AL. 2021. Tracing the genetic footprints of vertebrate landing in non-teleost ray-finned fishes. Cell 184(5):1377–1391. Google Scholar

    163.

    Bian, C., Y. Hu, V. Ravi, I. S. Kuznetsova, X. Shen, X. Mu, Y. Sun, X. You, J. Li, X. Li, ET AL. 2016. The Asian arowana (Scleropages formosus) genome provides new insights into the evolution of an early lineage of teleosts. Scientific Reports 6:24501. Google Scholar

    164.

    Bieńkowska-Wasiluk, M., and N. Bonde. 2015. A new Oligocene relative of the Caproidae (Teleostei: Acanthopterygii) from the Outer Carpathians, Poland. Bulletin of Geosciences 90(2):461–478. Google Scholar

    165.

    Bieńkowska-Wasiluk, M., N. Bonde, P. R. Møller, and A. Gaździcki. 2013. Eocene relatives of cod icefishes (Perciformes: Notothenioidei) from Seymour Island, Antarctica. Geological Quarterly 57(4):567–582. Google Scholar

    166.

    Bigelow, H. B. 1963. Order Isospondyli: characters and keys to suborders and families. In: B. Bigelow and W. C. Schroeder, eds. Soft-rayed bony fishes. Class Osteichthyes. Sturgeons, Gars, Tarpon, Ladyfish, Bonefish, Salmon, Charrs, Anchovies, Herring, Shads, Smelt, Capelin, et al. New Haven, CT: Sears Foundation of Marine Research, Yale University. pp. 89–106. (Memoir 1, Fishes of the Western North Atlantic 3.) Google Scholar

    167.

    Birge, T. L., G. M. Ralph, F. Di Dario, T. A. Munroe, R. W. Bullock, S. M. Maxwell, M. D. Santos, H. Hata, and K. E. Carpenter. 2021. Global conservation status of the world's most prominent forage fishes (Teleostei: Clupeiformes). Biological Conservation 253:108903.  https://doi.org/10.1016/j.biocon.2020.108903 Google Scholar

    168.

    Bista, I., J. M. D. Wood, T. Desvignes, S. A. Mccarthy, M. Matschiner, Z. Ning, A. Tracey, J. Torrance, Y. Sims, W. Chow, ET AL. 2023. Genomics of cold adaptations in the Antarctic notothenioid fish radiation. Nature Communications 14:3412.  https://doi.org/10.1038/s41467-023-38567-6 Google Scholar

    169.

    Blainville, H. M. D. 1818. Des ichthyolites du bassin de la Méditerranée. a. Des ichthyolites de Monte-Bolca, ou Vestena-Nuova, dans le Véronais. In: Une société de naturalistes et d'agriculteurs. Nouveau dictionnaire d'histoire naturelle, appliquée aux arts, à l'agriculture, à l'économie rurale et domestique, à la médicine, etc., tome 27. Paris: Deterville. pp. 334–361. Google Scholar

    170.

    Bleeker, P. 1849. Bijdrage tot de kennis der ichthyologische fauna van het eiland Madura, met beschrijving van eenige nieuwe soorten. Verhandelingen van het Bataviaasch Genootschap van Kunsten en Wetenschappen 22(8):1–16. Google Scholar

    171.

    Bleeker, P. 1852. Bijdrage tot de kennis der ichthijologische fauna van de Moluksche Eilanden. Visschen van Amboina en Ceram. Natuurkundig Tijdschrift voor Nederlandsch Indië 3:229–309. Google Scholar

    172.

    Bleeker, P. 1853a. Derde bijdrage tot de kennis der ichthyologische fauna van Amboina. Natuurkundig Tijdschrift voor Nederlandsch Indië 4:91–130. Google Scholar

    173.

    Bleeker, P. 1853b. Nalezingen op de ichthyologie van Japan. Verhandelingen van het Bataviaasch Genootschap van Kunsten en Wetenschappen 25(7):1–56. Google Scholar

    174.

    Bleeker, P. 1855. Bijdrage tot de kennis der ichthyologische fauna van de Batoe Eilanden. Natuurkundig Tijdschrift voor Nederlandsch Indië 8:305–328. Google Scholar

    175.

    Bleeker, P. 1856. Bijdrage tot de kennis der ichthyologische fauna van het eiland Boeroe. Natuurkundig Tijdschrift voor Nederlandsch Indië 11:383–414. Google Scholar

    176.

    Bleeker, P. 1858. Siluri. Vol. 1 of Ichthyologiae Archipelagi Indici Prodromus. Batavia: Lange. 370 pp. Google Scholar

    177.

    Bleeker, P. 1859. Enumeratio specierum piscium hucusque in Archipelago indico observatarum, adjectis habitationibusque, ubi descriptiones earum recentiores reperiuntur, nec non speciebus Musei Bleekeriani Bengalensibus, Japonicis, Capensibus Tasmanicisque. Acta Societatis Scientiarum Indo-Neerlandensis 6:1–276. Google Scholar

    178.

    Bleeker, P. 1862. Conspectus generum Labroideorum analyticus. Proceedings of the Zoological Society of London 1861, pt. 3, pp. 408–418. Google Scholar

    179.

    Bleeker, P. 1864a. Atlas ichthyologique des Indes orientales néêrlandaises: publié sous les auspices du gouvernement colonial néêrlandais, tome 4. Amsterdam: Frédéric Muller, editeur. 132 pp. Google Scholar

    180.

    Bleeker, P. 1864b. Poissons inédits indo-archipélagiques de l'ordre des Murènes. Nederlandsch Tijdschrift voor de Dierkunde 2:38–54. Google Scholar

    181.

    Bleeker, P. 1864c. Systema Muraenorum revisum. Nederlandsch Tijdschrift voor de Dierkunde. 2:113–122. Google Scholar

    182.

    Bleeker, P. 1866. Systema Balistidorum, Ostracionidorum, Gymnodontidorumque revisum. Nederlandsch Tijdschrift voor de Dierkunde 3:8–19. Google Scholar

    183.

    Bleeker, P. 1874. Poissons de Madagascar et de l'île la Réunion des collections de MM. Pollen et van Dam. In: P. L. Pollen and D. C. Van Dam, eds. Recherches sur la faune de Madagascar et de ses dépendances d'après les découvertes de François. Leide [Leiden]: E. J. Brill. pp. 1–106. Google Scholar

    184.

    Bloch, M. E. 1786. Naturgeschichte der ausländischen Fische, Volume 2. Berlin: Auf Kosten des Verfassers und in Commission bei dem Buchhändler Hr. Hesse. 160 pp. Google Scholar

    185.

    Bloch, M. E. 1788. Ueber zwey merkwürdige Fischarten. Abhandlungen der Böhmischen Gesellschaft der Wissenschaften 3:278–282. Google Scholar

    186.

    Bloch, M. E. 1792a. Beschreibung zweyer neuen Fische. Schriften der Gesellschaft Naturforschender Freunde zu Berlin 10:422–424. Google Scholar

    187.

    Bloch, M. E. 1792b. Naturgeschichte der ausländischen Fische, Volume 6. Berlin: Auf Kosten des Verfassers und in Commission bei dem Buchhändler Hr. Hesse. 174 pp. Google Scholar

    188.

    Bloch, M. E. 1795. Naturgeschichte der ausländischen Fische, Volume 9. Berlin: Auf Kosten des Verfassers und in Commission bei dem Buchhändler Hr. Hesse. 192 pp. Google Scholar

    189.

    Bloch, M. E., and J. G. Schneider. 1801. Systema ichthyologiae iconibus cx illustratum. Berlin: Sumtibus auctoris impressum et Bibliopolio Sanderiano commissum. 584 pp. Google Scholar

    190.

    Bloom, D. D., and J. P. Egan. 2018. Systematics of Clupeiformes and testing for ecological limits on species richness in a trans-marine/freshwater clade. Neotropical Ichthyology 16(3):e180095. https://doi.org/10.1590/1982-0224-20180095 Google Scholar

    191.

    Bloom, D. D., and N. R. Lovejoy. 2014. The evolutionary origins of diadromy inferred from a time-calibrated phylogeny for Clupeiformes (herring and allies). Proceedings of the Royal Society B, Biological Sciences 281(1778):20132081. Google Scholar

    192.

    Bloom, D. D., P. J. Unmack, A. E. Gosztonyi, K. R. Piller, and N. R. Lovejoy. 2012. It's a family matter: molecular phylogenetics of Atheriniformes and the polyphyly of the surf silversides (family: Notocheiridae). Molecular Phylogenetics and Evolution 62(3):1025–1030. Google Scholar

    193.

    Bloom, D. D., J. T. Weir, K. R. Piller, and N. R. Lovejoy. 2013. Do freshwater fishes diversify faster than marine fishes? A test using state-dependent diversification analyses and molecular phylogenetics of new world silversides (Atherinopsidae). Evolution 67(7):2040–2057. Google Scholar

    194.

    Blot, J. 1980. La faune ichthyologique des gisements du Monte Bolca (Province de Vérone, Italie): catalogue systématique présentat l'état actuel des recherches concernant cette faune. Bulletin du Muséum National d'Histoire Naturelle 2(4):339–396. Google Scholar

    195.

    Blum, S. D. 1988. The osteology and phylogeny of the Chaetodontidae (Pisces: Perciformes) [dissertation]. University of Hawaii. 365 pp. Google Scholar

    196.

    Blum, S. D. 1991. Dastilbe Jordan, 1910. In: J. G. Maisey, ed. Santana Fossils: An Illustrated Atlas. Neptune City, NJ: T. F. H. Publications. pp. 274–283. Google Scholar

    197.

    Boeseman, M. 1964. Notes on the fishes of western New Guinea: II. Lophichthys boschmai, a new genus and species from the Arafoera Sea. Zoologische Mededelingen 39(2):12–18. Google Scholar

    198.

    Bogan, S., F. L. Agnolin, and A. Scanferla. 2018. A new Andinichthyidae catfish (Ostariophysi, Siluriformes) from the Paleogene of northwestern Argentina. Journal of Vertebrate Paleontology 38(3):e1449117.  https://doi.org/10.1080/02724634.2018.1449117 Google Scholar

    199.

    Bohlen, J., and V. Šlechtová. 2009. Phylogenetic position of the fish genus Ellopostoma (Teleostei: Cypriniformes) using molecular genetic data. Ichthyological Exploration of Freshwaters 20:157–162. Google Scholar

    200.

    Böhlke, J. E. 1966. Order Lyomeri. In: C. W. Mead, ed. Orders Iniomi and Lyomeri: Lizardfishes, Other Iniomi, Deepsea Gulpers. New Haven, CT: Sears Foundation for Marine Research, Yale University. pp. 603–628. (Memoir 1, Fishes of the Western North Atlantic 5.) Google Scholar

    201.

    Bonaparte, C. L. 1831. Saggio di una distribuzione metodica degli animali vertebrati. Giornale Arcadico di Scienze Lettere ed Arti 52:155–189. Google Scholar

    202.

    Bonaparte, C. L. 1840. Prodromus systematis ichthyologiae. Nuovi Annali delle Scienze naturali Bologna Series 1 Anno 2 4:181–196, 272–277. Google Scholar

    203.

    Bonaparte, C. L. 1845. Specchio generale del sistema ittiologico. Atti della sesta Riunione degli Scienziati Italiani 6:386–390. Google Scholar

    204.

    Bonde, N. C. 1993. Osteoglossids (Teleostei: Osteoglossomorpha) of the Mesozoic—comments on their relationships. In: G. Arratia and G. Viohl, eds. Mesozoic Fishes: Systematics and Paleoecology; proceedings of the international meeting, Eichstätt, 1993. Munich: Verlag Dr. Friedrich Pfeil. pp. 273–284. Google Scholar

    205.

    Borden, W. C., T. Grande, and W. L. Smith. 2013. Comparative osteology and myology of the caudal fin in Paracanthopterygii (Teleostei: Acanthomorpha). In: G. Arratia, H.-P. Schultze, and M. V. H. Wilson, eds. Mesozoic Fishes 5: Global Diversity and Evolution. Munich: Verlag Dr. Friedrich Pfeil. pp. 419–455. Google Scholar

    206.

    Boulenger, G. A. 1891. Pisces. Zoological Record 28:1–41. Google Scholar

    207.

    Boulenger, G. A. 1895. Catalogue of the Perciform Fishes in the British Museum. London: Taylor and Francis. 394 pp. Google Scholar

    208.

    Boulenger, G. A. 1897. Descriptions of new fishes from the upper Congo. II. Annals and Magazine of Natural History, Series 6, 20:422–427. Google Scholar

    209.

    Boulenger, G. A. 1904a. A synopsis of the suborders and families of teleostean fishes. Annals and Magazine of Natural History, Series 7, 13:161–190. Google Scholar

    210.

    Boulenger, G. A. 1904b. Teleostei. Cambridge Natural History 7: 539–727. Google Scholar

    211.

    Boulenger, G. A. 1906. Fourth contribution to the ichthyology of Lake Tanganyika.—Report on the collection of fishes made by Dr. W. A. Cunnington during the third Tanganyika Expedition, 1904–1905. Transactions of the Zoological Society of London 17:537–601. Google Scholar

    212.

    Boulenger, G. A. 1909. Catalogue of the Fresh-water Fishes of Africa in the British Museum (Natural History), Volume 1. London: Taylor and Francis. 373 pp. Google Scholar

    213.

    Bowne, P. S. 1994. Systematics and morphology of the Gasterosteiformes. In: M. A. Bell and S. A. Foster, eds. The Evolutionary Biology of the Threespine Stickleback. Oxford: Oxford University Press. pp. 28–60. Google Scholar

    214.

    Braasch, I., A. R. Gehrke, J. J. Smith, K. Kawasaki, T. Manousaki, J. Pasquier, A. Amores, T. Desvignes, P. Batzel, J. Catchen, ET AL. 2016. The spotted gar genome illuminates vertebrate evolution and facilitates human-teleost comparisons. Nature Genetics 48:427–437. Google Scholar

    215.

    Bragança, P. H. N., P. F. Amorim, and W. J. E. M. Costa. 2018. Pantanodontidae (Teleostei, Cyprinodontiformes), the sister group to all other cyprinodontoid killifishes as inferred by molecular data. Zoosystematics and Evolution 94:137–145. Google Scholar

    216.

    Bragança, P. H. N., and W. J. E. M. Costa. 2018. Time-calibrated molecular phylogeny reveals a Miocene–Pliocene diversification in the Amazon miniature killifish genus Fluviphylax (Cyprinodontiformes: Cyprinodontoidei). Organisms Diversity and Evolution 18(3):345–353. Google Scholar

    217.

    Bragança, P. H. N.2019. Multigene fossil-calibrated analysis of the African lampeyes (Cyprinodontoidei: Procatopodidae) reveals an early Oligocene origin and Neogene diversification driven by palaeogeographic and palaeoclimatic events. Organisms Diversity and Evolution 19(2):303–320. Google Scholar

    218.

    Branson, B. A., and G. A. Moore. 1962. The lateralis components of the acoustico-lateralis system in the sunfish family Centrarchidae. Copeia 1962(1):1–108. Google Scholar

    219.

    Bravo, G. A., A. Antonelli, C. D. Bacon, K. Bartoszek, M. P. K. Blom, S. Huynh, G. Jones, L. L. Knowles, S. Lamichhaney, T. Marcussen, ET AL. 2019. Embracing heterogeneity: coalescing the tree of life and the future of phylogenomics. PeerJ 7:e6399.  https://doi.org/10.7717/peerj.6399 Google Scholar

    220.

    Breder, C. M., JR ., AND D. E. Rosen. 1966. Modes of Reproduction in Fishes. Garden City, NJ: Natural History Press. 941 pp. Google Scholar

    221.

    Breining, T., and R. Britz. 2000. Egg surface structure of three clingfish species, using scanning electron microscopy. Journal of Fish Biology 56(5):1129–1137. Google Scholar

    222.

    Bridge, T. W. 1904. Fishes. Cambridge Natural History 7:141–537. Google Scholar

    223.

    Brinkman, D. B., M. G. Newbrey, and A. G. Neuman. 2014. Diversity and paleoecology of actinopterygian fish from vertebrate microfossil localities of the Maastrichtian Hell Creek Formation of Montana. In: G. P. Wilson, W. A. Clemens, J. R. Horner, and J. H. Hartman, eds. Through the End of the Cretaceous in the Type Locality of the Hell Creek Formation in Montana and Adjacent Areas. pp. 247–270. (Geological Society of America Special Papers 503.) Google Scholar

    224.

    Brito, P. M., J. Alvarado-Ortega, and F. J. Meunier. 2017. Earliest known lepisosteoid extends the range of anatomically modern gars to the Late Jurassic. Scientific Reports 7:17830. Google Scholar

    225.

    Brito, P. M., F. J. Figueiredo, and M. E. C. Leal. 2020. A revision of Laeliichthys ancestralis Santos, 1985 (Teleostei: Osteoglossomorpha) from the Lower Cretaceous of Brazil: phylogenetic relationships and biogeographical implications. PLOS ONE 15(10):e0241009.  https://doi.org/10.1371/journal.pone.0241009 Google Scholar

    226.

    Britz, R. 1994. Ontogenic features of Luciocephalus (Perciformes, Anabantoidei) with a reversed hypothesis of anabantoid intrarelationships. Zoological Journal of the Linnean Society 112(4):491–508. Google Scholar

    227.

    Britz, R. 1996. Ontogeny of the ethmoidal region and hyopalatine arch in Macrognathus pancalus (Percomorpha, Mastacembeloidei): with critical remarks on mastacembeloid inter- and intrarelationships. American Museum Novitates 3181:1–18. Google Scholar

    228.

    Britz, R. 1997. Egg surface structure and larval cement glands in nandid and badid fishes, with remarks on phylogeny and biogeography. American Museum Novitates 3195:1–17. Google Scholar

    229.

    Britz, R. 2001. The genus Betta—monophyly and intrarelationships, with remarks on the subfamilies Macropodinae and Luciocephalinae (Teleostei: Osphronemidae). Ichthyological Exploration of Freshwaters 12:305–318. Google Scholar

    230.

    Britz, R., and P. Bartsch. 2003. The myth of dorsal ribs in gnathostome vertebrates. Proceedings of the Royal Society of London Series B, Biological Sciences 270(Suppl. 1):S1–S4. Google Scholar

    231.

    Britz, R., and K. W. Conway. 2011. The Cypriniformes tree of confusion. Zootaxa 2946(1):73–78.  https://doi.org/10.11646/zootaxa.2946.1.15 Google Scholar

    232.

    Britz, R., K. W. Conway, and L. Rüber. 2014. Miniatures, morphology and molecules: Paedocypris and its phylogenetic position (Teleostei, Cypriniformes). Zoological Journal of the Linnean Society 172:556–615. Google Scholar

    233.

    Britz, R., N. Dahanukar, V. K. Anoop, S. Philip, B. Clark, R. Raghavan, and L. Rüber. 2020. Aenigmachannidae, a new family of snakehead fishes (Teleostei: Channoidei) from subterranean waters of South India. Scientific Reports 10:16081. Google Scholar

    234.

    Britz, R., and G. D. Johnson. 2002. “Paradox Lost”: skeletal ontogeny of Indostomus paradoxus and its signficance for the phylogenetic relationships of Indostomidae (Teleostei, Gasterosteiformes). American Museum Novitates 3383:1–43. Google Scholar

    235.

    Britz, R.2003. On the homology of the posteriormost gill arch in polypterids (Cladistia, Actinopterygii). Zoological Journal of the Linnean Society 138:495–503. Google Scholar

    236.

    Britz, R.2010. Occipito-vertebral fusion in actinopterygians: conjecture, myth and reality. Part 1, non-teleosts. In: J. S. Nelson, H.-P. Schultze, and M. V. H. Wilson, eds. Origin and Phylogenetic Interrelationships of Teleosts. Munich: Verlag Dr. Friedrich Pfeil. pp. 77–93. Google Scholar

    237.

    Britz, R., F. Kakkassery, and R. Raghavan. 2014. Osteology of Kryptoglanis shajii, a stygobitic catfish (Teleostei: Siluriformes) from peninsular India with diagnosis of a new family Kryptoglanidae. Ichthyological Exploration of Freshwaters 24:193–207. Google Scholar

    238.

    Britz, R., and M. Kottelat. 1999. Two new species of the gasterosteiform fishes of the genus Indostomus (Teleostei: Indostomidae). Ichthyological Exploration of Freshwaters 10:327–336. Google Scholar

    239.

    Britz, R.. 2003. Descriptive osteology of the family Chaudhuriidae (Teleostei, Synbranchiformes, Mastacembeloidei), with a discussion of its relationships. American Museum Novitates 3418:1–62. Google Scholar

    240.

    Britz, R., M. Kottelat, and T. H. Hui. 2011. Fangfangia spinicleithralis, a new genus and species of miniature cyprinid fish from the peat swamp forests of Borneo (Teleostei: Cyprinidae). Ichthyological Exploration of Freshwaters 22:327–335. Google Scholar

    241.

    Britz, R., K. Kumar, and F. Baby. 2012. Pristolepis rubripinnis, a new species of fish from southern India (Teleostei: Percomorpha: Pristolepididae). Zootaxa 3345(1):59–68.  https://doi.org/10.11646/zootaxa.3345.1.3 Google Scholar

    242.

    Brothers, E. B. 1984. Otolith studies. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 50–57. (Special Publication 1.) Google Scholar

    243.

    Broughton, R. E. 2010. Phylogeny of teleosts based on mitochondrial sequences. In: J. S. Nelson, H.-P. Schultze, and M. V. H. Wilson, eds. Origin and Phylogenetic Interrelationships of Teleosts. Munich: Verlag Dr. Friedrich Pfeil. pp. 61–76. Google Scholar

    244.

    Broughton, R. E., R. Betancur-R, C. Li, G. Arratia, and G. Ortí. 2013. Multi-locus phylogenetic analysis reveals the pattern and tempo of bony fish evolution. PLOS Currents 5.  https://doi.org/currents.tol.2ca8041495ffafd0c92756e75247483e Google Scholar

    245.

    Broussonet, P. M. A. 1782. Ichtyologia, Sistens Piscium Descriptiones et Icones. London: Elmsy. 42 pp. Google Scholar

    246.

    Brownstein, C. D. 2022. Unappreciated Cenozoic eco-morphological diversification of stem gars revealed by a new large species. Acta Palaeontologica Polonica 67(3):559–568. Google Scholar

    247.

    Brownstein, C. D. 2023. Syngnathoid evolutionary history and the conundrum of fossil misplacement. Integrative Organismal Biology 5(1):obad011.  https://doi.org/10.1093/iob/obad011 Google Scholar

    248.

    Brownstein, C. D., D. Kim, O. D. Orr, G. M. Hogue, B. H. Tracy, M. W. Pugh, R. Singer, C. Myles-Mcburney, J. M. Mollish, J. W. Simmons, ET AL. 2022. Hidden species diversity in an iconic living fossil vertebrate. Biology Letters 18(11):20220395.  https://doi.org/10.1098/rsbl.2022.0395  Google Scholar

    249.

    Brownstein, C. D., and T. R. Lyson. 2022. Giant gar from directly above the Cretaceous–Palaeogene boundary suggests healthy freshwater ecosystems existed within thousands of years of the asteroid impact. Biology Letters 18(6):20220118.  https://doi.org/10.1098/rsbl.2022.0118 Google Scholar

    250.

    Brownstein, C. D., and T. J. Near. 2023. Evolutionary origins of the lampriform pelagic radiation. Zoological Journal of the Linnean Society, zlad142.  https://doi.org/10.1093/zoolinnean/zlad142  Google Scholar

    251.

    Brownstein, C. D., L. Yang, M. Friedman, and T. J. Near. 2023. Phylogenomics of the ancient and speciesdepauperate gars tracks 150 million years of continental fragmentation in the Northern Hemisphere. Systematic Biology 72(1):213–227. Google Scholar

    252.

    Brünnich, M. T. 1768. Ichthyologia Massiliensis, sistens piscium descriptiones eorumque apud incolas nomina. Accedunt Spolia Maris Adriatici. Hafniae et Lipsiae [Copenhagen and Leipzig]: Apud Rothii viduam et Proft. 110 pp. Google Scholar

    253.

    Brünnich, M. T. 1788. Om en ny fiskart, den draabeplettede pladefish, fanget ved Helsingör i Nordsöen 1786. Kongelige Danske Videnskabernes Selskabs Skrivter, Nye Samling af det 3:398–407. Google Scholar

    254.

    Buckup, P. A. 1993. The monophyly of the Characidiinae, a Neotropical group of characiform fishes (Teleostei: Ostariophysi). Zoological Journal of the Linnean Society 108:225–245. Google Scholar

    255.

    Buckup, P. A. 1998. Relationships of the Characidiinae and phylogeny of characiform fishes (Teleostei: Ostariophysi). In: L. R. Malabarba, R. E. Reis, R. P. Vari, Z. M. Lucena, and C. A. S. Lucena, eds. Phylogeny and Classification of Neotropical Fishes. Porto Alegre, Brazil: EDIPUCRS. pp. 123–144. Google Scholar

    256.

    Burgess, G. H. 1974. Evidence for the elevation to family status of the angelfishes (Pomacanthidae), previously considered to be a subfamily of the butterflyfish family, Chaetodontidae. Pacific Science 28(1):57–71. Google Scholar

    257.

    Bürgin, T. 1989. The oral jaw apparatus of the Indian Halibut, Psettodes erumei (Bloch and Schneider, 1801) (Teleostei; Pleuronectiformes)—a formal description with functional considerations. Progress in Zoology 35:463–466. Google Scholar

    258.

    Burns, M. D., and B. L. Sidlauskas. 2019. Ancient and contingent body shape diversification in a hyperdiverse continental fish radiation. Evolution 73(3):569–587. Google Scholar

    259.

    Burridge, C. P., R. M. Mcdowall, D. Craw, M. V. H. Wilson, and J. M. Waters. 2012. Marine dispersal as a pre-requisite for Gondwanan vicariance among elements of the galaxiid fish fauna. Journal of Biogeography 39(2):306–321. Google Scholar

    260.

    Burridge, C. P., and A. J. Smolenski. 2004. Molecular phylogeny of the Cheilodactylidae and Latridae (Perciformes: Cirrhitoidea) with notes on taxonomy and biogeography. Molecular Phylogenetics and Evolution 30(1):118–127. Google Scholar

    261.

    Byrne, L., F. Chapleau, and S. Aris-Brosou. 2018. How the Central American Seaway and an ancient Northern Passage affected flatfish diversification. Molecular Biology and Evolution 35(8):1982–1989. Google Scholar

    262.

    Cadenat, J. 1937. Recherches systématiques sur les poissons littoraux de la côte occidentale d'Afrique. Liste des poissons littoraux récoltés par le navire “Président Théodore-Tissier” au cours de sa cinquième croisière (1936). Revue des Travaux de l'Institut des Pêches Maritimes 10(fasc. 4, no. 40):425–462. Google Scholar

    263.

    Calcagnotto, D., S. A. Schaefer, and R. Desalle. 2005. Relationships among characiform fishes inferred from analysis of nuclear and mitochondrial gene sequences. Molecular Phylogenetics and Evolution 36(1):135–153. Google Scholar

    264.

    Calzoni, P., J. Amalfitano, L. Giusberti, G. Marramà, and G. Carnevale. 2023. Eocene Rhamphosidae (Teleostei: Syngnathiformes) from the Bolca Lagerstätte, Italy. Rivista Italiana di Paleontologia e Stratigrafia 129(3):573–607. Google Scholar

    265.

    Campanella, D., L. C. Hughes, P. J. Unmack, D. D. Bloom, K. R. Piller, and G. Ortí. 2015. Multi-locus fossil-calibrated phylogeny of Atheriniformes (Teleostei, Ovalentaria). Molecular Phylogenetics and Evolution 86:8–23. Google Scholar

    266.

    Campbell, M. A., M. E. Alfaro, M. Belasco, and J. A. López. 2017. Early-branching euteleost relationships: areas of congruence between concatenation and coalescent model inferences. PeerJ 5:e3548.  https://doi.org/10.7717/peerj.3548 Google Scholar

    267.

    Campbell, M. A., B. Chanet, J.-N. Chen, M.-Y. Lee, and W.-J. Chen. 2019. Origins and relationships of the Pleuronectoidei: molecular and morphological analysis of living and fossil taxa. Zoologica Scripta 48(5):640–656. Google Scholar

    268.

    Campbell, M. A., W.-J. Chen, and J. A. López. 2013. Are flatfishes (Pleuronectiformes) monophyletic? Molecular Phylogenetics and Evolution 69(3):664–673. Google Scholar

    269.

    Campbell, M. A.. 2014. Molecular data do not provide unambiguous support for the monophyly of flatfishes (Pleuronectiformes): a reply to Betancur-R and Ortí. Molecular Phylogenetics and Evolution 75:149–153. Google Scholar

    270.

    Campbell, M. A., and J. A. Lopéz. 2014. Mitochondrial phylogeography of a Beringian relict: the endemic freshwater genus of blackfish Dallia (Esociformes). Journal of Fish Biology 84(2):523–538. Google Scholar

    271.

    Campbell, M. A., J. A. López, T. Sado, and M. Miya. 2013. Pike and salmon as sister taxa: detailed intraclade resolution and divergence time estimation of Esociformes + Salmoniformes based on whole mitochondrial genome sequences. Gene 530(1):57–65. Google Scholar

    272.

    Campbell, M. A., J. G. Nielsen, T. Sado, C. Shinzato, M. Kanda, T. P. Satoh, and M. Miya. 2017. Evolutionary affinities of the unfathomable Parabrotulidae: molecular data indicate placement of Parabrotula within the family Bythitidae, Ophidiiformes. Molecular Phylogenetics and Evolution 109:337–342. Google Scholar

    273.

    Campbell, M. A., T. Sado, C. Shinzato, R. Koyanagi, M. Okamoto, and M. Miya. 2018. Multilocus phylogenetic analysis of the first molecular data from the rare and monotypic Amarsipidae places the family within the Pelagia and highlights limitations of existing data sets in resolving pelagian interrelationships. Molecular Phylogenetics and Evolution 124:172–180. Google Scholar

    274.

    Campbell, M. A., P. Tongboonkua, B. Chanet, and W.-J. Chen. 2020. The distribution of the recessus orbitalis across flatfishes (order: Pleuronectiformes). Journal of Fish Biology 97(1):293–297. Google Scholar

    275.

    Cantalice, K. M., and J. Alvarado-Ortega. 2016. Eekaulostomus cuevasae gen. and sp. nov., an ancient armored trumpetfish (Aulostomoidea) from Danian (Paleocene) marine deposits of Belisario Domínguez, Chiapas, southeastern Mexico. Palaeontologia Electronica 19.3.53A:1–24. Google Scholar

    276.

    Cantalice, K. M., J. Alvarado-Ortega, and D. R. Bellwood. 2020. †Chaychanus gonzalezorum gen. et sp. nov.: a damselfish fossil (Percomorphaceae; Pomacentridae), from the Early Paleocene outcrop of Chiapas, southeastern Mexico. Journal of South American Earth Sciences 98:102322.  HTTPS://DOI.ORG/10.1016/J.JSAMES.2019.102322  Google Scholar

    277.

    Cantalice, K. M., J. Alvarado-Ortega, D. R. Bellwood, and A. C. Siqueira. 2022. Rising from the ashes: the biogeographic origins of modern coral reef fishes. BioScience 72(8):769–777. Google Scholar

    278.

    Cantalice, K. M., B. A. Than-Marchese, and E. Villalobos-Segura. 2021. A new Cenomanian acanthomorph fish from the El Chango quarry (Chiapas, south-eastern Mexico) and its implications for the early diversification and evolutionary trends of acanthopterygians. Papers in Palaeontology 7(3):1699–1726. Google Scholar

    279.

    Cantino, P. D., and K. De Queiroz. 2020. International Code of Phylogenetic Nomenclature (PhyloCode). Boca Raton: CRC Press. 149 pp. Google Scholar

    280.

    Cantor, T. 1842. General features of Chusan, with remarks on the flora and fauna of that island. Annals and Magazine of Natural History, Series 1, 9(60):481–493. Google Scholar

    281.

    Capobianco, A., and M. Friedman. 2019. Vicariance and dispersal in southern hemisphere freshwater fish clades: a palaeontological perspective. Biological Reviews 94(2):662–699. Google Scholar

    282.

    Carmichael, D. 1819. Description of four species of fish found on the coast of Tristan da Cunha. Transactions of the Linnean Society of London 12(2):500–502. Google Scholar

    283.

    Carnevale, G. 2007. Fossil fishes from the Serravallian (Middle Miocene) of Torricella Peligna, Italy. Palaeontographia Italica 91:1–67. Google Scholar

    284.

    Carnevale, G., and A. F. Bannikov. 2019. A dragonet (Teleostei, Callionymoidei) from the Eocene of Monte Bolca, Italy. Bollettino della Società Paleontologica Italiana 58(3):295–307. Google Scholar

    285.

    Carnevale, G., A. F. Bannikov, G. Marramà, J. C. Tyler, and R. Zorzin. 2014. 5. The Pesciara-Monte Postale Fossil Lagerstätte: 2. Fishes and other vertebrates. In: C. A. Papazzoni, L. Giusberti, G. Carnevale, G. Roghi, D. Bassi, and R. Zorzin, eds. The Bolca Fossil-Lagerstätten: A Window into the Eocene World. Milan: Rendiconti della Società Paleontologica Italiana. pp. 37–63. Google Scholar

    286.

    Carnevale, G., and B. B. Collette. 2014. †Zappaichthys harzhauseri, gen. et sp. nov., a new Miocene toadfish (Teleostei, Batrachoidiformes) from the Paratethys (St. Margarethen in Burgenland, Austria), with comments on the fossil record of batrachoidiform fishes. Journal of Vertebrate Paleontology 34(5):1005–1017. Google Scholar

    287.

    Carnevale, G., and G. D. Johnson. 2015. A Cretaceous cusk-eel (Teleostei, Ophidiiformes) from Italy and the Mesozoic diversification of percomorph fishes. Copeia 103(4):771–791. Google Scholar

    288.

    Carnevale, G., L. Pellegrino, and J. C. Tyler. 2021. Evolution and fossil record of the ocean sunfishes. In: T. M. Thys, G. C. Hays, and J. D. R. Houghton, eds. The Ocean Sunfishes. Boca Raton, FL: CRC. pp. 1–17. Google Scholar

    289.

    Carnevale, G., and T. W. Pietsch. 2009. An Eocene frogfish from Monte Bolca, Italy: the earliest known skeletal record for the family. Palaeontology 52(4):745–752. Google Scholar

    290.

    Carnevale, G.. 2010. Eocene handfishes from Monte Bolca, with description of a new genus and species, and a phylogeny of the family Brachionichthyidae (Teleostei: Lophiiformes). Zoological Journal of the Linnean Society 160:621–647. Google Scholar

    291.

    Carnevale, G.. 2011. Batfishes from the Eocene of Monte Bolca. Geological Magazine 148(3):461–472. Google Scholar

    292.

    Carnevale, G.. 2012. †Caruso, a new genus of anglerfishes from the Eocene of Monte Bolca, Italy, with a comparative osteology and phylogeny of the teleost family Lophiidae. Journal of Systematic Palaeontology 10(1):47–72. Google Scholar

    293.

    Carnevale, G., T. W. Pietsch, N. Bonde, M. E. C. Leal, and G. Marramà. 2020. †Neilpeartia ceratoi, gen. et sp. nov., a new frogfish from the Eocene of Bolca, Italy. Journal of Vertebrate Paleontology 40(2):e1778711.  https://doi.org/10.1080/02724634.2020.1778711  Google Scholar

    294.

    Carnevale, G., and A. Rindone. 2011. The teleost fish Paravinciguerria praecursor Arambourg, 1954 in the Cenomanian of north-eastern Sicily. Bollettino della Società Paleontologica Italiana 50(1):1–10. Google Scholar

    295.

    Caron, A., V. Venkataraman, K. Tietjen, and M. I. Coates. 2023. A fish for Phoebe: a new actinopterygian from the Upper Carboniferous Coal Measures of Saddleworth, Greater Manchester, UK, and a revision of Kansasiella eatoni. Zoological Journal of the Linnean Society 198:957–981. Google Scholar

    296.

    Carpenter, K. E. 1988. Fusilier Fishes of the World. Vol. 8 of FAO Species Catalogue. Rome: Food and Agriculture Organization of the United Nations. 75 pp. (FAO Fisheries Synopsis 125.) Google Scholar

    297.

    Carpenter, K. E., and G. R. Allen. 1989. Emperor Fishes and Large-eye Breams of the World (Family Lethrinidae). Vol. 8 of FAO Species Catalogue. 118 pp. (FAO Fisheries Synopsis 125.) Google Scholar

    298.

    Carvalho, M., F. A. Bockmann, and M. R. De Carvalho. 2013. Homology of the fifth epibranchial and accessory elements of the ceratobranchials among gnathostomes: insights from the development of Ostariophysans. PLOS ONE 8(4):e62389.  https://doi.org/10.1371/journal.pone.0062389 Google Scholar

    299.

    Casier, E. 1946. La faune ichthyologique de l'Ypresien de la Belgique. Brussells: Musée Royal d'Histoire Naturelle de Belgique. 267 pp. (Memoires 104.) Google Scholar

    300.

    Casier, E. 1966. Faune Ichthyologique du London Clay. London: British Museum (Natural History). 496 pp. Google Scholar

    301.

    Castle, P. H. J. 1977. Results of the research cruises of FRV “Walther Herwig” to South America. L. A new genus and species of bobtail eel (Anguilliformes, Cyemidae) from the South Atlantic. Archiv für Fischereiwissenschaft 28(2–3):69–79. Google Scholar

    302.

    Cavagnetto, C., and J. Gaudant. 2000. A palynoflora of Palaeocene age from the fossiliferous sapropels of the Boltyshka depression, Central Ukraine. Newsletters on Stratigraphy 38(1):39–56. Google Scholar

    303.

    Cavender, T. 1969. An Oligocene Mudminnow (Family Umbridae) from Oregon with Remarks on Relationships within the Esocoidei. Ann Arbor: University of Michigan Museum of Zoology. 33 pp. (Occasional Papers 660.) Google Scholar

    304.

    Cavender, T. M., and M. M. Coburn. 1992. Phylogenetic relationships of North American Cyprinidae. In: R. L. Mayden, ed. Systematics, Historical Ecology, and North American Freshwater Fishes. Stanford, CA: Stanford University Press. pp. 293–327. Google Scholar

    305.

    Cavin, L. 1999. A new Clupavidae (Teleostei, Ostariophysi) from the Cenomanian of Daoura (Morocco). Comptes rendus de l'Académie des sciences Série 2A: sciences de la terre et des planètes 329:689–695. Google Scholar

    306.

    Cavin, L. 2001. Osteology and phylogenetic relationships of the teleost Goulmimichthys arambourgi Cavin, 1995, from the Upper Cretaceous of Goulmima, Morocco. Eclogae Geologicae Helvetiae 94:509–535. Google Scholar

    307.

    Cavin, L. 2017. Freshwater Fishes: 250 Years of Evolutionary History. London: ISTE Press. 199 pp. Google Scholar

    308.

    Cavin, L., A. Alexopoulos, and A. Piuz. 2012. Late Cretaceous (Maastrichtian) ray-finned fishes from the island of Gavdos, southern Greece, with comments on the evolutionary history of the aulopiform teleost Enchodus. Bulletin de la Société Géologique de France 183(6):561–572. Google Scholar

    309.

    Cavin, L., U. Deesri, and V. Suteethorn. 2013. Osteology and relationships of Thaiichthys nov. gen.: a Ginglymodi from the Late Jurassic–Early Cretaceous of Thailand. Palaeontology 56(1):183–208. Google Scholar

    310.

    Cavin, L., and P. L. Forey. 2001. Osteology and systematic affinities of Palaeonotopterus greenwoodi Forey 1997 (Teleostei: Osteoglossomorpha). Zoological Journal of the Linnean Society 133:25–52. Google Scholar

    311.

    Cavin, L., and V. Suteethorn. 2006. A new semionotiform (Actinopterygii, Neopterygii) from Upper Jurassic–Lower Cretaceous deposits of north-east Thailand, with comments on the relationships of semionotiforms. Palaeontology 49(2):339–353. Google Scholar

    312.

    Cellinese, N., and C. Dell. 2020. RegNum—the international clade names repository [online database]. Accessed 2 December 2023.  https://www.phyloregnum.orgGoogle Scholar

    313.

    Chakrabarty, P. 2010. The transitioning state of systematic ichthyology. Copeia 2010(3):513–515. Google Scholar

    314.

    Chakrabarty, P., M. P. Davis, and J. S. Sparks. 2012. The first record of a trans-oceanic sister-group relationship between obligate vertebrate troglobites. PLOS ONE 7(8):e44083.  https://doi.org/10.1371/journal.pone.0044083 Google Scholar

    315.

    Chakrabarty, P., B. C. Faircloth, F. Alda, W. B. Ludt, C. D. Mcmahan, T. J. Near, A. Dornburg, J. S. Albert, J. Arroyave, ET AL. 2017. Phylogenomic systematics of ostariophysan fishes: ultraconserved elements support the surprising non-monophyly of Characiformes. Systematic Biology 66(6):881–895. Google Scholar

    316.

    Chakrabarty, P., J. A. Prejean, and M. L. Niemiller. 2014. The Hoosier cavefish, a new and endangered species (Amblyopsidae, Amblyopsis) from the caves of southern Indiana. ZooKeys 412:41–57. Google Scholar

    317.

    Chanet, B. 1994. Eobuglossus eocenicus (Woodward 1910) from the Upper Lutetian of Egypt, one of the oldest soleids (Teleostei, Pleuronectiformes). Neues Jahrbuch für Geologie und Paläontologie, Monatshefte 1994:391–398. Google Scholar

    318.

    Chanet, B. 1995. Joleaudichthys sadeki Chabanaud; 1937: Pleuronectiform fish from Egyptian Eocene. First steps on the history of Pleuronectiformes (Pisces; Teleostei). Géobios 28(Suppl. 2):189–191. Google Scholar

    319.

    Chanet, B. 1997. A cladistic reappraisal of the fossil flatfishes record consequences on the phylogeny of the Pleuronectiformes (Osteichthyes: Teleostei). Annales des Sciences naturelles—Zoologie et Biologie Animale, 13e Série, 18(3):105–117. Google Scholar

    320.

    Chanet, B. 1999. Supposed and true flatfishes [Teleostei: Pleuronectiformes] from the Eocene of Monte Bolca, Italy. Studi e Ricerche sui Giacimenti Terziari di Bolca 8:220–243. Google Scholar

    321.

    Chanet, B. 2003. Interrelationships of scophthalmid fishes (Pleuronectiformes: Scophthalmidae). Cybium 27(4):275–286. Google Scholar

    322.

    Chanet, B., C. Guintard, E. Betti, C. Gallut, A. Dettaï, and G. Lecointre. 2013. Evidence for a close phylogenetic relationship between the teleost orders Tetraodontiformes and Lophiiformes based on an analysis of soft anatomy. Cybium 37(3):179–198. Google Scholar

    323.

    Chanet, B., J. Mondéjar-Fernández, and G. Lecointre. 2020. Flatfishes interrelationships revisited based on anatomical characters. Cybium 44(1):9–18. Google Scholar

    324.

    Chang, J., D. L. Rabosky, S. A. Smith, and M. E. Alfaro. 2019. An R package and online resource for macroevolutionary studies using the ray-finned fish tree of life. Methods in Ecology and Evolution 10(7):1118–1124. Google Scholar

    325.

    Chang, M.-M., and G. Chen. 2008. Fossil cypriniformes from China and its adjacent areas and their paleobiogeographical implications. In: L. Cavin, A. Longbottom, and M. Richter, eds. Fishes and the Break-up of Pangaea. London: Geological Society of London. pp. 337–350. (Special Publications 295.) Google Scholar

    326.

    Chang, M.-M., and J. G. Maisey. 2003. Redescription of †Ellimma branneri and †Diplomystus shengliensis, and relationships of some basal clupeomorphs. American Museum Novitates 3404:1–35. Google Scholar

    327.

    Chapleau, F. 1993. Pleuronectiform relationships: a cladistic reassessment. Bulletin of Marine Science 52:516–540. Google Scholar

    328.

    Chen, F., G. Xue, Y. Wang, H. Zhang, P. D. Clift, Y. Xing, J. He, J. S. Albert, J. Chen, and P. Xie. 2023. Evolution of the Yangtze River and its biodiversity. The Innovation 4(3):100417.  https://doi.org/10.1016/j.xinn.2023.100417 Google Scholar

    329.

    Chen, G.-J., M.-M. Chang, and Q. Wang. 2010. Redescription of †Cobitis longipectoralis Zhou, 1992 (Cypriniformes: Cobitidae) from late early Miocene of East China. Science China Earth Sciences 53:945–955. Google Scholar

    330.

    Chen, G.-J., W. Liao, and X.-Q. Lei. 2015. First fossil cobitid (Teleostei: Cypriniformes) from early-middle Oligocene deposits of South China. Vertebrata PalAsiatica 53(4):299–309. Google Scholar

    331.

    Chen, J. N., J. A. López, S. Layoue, M. Miya, and W.-J. Chen. 2014. Phylogeny of the Elopomorpha (Teleostei): evidence from six nuclear and mitochondrial markers. Molecular Phylogenetics and Evolution 70:152–161. Google Scholar

    332.

    Chen, M.-Y., D. Liang, and P. Zhang. 2015. Selecting question-specific genes to reduce incongruence in phylogenomics: a case study of jawed vertebrate backbone phylogeny. Systematic Biology 64(6):1104–1120. Google Scholar

    333.

    Chen, W.-J., C. Bonillo, and G. Lecointre. 2003. Repeatability of clades as a criterion of reliability: a case study for molecular phylogeny of Acanthomorpha (Teleostei) with larger number of taxa. Molecular Phylogenetics and Evolution 26(2):262–288. Google Scholar

    334.

    Chen, W.-J., S. Lavoué, L. B. Beheregaray, and R. L. Mayden. 2014. Historical biogeography of a new anti-tropical clade of temperate freshwater fishes. Journal of Biogeography 41(9):1806–1818. Google Scholar

    335.

    Chen, W.-J., S. Lavoué, and R. L. Mayden. 2013. Evolutionary origin and early biogeography of otophysan fishes (Ostariophysi: Teleostei). Evolution 67(8):2218–2239. Google Scholar

    336.

    Chen, W.-J., V. Lheknim, and R. L. Mayden. 2009. Molecular phylogeny of the Cobitoidea (Teleostei: Cypriniformes) revisited: position of enigmatic loach Ellopostoma resolved with six nuclear genes. Journal of Fish Biology 75(9):2197–2208. Google Scholar

    337.

    Chen, W.-J., and R. L. Mayden. 2009. Molecular systematics of the Cyprinoidea (Teleostei: Cypriniformes), the world's largest clade of freshwater fishes: further evidence from six nuclear genes. Molecular Phylogenetics and Evolution 52(2):544–549. Google Scholar

    338.

    Chen, W.-J., M. Miya, K. Saitoh, and R. L. Mayden. 2008. Phylogenetic utility of two existing and four novel nuclear gene loci in reconstructing tree of life of ray-finned fishes: the order Cypriniformes (Ostariophysi) as a case study. Gene 423(2):125–134. Google Scholar

    339.

    Chen, W.-J., F. Santini, G. Carnevale, J. N. Chen, S. H. Liu, S. Lavoue, and R. L. Mayden. 2014. New insights on early evolution of spiny-rayed fishes (Teleostei: Acanthomorpha). Frontiers in Marine Science 1:53.  https://doi.org/10.3389/fmars.2014.00053 Google Scholar

    340.

    Chen, X. Y., and G. Arratia. 1994. Olfactory organ of acipenseriformes and comparison with other actinopterygians: patterns of diversity. Journal of Morphology 222(3):241–267. Google Scholar

    341.

    Chereshnev, I. A., O. A. Radchenko, and A. V. Petrovskaya. 2013. Relationships and position of the taxa of the subfamily Xiphisterinae in the system of the suborder Zoarcoidei (Perciformes). Russian Journal of Marine Biology 39:276–286. Google Scholar

    342.

    Cione, A. L. 1987. The late Cretaceous fauna of Los Alamitos, Patagonia, Argentina, Part II: the fishes. Revista del Museo Argentino de ciencias naturales “Bernardino Rivadavia” e Instituto Nacional de Investigacion de als Ciencas Naturales: Paleontologia 3:111–120. Google Scholar

    343.

    Cione, A. L., M. D. L. M. Azpelicueta, and D. R. Bellwood. 1995. An oplegnathid fish from the Eocene of Antarctica. Palaeontology 37(4):931–940. Google Scholar

    344.

    Clardy, T. R. 2014. Phylogenetic systematics of the prickleback family Stichaeidae (Cottiformes: Zoarcidae) using morphological data [dissertation]. College of William and Mary. 188 pp. Dissertations, Theses, and Masters Projects, Paper 1539616612.  https://doi.org/10.25773/v5-fyer-5n47  Google Scholar

    345.

    Clark, E., and M. Pohle. 1996. Trichonotus halstead, a new sand-diving fish from Papua New Guinea. Environmental Biology of Fishes 45:1–11. Google Scholar

    346.

    Clark, H. W. 1937. New fishes from the Templeton Crocker expedition of 1934–35. Copeia 1937(2):88–91. Google Scholar

    347.

    Clausen, H. S. 1959. Denticipitidae, a new family of primitive isospondylous teleosts from West African fresh-water. Videnskabelige Meddelelser fra Dansk Naturhistorisk Forening 121:141–156. Google Scholar

    348.

    Close, R. A., Z. Johanson, J. C. Tyler, R. C. Harrington, and M. Friedman. 2016. Mosaicism in a new Eocene pufferfish highlights rapid morphological innovation near the origin of crown tetraodontiforms. Palaeontology 59(4):499–514. Google Scholar

    349.

    Coates, M. I. 1998. Actinopterygians from the Namurian of Bearsden, Scotland, with comments on early actinopterygian neurocrania. Zoological Journal of the Linnean Society 122:27–59. Google Scholar

    350.

    Coates, M. I. 1999. Endocranial preservation of a Carboniferous actinopterygian from Lancashire, UK, and the interrelationships of primitive actinopterygians. Philosophical Transactions of the Royal Society of London Series B, Biological Sciences 354(1382):435–462. Google Scholar

    351.

    Coelho, M. V., C. Cupello, and P. Brito. 2018. Morphological variations in the dorsal fin finlets of extant polypterids raise questions about their taxonomical validity. PeerJ 6:e5083.  https://doi.org/10.7717/peerj.5083 Google Scholar

    352.

    Cohen, D. M. 1984. Gadiformes: overview. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 259–265. (Special Publication 1.) Google Scholar

    353.

    Cohen, D. M., T. Inada, T. Iwamoto, and N. Scialabba. 1990. Gadiform Fishes of the World (Order Gadiformes). Vol. 10 of FAO Species Catalogue. 442 pp. (FAO Fisheries Synopsis 125.) Google Scholar

    354.

    Cohen, D. M., and J. G. Nielsen. 1978. Guide to the Identififcation of Genera of the Fish Order Ophidiiformes with a Tentative Classification of the Order. Washington, DC: National Oceanic and Atmospheric Administration. 72 pp. (NOAA Technical Report NMFS Circular 417.) Google Scholar

    355.

    Colgan, D. J., C. G. Zhang, and J. R. Paxton. 2000. Phylogenetic investigations of the Stephanoberyciformes and Beryciformes, particularly whalefishes (Euteleostei: Cetomimidae), based on partial 12S rDNA and 16S rDNA sequences. Molecular Phylogenetics and Evolution 17(1):15–25. Google Scholar

    356.

    Collar, D. C., S. Tremaine, R. C. Harrington, H. T. Beckett, and M. Friedman. 2022. Mosaic adaptive peak shifts underlie body shape diversification in pelagiarian fishes (Acanthomorpha: Percomorpha). Biological Journal of the Linnean Society 137(2):324–340. Google Scholar

    357.

    Collette, B. B. 2003. Family Belonidae Bonaparte 1832: needlefishes. California Academy of Sciences Annotated Checklists of Fishes 16:1–22. Google Scholar

    358.

    Collette, B. B. 2004a. Family Hemiramphidae Gill 1859: halfbeaks. California Academy of Sciences Annotated Checklists of Fishes 22:1–35. Google Scholar

    359.

    Collette, B. B. 2004b. Family Scomberesocidae Müller 1843: sauries. California Academy of Sciences Annotated Checklists of Fishes 21:1–6. Google Scholar

    360.

    Collette, B. B., and K. E. Bemis. 2019a. Family Belonidae. In: B. B. Collette, K. E. Bemis, N. V. Parin, and I. B. Shakhovskoy. Order Beloniformes: Needlefishes, Sauries, Halfbeaks, and Flyingfishes. B. B. Collette and T. J. Near, eds. New Haven, CT: Peabody Museum of Natural History, Yale University. pp. 5–77. (Memoir 1, Fishes of the Western North Atlantic 10.) Google Scholar

    361.

    Collette, B. B.2019b. Family Hemiramphidae. In: B. B. Collette, K. E. Bemis, N. V. Parin, and I. B. Shakhovskoy. Order Beloniformes: Needlefishes, Sauries, Halfbeaks, and Flyingfishes. B. B. Collette and T. J. Near, eds. New Haven, CT: Peabody Museum of Natural History, Yale University. pp. 89–147. (Memoir 1, Fishes of the Western North Atlantic 10.) Google Scholar

    362.

    Collette, B. B.2019c. Order Beloniformes: Needlefishes, Sauries, Halfbeaks, and Flyingfishes. In: B. B. Collette, K. E. Bemis, N. V. Parin, and I. B. Shakhovskoy. Order Beloniformes: Needlefishes, Sauries, Halfbeaks, and Flyingfishes. B. B. Collette and T. J. Near, eds. New Haven, CT: Peabody Museum of Natural History, Yale University. pp. 1–4. (Memoir 1, Fishes of the Western North Atlantic 10.) Google Scholar

    363.

    Collette, B. B., G. E. Mcgowen, N. V. Parin, and S. Mito. 1984. Beloniformes: development and relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 335–354. (Special Publication 1.) Google Scholar

    364.

    Collette, B. B., and C. E. Nauen. 1983. Scombrids of the World. Vol. 2 of FAO Species Catalogue. 137 pp. (FAO Fisheries Synopsis 125.) Google Scholar

    365.

    Collette, B. B., T. Potthoff, W. J. Richards, S. Ueyanagi, J. L. Russo, and Y. Nishikawa. 1984. Scombridei: development and relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 591–620. (Special Publication 1.) Google Scholar

    366.

    Collins, R. A., R. Britz, and L. Rüber. 2015. Phylogenetic systematics of leaffishes (Teleostei: Polycentridae, Nandidae). Journal of Zoological Systematics and Evolutionary Research 53(4):259–272. Google Scholar

    367.

    Conway, K. W. 2011. Osteology of the South Asian genus Psilorhynchus McClelland, 1839 (Teleostei: Ostariophysi: Psilorhynchidae), with investigation of its phylogenetic relationships within the order Cypriniformes. Zoological Journal of the Linnean Society 163(1):50–154. Google Scholar

    368.

    Conway, K. W., M. V. Hirt, L. Yang, R. L. Mayden, and A. M. Simons. 2010. Cypriniformes: systematics and paleontology. In: J. S. Nelson, H.-P. Schultze, and M. V. H. Wilson, eds. Origin and Phylogenetic Interrelationships of Teleosts. Munich: Verlag Dr. Friedrich Pfeil. pp. 295–316. Google Scholar

    369.

    Conway, K. W., and R. L. Mayden. 2007. The gill arches of Psilorhynchus (Ostariophysi: Psilorhynchidae). Copeia 2007(2):267–280. Google Scholar

    370.

    Cooper, J. A., and F. Chapleau. 1998a. Monophyly and intrarelationships of the family Pleuronectidae (Pleuronectiformes), with a revised classification. Fishery Bulletin 96:686–726. Google Scholar

    371.

    Cooper, J. A.. 1998b. Phylogenetic status of Paralichthodes algoensis (Pleuronectiformes: Paralichthodidae). Copeia 1998(2):477–481. Google Scholar

    372.

    Cope, E. D. 1865. Partial catalogue of the cold-blooded Vertebrata of Michigan. Part II. Proceedings of the Academy of Natural Science Philadelphia 17(2):78–88. Google Scholar

    373.

    Cope, E. D. 1871a. Contribution to the ichthyology of the Lesser Antilles. Transactions of the American Philosophical Society 14:445–483. Google Scholar

    374.

    Cope, E. D. 1871b. Observations on the systematic relations of the fishes. American Naturalist 5(8–9):579–593. Google Scholar

    375.

    Cope, E. D. 1871c. On the fishes of the Tertiary shales of Green River, Wyoming Territory. In: F. V. Hayden, ed. Preliminary Report of the United States Geological Survey of Wyoming and Portions of Contiguous Territories. Washington, DC: Government Printing Office. pp. 425–431. Google Scholar

    376.

    Cope, E. D. 1873. A contribution to the ichthyology of Alaska. Proceedings of the American Philosophical Society 13(90):24–32. Google Scholar

    377.

    Cope, E. D. 1877a. A contribution to the knowledge of the ichthyological fauna of the Green River Shales. In: F. V. Hayden, ed. Bulletin of the United States Geological and Geographical Survey of the Territories, Volume 3. Washington, DC: Government Printing Office. pp. 807–819. Google Scholar

    378.

    Cope, E. D. 1877b. On the classification of the extinct fishes of the lower types. Proceedings of the American Association for the Advancement of Science 26:292–300. Google Scholar

    379.

    Cope, E. D. 1878. Descriptions of fishes from the cretaceous and tertiary deposits west of the Mississippi River. Bulletin of the United States Geological and Geographical Survey of the Territories, Volume 4. Washington, DC: Government Printing Office. pp. 67–77. Google Scholar

    380.

    Cope, E. D. 1883. On a new and extinct genus and species of Percidae from Dakota Territory. American Journal of Science 25:414–416. Google Scholar

    381.

    Costa, W. J. E. M. 1998. Phylogeny and classification of the Cyprinodontiformes (Euteleostei: Atherinomorpha): a reappraisal. In: L. R. Malabarba, R. E. Reis, R. P. Vari, Z. M. Lucena, and C. A. S. Lucena, eds. Phylogeny and Classification of Neotropical Fishes. Porto Alegre, Brazil: EDIPUCRS. pp. 537–560. Google Scholar

    382.

    Costa, W. J. E. M. 2004. Relationships and redescription of Fundulus brasiliensis (Cyprinodontiformes: Rivulidae), with description of a new genus and notes on the classification of the Aplocheiloidei. Ichthyological Exploration of Freshwaters 15:105–120. Google Scholar

    383.

    Costa, W. J. E. M. 2011. Redescription and phylogenetic position of the fossil killifish †Carrionellus diumortuus White from the Lower Miocene of Ecuador (Teleostei: Cyprinodontiformes). Cybium 35(3):181–187. Google Scholar

    384.

    Costa, W. J. E. M. 2012a. The caudal skeleton of extant and fossil cyprinodontiform fishes (Teleostei: Atherinomorpha): comparative morphology and delimitation of phylogenetic characters. Vertebrate Zoology 62(2):161–180. Google Scholar

    385.

    Costa, W. J. E. M. 2012b. Oligocene killifishes (Teleostei: Cyprinodontiformes) from southern France: relationships, taxonomic position, and evidence of internal fertilization. Vertebrate Zoology 62(3):371–386. Google Scholar

    386.

    Costa, W. J. E. M. 2016. Comparative morphology and classification of South American cynopoeciline killifishes (Cyprinodontiformes: Aplocheilidae), with notes on family-group names used for aplocheiloids. Vertebrate Zoology 66(2):125–140. Google Scholar

    387.

    Costa, W. J. E. M., P. F. Amorim, and J. L. O. Mattos. 2017. Molecular phylogeny and timing of diversification in South American Cynolebiini seasonal killifishes. Molecular Phylogenetics and Evolution 116:61–68. Google Scholar

    388.

    Covain, R., S. Fisch-Muller, C. Oliveira, J. H. Mol, J. I. Montoya-Burgos, and S. Dray. 2016. Molecular phylogeny of the highly diversified catfish subfamily Loricariinae (Siluriformes, Loricariidae) reveals incongruences with morphological classification. Molecular Phylogenetics and Evolution 94(Part B):492–517. Google Scholar

    389.

    Cowman, P. F., D. R. Bellwood, and L. Van Herwerden. 2009. Dating the evolutionary origins of wrasse lineages (Labridae) and the rise of trophic novelty on coral reefs. Molecular Phylogenetics and Evolution 52(3):621–631. Google Scholar

    390.

    Craig, M. T., and P. A. Hastings. 2007. A molecular phylogeny of the groupers of the subfamily Epinephelinae (Serranidae) with a revised classification of the Epinephelini. Ichthyological Research 54:1–17. Google Scholar

    391.

    Cui, L., Y. Dong, R. Cao, X. Zhou, and S. Lu. 2020. Characterization of the complete mitochondrial genome of Arius dispar (Siluriformes: Ariidae) and phylogenetic analysis among sea catfishes. Journal of Ocean University of China 19:1198–1206. Google Scholar

    392.

    Cunha, C., N. Mesquita, T. E. Dowling, A. Gilles, and M. M. Coelho. 2002. Phylogenetic relationships of Eurasian and American cyprinids using cytochrome b sequences. Journal of Fish Biology 61(4):929–944. Google Scholar

    393.

    Cuvier, G. 1816. Le Règne Animal Distribué d'Après son Organisation pour Servir de Base à l'Histoire Naturelle des Animaux et d'Introduction à l'Anatomie Comparée. Tome 2, Les reptiles, les poissons, les mollusques et les annélides. 1st ed. Paris: Chez Déterville. 532 pp. Google Scholar

    394.

    Cuvier, G. 1829. Le Règne Animal Distribué d'Après son Organisation, pour Servir de Base à l'Histoire Naturelle des Animaux et d'Introduction à l'Anatomie Comparée, Tome 2, Les reptiles, les poissons, les mollusques et les annélides. 2nd ed. Paris: Chez Déterville. 406 pp. Google Scholar

    395.

    Cuvier, G., and A. Valenciennes. 1828. Histoire Naturelle des Poissons. Tome 2. Livre troisième, Des Poissons de la Famille des Perches, ou des Percoïdes. Paris: F. G. Levrault. 490 pp. Google Scholar

    396.

    Cuvier, G.. 1829a. Histoire Naturelle des Poissons. Tome 3. Suite du livre troisième, Des Percoïdes à Dorsale Unique à Sept Rayons Branchiaux et à Dents en Velours ou en Cardes. Paris: F. G. Levrault. 500 pp. Google Scholar

    397.

    Cuvier, G.. 1829b. Histoire Naturelle des Poissons. Tome 4. Livre quatrième, Des Acanthoptérygiens à Joue Cuirassée. Paris: F. G. Levrault. 518 pp. Google Scholar

    398.

    Cuvier, G.. 1830. Histoire Naturelle des Poissons. Tome 5. Livre cinquième, Des Sciénoïdes. Paris: F. G. Levrault. 499 pp. Google Scholar

    399.

    Cuvier, G.. 1831. Histoire Naturelle des Poissons. Tome 7. Livre septième, Des Squamipennes. Livre huitième, Des Poissons à Pharyngiens Labyrinthiformes. Paris: F. G. Levrault. 531 pp. Google Scholar

    400.

    Cuvier, G.. 1833. Histoire Naturelle des Poissons. Tome 9. Suite du livre neuvième, Des Scombéroïdes. Paris: F. G. Levrault. 512 pp. Google Scholar

    401.

    Cuvier, G.. 1837. Histoire Naturelle des Poissons. Tome 12. Suite du livre quatorzième, Gobioïdes. Livre quinzième, Acanthoptérygiens à Pectorales Pédiculées. Paris: F. G. Levrault. 507 pp. Google Scholar

    402.

    Cuvier, G.. 1846. Histoire Naturelle des Poissons. Tome 18. Suite du livre dix-huitième, Cyprinoïdes. Livre dix-neuvième, Des Ésoces ou Lucioïdes. Paris: P. Bertrand. 505 pp. Google Scholar

    403.

    Daane, J. M., N. Blum, J. Lanni, H. Boldt, M. K. Iovine, C. W. Higdon, S. L. Johnson, N. R. Lovejoy, and M. P. Harris. 2021. Modulation of bioelectric cues in the evolution of flying fishes. Current Biology 31(22):5052–5061.e8.  https://doi.org/10.1016/j.cub.2021.08.054 Google Scholar

    404.

    Daget, J. 1950. Révision des affinités phylogénétiques des Polyptéridés. Dakar, Senegal: Institut Français d'Afrique Noire. 178 pp. (Mémoires de l'Institut Français d'Afrique Noire 11.) Google Scholar

    405.

    Daget, J., M. Gayet, F. J. Meunier, and J.-Y. Sire. 2001. Major discoveries on the dermal skeleton of fossil and Recent polypteriforms: a review. Fish and Fisheries 2(2):113–124. Google Scholar

    406.

    Dai, W., M. Zou, L. Yang, K. Du, W. Chen, Y. Shen, R. L. Mayden, and S. He. 2018. Phylogenomic perspective on the relationships and evolutionary history of the major otocephalan lineages. Scientific Reports 8:205. Google Scholar

    407.

    Damerau, M., M. Freese, and R. Hanel. 2018. Multigene phylogeny of jacks and pompanos (Carangidae), including placement of monotypic vadigo Campogramma glaycos. Journal of Fish Biology 92(1):190–202. Google Scholar

    408.

    Danil′ Chenko, P. G. 1960. Kostistye ryby Maĭkopskikh otlozheniĭ Kavkaza. Moscow: Izdatel′stvo Akademii͡a nauk SSSR. 208 pp. (Trudy Paleontologicheskogo Instituta 78.) Google Scholar

    409.

    Danil 1968. Ryby verkhnego paleostena Turkmenii. In: D. V. Obruchev, ed. Ocherki po filogenii i sistematike iskopayemykh ryb i beschelyustnykh. Moscow: Nauka. pp. 113–156. Google Scholar

    410.

    Danilit′chenko, P. G. 1962. Fishes from the Dabakhanian Formation of Georgia. Paleontologischeskii Zhurnal 1:111–126. Google Scholar

    411.

    Da Silva, J. P. C. B., A. Datovo, and G. D. Johnson. 2019. Phylogenetic interrelationships of the eel families Derichthyidae and Colocongridae (Elopomorpha: Anguilliformes) based on the pectoral skeleton. Journal of Morphology 280(7):934–947. Google Scholar

    412.

    Da Silva, J. P. C. B., and G. D. Johnson. 2018. Reconsidering pectoral girdle and fin morphology in Anguillidae (Elopomorpha: Anguilliformes). Journal of Fish Biology 93(2):420–423. Google Scholar

    413.

    Datovo, A., and R. M. C. Castro. 2012. Anatomy and evolution of the mandibular, hyopalatine, and opercular muscles in characiform fishes (Teleostei: Ostariophysi). Zoology 115(2):84–116. Google Scholar

    414.

    Datovo, A., M. C. C. De Pinna, and G. D. Johnson. 2014. The infrabranchial musculature and its bearing on the phylogeny of percomorph fishes (Osteichthyes: Teleostei). PLOS ONE 9(10):e110129.  https://doi.org/10.1371/journal.pone.0110129 Google Scholar

    415.

    Datovo, A., and P. P. Rizzato. 2018. Evolution of the facial musculature in basal ray-finned fishes. Frontiers in Zoology 15:40. Google Scholar

    416.

    Datovo, A., and R. P. Vari. 2014. The adductor mandibulae muscle complex in lower teleostean fishes (Osteichthyes: Actinopterygii): comparative anatomy, synonymy, and phylogenetic implications. Zoological Journal of the Linnean Society 171(3):554–622. Google Scholar

    417.

    Davesne, D. 2017. A fossil unicorn crestfish (Teleostei, Lampridiformes, Lophotidae) from the Eocene of Iran. PeerJ 5:e3381. https://doi.org/10.7717/peerj.3381 Google Scholar

    418.

    Davesne, D., G. Carnevale, and M. Friedman. 2017. Bajaichthys elegans from the Eocene of Bolca (Italy) and the overlooked morphological diversity of Zeiformes (Teleostei, Acanthomorpha). Palaeontology 60(2):255–268. Google Scholar

    419.

    Davesne, D., M. Friedman, V. Barriel, G. Lecointre, P. Janvier, C. Gallut, and O. Otero. 2014. Early fossils illuminate character evolution and interrelationships of Lampridiformes (Teleostei, Acanthomorpha). Zoological Journal of the Linnean Society 172(2):475–498. Google Scholar

    420.

    Davesne, D., C. Gallut, V. Barriel, P. Janvier, G. Lecointre, and O. Otero. 2016. The phylogenetic intrarelationships of spiny-rayed fishes (Acanthomorpha, Teleostei, Actinopterygii): fossil taxa increase the congruence of morphology with molecular data. Frontiers in Ecology and Evolution 4:129.  https://doi.org/10.3389/fevo.2016.00129 Google Scholar

    421.

    Davis, M. P. 2010. Evolutionary relationships of the Aulopiformes (Euteleostei: Cyclosquamata): a molecular and total evidence approach. In: J. S. Nelson, H.-P. Schultze, and M. V. H. Wilson, eds. Origin and Phylogenetic Interrelationships of Teleosts. Munich: Verlag Dr. Friedrich Pfeil. pp. 431–470. Google Scholar

    422.

    Davis, M. P., G. Arratia, and T. M. Kaiser. 2013. The first fossil shellear and its implications for the evolution and divergence of the Kneriidae (Teleostei: Gonorynchiformes). In: G. Arratia, H.-P. Schultze, and M. V. H. Wilson, eds. Mesozoic Fishes 5: Global Diversity and Evolution. Munich: Verlag Dr. Friedrich Pfeil. pp. 325–362. Google Scholar

    423.

    Davis, M. P., and C. Fielitz. 2010. Estimating divergence times of lizardfishes and their allies (Euteleostei: Aulopiformes) and the timing of deep-sea adaptations. Molecular Phylogenetics and Evolution 57(3):1194–1208. Google Scholar

    424.

    Davis, M. P., N. I. Holcroft, E. O. Wiley, J. S. Sparks, and W. L. Smith. 2014. Species-specific bioluminescence facilitates speciation in the deep sea. Marine Biology 161:1139–1148. Google Scholar

    425.

    Davis, M. P., J. S. Sparks, and W. L. Smith. 2016. Repeated and widespread evolution of bioluminescence in marine fishes. PLOS ONE 11(6):e0155154.  https://doi.org/10.1371/journal.pone.0155154 Google Scholar

    426.

    Day, J. J., A. Fages, K. J. Brown, E. J. Vreven, M. L. J. Stiassny, R. Bills, J. P. Friel, and L. Rüber. 2017. Multiple independent colonizations into the Congo Basin during the continental radiation of African Mastacembelus spiny eels. Journal of Biogeography 44(10):2308–2318. Google Scholar

    427.

    Dean, B. 1895. Fishes, Living and Fossil: An Outline of their Forms and Probable Relationships. New York: Macmillan and Co. 300 pp. (Columbia Biological Series 3.) Google Scholar

    428.

    De Figueiredo, F. J. 2005. Reassessment of the morphology of Scombroclupeoides scutata Woodward, 1908, a teleostean fish from the Early Cretaceous of Bahia, with comments on its relationships. Arquivos do Museu Nacional do Rio de Janeiro 63(3):507–522. Google Scholar

    429.

    De Figueiredo, F. J. 2009. A new clupeiform fish from the Lower Cretaceous (Barremian) of Sergipe-Alagoas Basin, northeastern Brazil. Journal of Vertebrate Paleontology 29(4):993–1005. Google Scholar

    430.

    De Figueiredo, F. J., and V. Gallo. 2004. A new teleost fish from the Early Cretaceous of northeastern Brazil. Boletim do Museu Nacional do Rio de Janeiro 73:1–23. Google Scholar

    431.

    De Figueiredo, F. J., V. Gallo, and A. F. P. Delarmelina. 2012. A new protacanthopterygian fish from the Upper Cretaceous (Turonian) of the Pelotas Basin, southern Brazil. Cretaceous Research 34:116–123. Google Scholar

    432.

    De Figueiredo, F. J., V. Gallo, and M. E. C. Leal. 2012. Phylogenetic relationships of the elopomorph fish †Paraelops cearensis Silva Santos revisited: evidence from new specimens. Cretaceous Research 37:148–154. Google Scholar

    433.

    Delbarre, D. J., D. Davesne, and M. Friedman. 2016. Anatomy and relationships of †Aipichthys pretiosus and †‘Aipichthys’ nuchalis (Acanthomorpha: Lampridomorpha), with a review of Late Cretaceous relatives of oarfishes and their allies. Journal of Systematic Palaeontology 14(7):545–567. Google Scholar

    434.

    Demartini, E. E., and T. J. Donaldson. 1996. Color morph-habitat relations to the arc-eye hawkfish Paracirrhites arcatus (Pisces: Cirrhitidae). Copeia 1996(2):362–371. Google Scholar

    435.

    Denton, J. S. S. 2014. Seven-locus molecular phylogeny of Myctophiformes (Teleostei; Scopelomorpha) highlights the utility of the order for studies of deep-sea evolution. Molecular Phylogenetics and Evolution 76:270–292. Google Scholar

    436.

    Denys, G. P. J., A. Dettaï, H. Persat, M. Hautecœur, and P. Keith. 2014. Morphological and molecular evidence of three species of pikes Esox spp. (Actinopterygii, Esocidae) in France, including the description of a new species. Comptes Rendus Biologies 337(9):521–534. Google Scholar

    437.

    Denys, G. P. J., T. Lauga, G. B. Delmastro, and A. Dettaï. 2018. S7 characterization of Western European pikes Esox spp. (Actinopterygii, Esociformes). Cybium 42(3):221–228. Google Scholar

    438.

    De Pinna, M. C. C. 1993. Higher-level phylogeny of Siluriformes, with a new classification of the order (Teleostei, Ostariophysi) [dissertation]. New York: City University of New York. pp. 482. Google Scholar

    439.

    De Pinna, M. C. C. 1996. Teleostean monophyly. In: M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson, eds. Interrelationships of Fishes. San Diego: Academic Press. pp. 147–162. Google Scholar

    440.

    De Pinna, M. C. C. 1998. Phylogenetic relationships of Neotropical Siluriformes (Teleostei: Ostariophysi): historical overview and synthesis of hypotheses. In: L. R. Malabarba, R. E. Reis, R. P. Vari, Z. M. Lucena, and C. A. S. Lucena, eds. Phylogeny and Classification of Neotropical Fishes. Porto Alegre, Brazil: EDIPUCRS. pp. 279–330. Google Scholar

    441.

    De Pinna, M. C. C., and F. Di Dario. 2010. The branchial arches of the primitive clupeomorph fish Denticeps clupeoides, and their phylogenetic implication (Clupeiformes, Denticipitidae). In: J. S. Nelson, H.-P. Schultze, and M. V. H. Wilson, eds. Origin and Phylogenetic Interrelationships of Teleosts. Munich: Verlag Dr. Friedrich Pfeil. pp. 251–268. Google Scholar

    442.

    De Pinna, M. C. C., J. Zuanon, L. Rapp Py-Daniel, and P. Petry. 2018. A new family of Neotropical freshwater fishes from deep fossorial Amazonian habitat, with a reappraisal of morphological characiform phylogeny (Teleostei: Ostariophysi). Zoological Journal of the Linnean Society 182(1):76–106. Google Scholar

    443.

    De Queiroz, K., P. D. Cantino, and J. A. Gauthier. 2020a. Introduction. In: K. De Queiroz, P. D. Cantino, and J. A. Gauthier, eds. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC Press. pp. xv–xxvii. Google Scholar

    444.

    De Queiroz, K., eds. 2020b. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC Press. 1324 pp. Google Scholar

    445.

    De Queiroz, K., and J. Gauthier. 1990. Phylogeny as a central principle in taxonomy: Phylogenetic definition of taxon names. Systematic Zoology 39(4):307–322. Google Scholar

    446.

    De Queiroz, K.. 1992. Phylogenetic taxonomy. Annual Review of Ecology and Systematics 23:449–480. Google Scholar

    447.

    De Queiroz, K.. 1994. Toward a phylogenetic system of biological nomenclature. Trends in Ecology and Evolution 9(1):27–31. Google Scholar

    448.

    Derome, N., W.-J. Chen, A. Dettaï, C. Bonillo, and G. Lecointre. 2002. Phylogeny of Antarctic dragonfishes (Bathydraconidae, Notothenioidei, Teleostei) and related families based on their anatomy and two mitochondrial genes. Molecular Phylogenetics and Evolution 24(1):139–152. Google Scholar

    449.

    Derouen, V., W. B. Ludt, H.-C. Ho, and P. Chakrabarty. 2015. Examining evolutionary relationships and shifts in depth preferences in batfishes (Lophiiformes: Ogcocephalidae). Molecular Phylogenetics and Evolution 84(Part A):27–33. Google Scholar

    450.

    De Sousa, R. P. C., G. C. Silva-Oliveira, I. O. Furo, A. B. De Oliveira-Filho, C. D. B. De Brito, L. Rabelo, A. Guimarães-Costa, E. H. C. De Oliveira, and M. Vallinoto. 2021. The role of the chromosomal rearrangements in the evolution and speciation of Elopiformes fishes (Teleostei; Elopomorpha). Zoologischer Anzeiger 290:40–48. Google Scholar

    451.

    Dettaï, A., M. Berkani, A. C. Lautredou, A. Couloux, G. Lecointre, C. Ozouf-Costaz, and C. Gallut. 2012. Tracking the elusive monophyly of nototheniid fishes (Teleostei) with multiple mitochondrial and nuclear markers. In: Molecular and Genetic Advances to Understanding Evolution and Biodiversity in the Polar Regions, edited by C. Verde, P. Convey, and G. Di Prisco. Special issue, Marine Genomics 8:49–58. Google Scholar

    452.

    Dettaï, A., and G. Lecointre. 2004. In search of notothenioid (Teleostei) relatives. Antarctic Science 16:71–85. Google Scholar

    453.

    Dettaï, A.. 2005. Further support for the clades obtained by multiple molecular phylogenies in the acanthomorph bush. Comptes Rendus Biologies 328(7):647–689. Google Scholar

    454.

    Dettaï, A.. 2008. New insights into the organization and evolution of vertebrate IRBP genes and utility of IRBP gene sequences for the phylogenetic study of the Acanthomorpha (Actinopterygii: Teleostei). Molecular Phylogenetics and Evolution 48(1):258–269. Google Scholar

    455.

    Díaz-Cruz, J. A., J. Alvarado-Ortega, and S. Giles. 2020. A long snout enchodontid fish (Aulopiformes: Enchodontidae) from the Early Cretaceous deposits at the El Chango quarry, Chiapas, southeastern Mexico: a multi-approach study. Palaeontolgia Electronica 23(2):a30.  https://doi.org/10.26879/1065 Google Scholar

    456.

    Díaz-Cruz, J. A., J. Alvarado-Ortega, M. M. Ramírez-Sánchez, E. L. Bernard, L. Allington-Jones, and M. Graham. 2021. Phylogenetic morphometrics, geometric morphometrics and the Mexican fossils to understand evolutionary trends of enchodontid fishes. Journal of South American Earth Sciences 111:103492.  https://doi.org/10.1016/j.jsames.2021.103492 Google Scholar

    457.

    Di Dario, F. 2004. Homology between the recessus lateralis and cephalic sensory canals, with the proposition of additional synapomorphies for the Clupeiformes and the Clupeoidei. Zoological Journal of the Linnean Society 141(2):257–270. Google Scholar

    458.

    Di Dario, F. 2009. Chirocentrids as engrauloids: evidence from suspensorium, branchial arches, and infraorbital bones (Clupeomorpha, Teleostei). Zoological Journal of the Linnean Society 156(2):363–383. Google Scholar

    459.

    Di Dario, F., and M. C. C. De Pinna. 2006. The supratemporal system and the pattern of ramification of cephalic sensory canals in Denticeps clupeoides (Denticipitoidei, Teleostei): additional evidence for monophyly of Clupeiformes and Clupeoidei. Papéis Avulsos de Zoologia 46(10):107–123. Google Scholar

    460.

    Dietze, K. 2009. Morphology and phylogenetic relationships of certain neoteleostean fishes from the Upper Cretaceous of Sendenhorst, Germany. Cretaceous Research 30(3):559–574. Google Scholar

    461.

    Dillman, C. B., D. E. Bergstrom, D. B. Noltie, T. P. Holtsford, and R. L. Mayden. 2011. Regressive progression, progressive regression or neither? Phylogeny and evolution of the Percopsiformes (Teleostei, Paracanthopterygii). Zoologica Scripta 40(1):45–60. Google Scholar

    462.

    Dillman, C. B., B. L. Sidlauskas, and R. P. Vari. 2016. A morphological supermatrix-based phylogeny for the Neotropical fish superfamily Anostomoidea (Ostariophysi: Characiformes): phylogeny, missing data and homoplasy. Cladistics 32(3):276–296. Google Scholar

    463.

    Dimmick, W. W., and A. Larson. 1996. A molecular and morphological perspective on the phylogenetic relationships of the otophysan fishes. Molecular Phylogenetics and Evolution 6(1):120–133. Google Scholar

    464.

    Ding, Z., Y. Xu, W. Chen, Y. Liu, C. Wang, Y. Niu, K. Zhang, Y. Wang, and L. Yang. 2023. Stronger selective constraints on the mitochondrial genome in flying fishes. Frontiers in Marine Science 10:1168417.  https://doi.org/10.3389/fmars.2023.1168417 Google Scholar

    465.

    Diogo, R. 2003. Higher-level phylogeny of Siluriformes—an overview. In: G. Arratia, B. G. Kapoor, M. Chardon, and R. Diogo, eds. Catfishes, Volume 1. Enfield, NH: Science Publishers. pp. 353–384. Google Scholar

    466.

    Diogo, R.2004. Morphological Evolution, Adaptations, Homoplasies, Constraints, and Evolutionary Trends: Catfishes as a Case Study on General Phylogeny and Macroevolution. Enfield, NH: Science Publishers. 491 pp. Google Scholar

    467.

    Diogo, R.2007. The Origin of Higher Clades: Osteology, Myology, Phylogeny, and the Evolution of Bony Fishes and the Rise of Tetrapods. Enfield, NH: Science Publishers. 367 pp. Google Scholar

    468.

    Diogo, R.2009. Origin, evolution and homologies of the Weberian apparatus: a new insight. International Journal of Morphology 27:333–354. Google Scholar

    469.

    Diogo, R., I. Doadrio, and P. Vandewalle. 2008. Teleostean phylogeny based on osteological and myological characters. International Journal of Morphology 26:463–522. Google Scholar

    470.

    Doiuchi, R., T. Sato, and T. Nakabo. 2004. Phylogenetic relationships of the stromateoid fishes (Perciformes). Ichthyological Research 51:202–212. Google Scholar

    471.

    Dollo, L. 1904. Poissons. Expedition antarctique Belge. Resultats du voyage du S. Y. Belgica en 1897–1898–1899 sous le commandement de A. de Gerlache de Gomery. Antwerp: J.-E. Buschmann. 240 pp. Google Scholar

    472.

    Dornburg, A., M. Friedman, and T. J. Near. 2015. Phylogenetic analysis of molecular and morphological data highlights uncertainty in the relationships of fossil and living species of Elopomorpha (Actinopterygii: Teleostei). Molecular Phylogenetics and Evolution 89:205–218. Google Scholar

    473.

    Dornburg, A., and T. J. Near. 2021. The emerging phylogenetic perspective on the evolution of actinopterygian fishes. Annual Review of Ecology, Evolution, and Systematics 52:427–452. Google Scholar

    474.

    Dornburg, A., J. P. Townsend, W. Brooks, E. Spriggs, R. I. Eytan, J. A. Moore, P. C. Wainwright, A. R. Lemmon, E. M. Lemmon, and T. J. Near. 2017. New insights on the sister lineage of percomorph fishes with an anchored hybrid enrichment dataset. Molecular Phylogenetics and Evolution 110:27–38. Google Scholar

    475.

    Dunn, C. W., A. Hejnol, D. Q. Matus, K. Pang, W. E. Browne, S. A. Smith, E. Seaver, G. W. Rouse, M. Obst, G. D. Edgecombe, ET AL. 2008. Broad phylogenomic sampling improves resolution of the animal tree of life. Nature 452:745–749. Google Scholar

    476.

    Dunn, J. R. 1989. A provisional phylogeny of gadid fishes based on adult and early life-history characters. In: D. M. Cohen, ed. Papers on the Systematics of Gadiform Fishes. Los Angeles: Natural History Museum of Los Angeles County. pp. 209–235. Google Scholar

    477.

    Duong, T. Y., L. T. K. Pham, X. T. K. Le, N. T. T. Nguyen, S. A. M. Nor, and T. H. Le. 2023. Mitophylogeny of pangasiid catfishes and its taxonomic implications for Pangasiidae and the suborder Siluroidei. Zoological Studies 62:48. Google Scholar

    478.

    Dutheil, D. B. 1999. An overview of the freshwater fish fauna from the Kem Kem beds (Late Cretaceous: Cenomanian) of southeastern Morocco. In: G. Arratia and H.-P. Schultze, eds. Mesozoic Fishes 2: Systematics and Fossil Record. Munich: Verlag Dr. Friedrich Pfeil. pp. 553–563. Google Scholar

    479.

    Dyer, B. S. 2006. Systematic revision of the South American silversides (Teleostei, Atheriniformes). Biocell 30:69–88. Google Scholar

    480.

    Dyer, B. S., and B. Chernoff. 1996. Phylogenetic relationships among atheriniform fishes (Teleostei: Atherinomorpha). Zoological Journal of the Linnean Society 117(1):1–69. Google Scholar

    481.

    Dyldin, Y., L. Hanel, R. Fricke, A. Orlov, V. Romanov, J. Plesnik, E. Interesova, D. Vorobiev, and M. Kochetkova. 2020. Fish diversity in freshwater and brackish water ecosystems of Russia and adjacent waters. Publications of the Seto Marine Biological Laboratory 45:47–116. Google Scholar

    482.

    Eastman, J. T., and R. R. Eakin. 2021. Checklist of the species of notothenioid fishes. Antarctic Science 33(3):273–280. Google Scholar

    483.

    Ebert, M. 2018. Cerinichthys koelblae, gen. et sp. nov., from the Upper Jurassic of Cerin, France, and its phylogenetic setting, leading to a reassessment of the phylogenetic relationships of Halecomorphi (Actinopterygii). Journal of Vertebrate Paleontology 38:e1420071.  https://doi.org/10.1080/02724634.2017.1420071 Google Scholar

    484.

    Echelle, A. A., E. W. Carson, A. F. Echelle, R. A. Van Den Bussche, T. E. Dowling, and A. Meyer. 2005. Historical biogeography of the New-World pupfish genus Cyprinodon (Teleostei: Cyprinodontidae). Copeia 2005(2):320–339. Google Scholar

    485.

    Echelle, A. A., L. Fuselier, R. A. Van Den Bussche, C. M. L. Rodriguez, and M. L. Smith. 2006. Molecular systematics of Hispaniolan pupfishes (Cyprinodontidae: Cyprinodon): implications for the biogeography of insular Caribbean fishes. Molecular Phylogenetics and Evolution 39(3):855–864. Google Scholar

    486.

    Egan, J. P., D. D. Bloom, C.-H. Kuo, M. P. Hammer, P. Tongnunui, S. P. Iglésias, M. Sheaves, C. Grudpan, and A. M. Simons. 2018. Phylogenetic analysis of trophic niche evolution reveals a latitudinal herbivory gradient in Clupeoidei (herrings, anchovies, and allies). Molecular Phylogenetics and Evolution 124:151–161. Google Scholar

    487.

    Eigenmann, C. H. 1909. Reports on the expedition to British Guiana of the Indiana University and the Carnegie Museum, 1908. Report no. 1. Some new genera and species of fishes from British Guiana. Annals of the Carnegie Museum 6(1):4–54. Google Scholar

    488.

    Eigenmann, C. H. 1922. The Fishes of Western South America, Part I. The Fresh-water Fishes of Northwestern South America, Including Colombia, Panama, and the Pacific Slopes of Ecuador and Peru, Together with an Appendix upon the Fishes of the Rio Meta in Colombia. Pittsburgh: Carnegie Institute. 346 pp. (Memoirs of the Carnegie Museum 9, no. 1.) Google Scholar

    489.

    Eigenmann, C. H., and R. S. Eigenmann. 1890. A Revision of the South American Nematognathi or Cat-fishes. San Francisco: California Academy of Sciences. 508 pp. (Occasional Papers 1.) Google Scholar

    490.

    Endo, H. 2002. Phylogeny of the order Gadiformes (Teleostei, Paracanthopterygii). Memoirs of the Graduate School of Fisheries Sciences 49:75–149. Google Scholar

    491.

    Espíndola, V. C., G. D. Johnson, and M. C. C. De Pinna. 2023. Facial and opercular muscles in the Anguilliformes (Elopomorpha: Teleostei): comparative anatomy and phylogenetic implications for the basal position of Protanguilla. Journal of Morphology 284(3):e21556.  https://doi.org/10.1002/jmor.21556 Google Scholar

    492.

    Everly, A. W. 2002. Stages of development of the Goosefish, Lophius americanus, and comments on the phylogenetic significance of the development of the luring apparatus in Lophiiformes. Environmental Biology of Fishes 64:393–417. Google Scholar

    493.

    Evseenko, S. A., N. V. Gordeeva, Y. Y. Bolshakova, and S. H. Kobyliansky. 2018. Morphology and molecular phylogenetic relationships of Barathronus multidens (Ophidiiformes: Bythitidae). Cybium 42(2):137–141. Google Scholar

    494.

    Eytan, R. I., B. R. Evans, A. Dornburg, A. R. Lemmon, E. M. Lemmon, P. C. Wainwright, and T. J. Near. 2015. Are 100 enough? Inferring acanthomorph teleost phylogeny using Anchored Hybrid Enrichment. BMC Evolutionary Biology 15:113.  https://doi.org/10.1186/s12862-015-0415-0 Google Scholar

    495.

    Fabricus, O. 1780. Fauna Groenlandica, Systematice Sistens Animalia Groenlandiae Occidentalis Hactenus Indagata, Quoad Nomen Specificum, Trivale, Vernaculumque. Copenhagen: I. G. Rothe. 452 pp. Google Scholar

    496.

    Fahay, M. P., and D. F. Markle. 1984. Gadiformes: development and relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 265–283. (Special Publication 1.) Google Scholar

    497.

    Faircloth, B. C., F. Alda, K. Hoekzema, M. D. Burns, C. Oliveira, J. S. Albert, B. F. Melo, L. E. Ochoa, F. F. Roxo, P. Chakrabarty, ET AL. 2020. A target enrichment bait set for studying relationships among ostariophysan fishes. Copeia 108(1):47–60. Google Scholar

    498.

    Faircloth, B. C., L. Sorenson, F. Santini, and M. E. Alfaro. 2013. A phylogenomic perspective on the radiation of ray-finned fishes based upon targeted sequencing of ultraconserved elements (UCEs). PLOS ONE 8(6): e65923.  https://doi.org/10.1371/journal.pone.0065923 Google Scholar

    499.

    Fang, F., M. Norén, T.-Y. Liao, M. Källersjö, and S. O. Kullander. 2009. Molecular phylogenetic interrelationships of the south Asian cyprinid genera Danio, Devario and Microrasbora (Teleostei, Cyprinidae, Danioninae). Zoologica Scripta 38(3):237–256. Google Scholar

    500.

    [FAO] Food And Agriculture Organization Of The United Nations. 2020. The State of World Fisheries and Aquaculture 2020: Sustainability in Action. Rome: FAO. Accessed 3 November 2023.  https://www.fao.org/state-of-fisheries-aquaculture/2020/en/  Google Scholar

    501.

    Fara, E., M. Gayet, and L. Taverne. 2010. The fossil record of Gonorynchiformes. In: T. Grande, F. J. Poyato-Ariza, and R. Diogo, eds. Gonorynchiformes and Ostariophysan Relationships: A Comprehensive Review. Enfield, NH: Science Publishers. pp. 173–226. Google Scholar

    502.

    Feng, D.-H., G. H. Xu, X.-Y. Ma, and Y. Ren. 2023. Taxonomic revision of Sinoeugnathus kueichowensis (Halecomorphi, Holostei) from the Middle Triassic of Guizhou and Yunnan, China. Vertebrata PalAsiatica 61(3):161–181. Google Scholar

    503.

    Ferraris, C. J. 2007. Checklist of catfishes, recent and fossil (Osteichthyes: Siluriformes), and catalogue of siluriform primary types. Zootaxa 1418(1):1–628.  https://doi.org/10.11646/zootaxa.1418.1.1 Google Scholar

    504.

    Ferraris, C. J., C. D. De Santana, and R. P. Vari. 2017. Checklist of Gymnotiformes (Osteichthyes: Ostariophysi) and catalogue of primary types. Neotropical Ichthyology 15:e160067.  https://doi.org/10.1590/1982-0224-20160067 Google Scholar

    505.

    Ferrer Aledo, J. 1930. Catálogo de los peces de Menorca. Revista de Menorca 29:225–259. Google Scholar

    506.

    Fessler, J. L., and M. W. Westneat. 2007. Molecular phylogenetics of the butterflyfishes (Chaetodontidae): taxonomy and biogeography of a global coral reef fish family. Molecular Phylogenetics and Evolution 45(1):50–68. Google Scholar

    507.

    Fielitz, C. 2002. A new Late Cretaceous (Turonian) basal euteleostean fish from Lac des Bois of the Northwest Territories of Canada. Canadian Journal of Earth Sciences 39(11):1579–1590. Google Scholar

    508.

    Fielitz, C. 2004. The phylogenetic relationships of the †Enchodontidae (Teleostei: Aulopiformes). In: G. Arratia, M. V. H. Wilson, and R. Cloutier, eds. Recent Advances in the Origin and Early Radiation of Vertebrates. Munich: Verlag Dr. Friedrich Pfeil. pp. 619–634. Google Scholar

    509.

    Fielitz, C., and D. Bardack. 1992. Deltaichthys albuloides, a new and unusually preserved albulid (Teleostei) probably from the Cretaceous of Texas. Journal of Vertebrate Paleontology 12(2):133–141. Google Scholar

    510.

    Fielitz, C., and K. A. González-Rodríguez. 2010. A new species of Enchodus (Aulopiformes: Enchodontidae) from the Cretaceous (Albian to Cenomanian) of Zimapan, Hidalgo, Mexico. Journal of Vertebrate Paleontology 30(5):1343–1351. Google Scholar

    511.

    Filleul, A. 2000. Baugeichthys caeruleus, gen. et sp nov., a new albuliform fish from the Hauterivian of the Massif des Bauges (France). Journal of Vertebrate Paleontology 20(4):637–644. Google Scholar

    512.

    Filleul, A., and S. Lavoue. 2001. Basal teleosts and the question of elopomorph monophyly. Morphological and molecular approaches. Comptes rendus de l'Académie des sciences Série 3: sciences de la vie 324(4):393–399. Google Scholar

    513.

    Filleul, A., and J. G. Maisey. 2004. Redescription of Santanichthys diasii (Otophysi, Characiformes) from the Albian of the Santana formation and comments on its implications for otophysan relationships. American Museum Novitates 3455:1–21. Google Scholar

    514.

    Fink, S. V., and W. L. Fink. 1981. Interrelationships of the ostariophysan fishes (Teleostei). Zoological Journal of the Linnean Society 72(4):297–353. Google Scholar

    515.

    Fink, S. V.. 1996. Interrelationships of ostariophysan fishes (Teleostei). In: M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson, eds. Interrelationships of Fishes. San Diego: Academic Press. pp. 209–249. Google Scholar

    516.

    Fink, W. L. 1984a. Basal euteleosts: relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 202–206. (Special Publication 1.) Google Scholar

    517.

    Fink, W. L. 1984b. Stomiiformes: relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 181–184. (Special Publication 1.) Google Scholar

    518.

    Fink, W. L. 1985. Phylogenetic Interrelationships of the Stomiid Fishes (Teleostei: Stomiiformes). Ann Arbor: University of Michigan Museum of Zoology. 127 pp. (Miscellaneous Publications 171.) Google Scholar

    519.

    Fink, W. L., and S. H. Weitzman. 1982. Relationships of the stomiiform fishes (Teleostei), with a description of Diplophos. Bulletin of the Museum of Comparative Zoology 150(2):31–93. Google Scholar

    520.

    Fitzinger, L. J. F. J. 1832. Ueber die Ausarbeitung einer Fauna des Erzherzogthumes Öesterreich, nebst einer systematischen Aufzählung der in diesem Lande vorkommenden Säugethiere, Reptilien und Fische, als Prodrum einer Fauna derfelben. Beiträge zur Landeskunde Österreichs unter der Enns 1:280–340. Google Scholar

    521.

    Fletcher, G. L., M. A. Shears, M. J. King, M. H. Kao, P. L. Davies, and C. L. Hew. 1988. Antifreeze protein genes: physiological regulation and potential value to the genetic engineering of freeze resistant fish. In: R. C. Ryans, ed. Fish Physiology, Fish Toxicology, and Fisheries Management: Proceedings of an International Symposium, Guangzhou, PRC, September 14–16, 1988. Athens, GA: U.S. Environmental Protection Agency. pp. 228–247. Google Scholar

    522.

    Forey, P. L. 1973a. Relationships of elopomorphs. In: P. H. Greenwood, R. S. Miles, and C. Patterson, eds. Interrelationships of Fishes. London: Academic Press. pp. 351–368. Google Scholar

    523.

    Forey, P. L. 1973b. A revision of the elopiform fishes, fossil and Recent. Bulletin of the British Museum (Natural History), Geology, Supplement 10:1–222. Google Scholar

    524.

    Forey, P. L. 1980. Latimeria: a paradoxical fish. Proceedings of the Royal Society of London Series B, Biological Sciences 208(1172):369–384. Google Scholar

    525.

    Forey, P. L. 1997. A Cretaceous notopterid (Pisces: Osteoglossomorpha) from Morocco. South African Journal of Science 93:564–569. Google Scholar

    526.

    Forey, P. L. 2004. A three-dimensional skull of a primative clupeomorph from the Cenomanian English Chalk, and implications for the evolution of the clupeomorph acusticolateralis system. In: G. Arratia and A. Tintori, eds. Mesozoic Fishes 3: Systematics, Paleoenvironments and Biodiversity. Munich: Verlag Dr. Friedrich Pfeil. pp. 405–427. Google Scholar

    527.

    Forey, P. L., and E. J. Hilton. 2010. Two new Tertiary osteoglossid fishes (Teleostei: Osteoglossomorpha) with notes on the history of the family. In: D. K. Elliot, J. G. Maisey, X. Yu, and D. Miao, eds. Morphology, Phylogeny, and Paleobiogeography of Fossil Fishes. Munich: Verlag Dr. Friedrich Pfeil. pp. 215–246. Google Scholar

    528.

    Forey, P. L., D. T. J. Littlewood, P. Ritchie, and A. Meyer. 1996. Interrelationships of elopomorph fishes. In: M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson, eds. Interrelationships of Fishes. San Diego: Academic Press. pp. 174–191. Google Scholar

    529.

    Forey, P. L., Y. Lu, C. Patterson, and C. E. Davis. 2003. Fossil fishes from the Cenomanian (Upper Cretaceous) of Namoura, Lebanon. Journal of Systematic Palaeontology 1(4):230–330. Google Scholar

    530.

    Forey, P. L., and J. G. Maisey. 2010. Structure and relationships of †Brannerion (Albuloidei), an Early Cretaceous teleost from Brazil. In: J. S. Nelson, H.-P. Schultze, and M. V. H. Wilson, eds. Origin and Phylogenetic Interrelationships of Teleosts. Munich: Verlag Dr. Friedrich Pfeil. pp. 183–218. Google Scholar

    531.

    Forster, J. R. 1773. An account of some curious fishes, sent from Hudson's Bay; by Mr. John Reinhold Forster, F. R. S. in a letter to Thomas Pennant, Esq; F. R. S. Philosophical Transactions of the Royal Society of London 63:149–160. Google Scholar

    532.

    Fowler, H. W. 1934. Descriptions of new fishes obtained 1907 to 1910, chiefly in the Philippine Islands and adjacent seas. Proceedings of the Academy of Natural Sciences of Philadelphia 85:233–367. Google Scholar

    533.

    Fowler, H. W. 1951. Analysis of the fishes of Chile. Revista Chilena de Historia Natural 51–53:263–326. Google Scholar

    534.

    Francisco, S. M., L. Congiu, S. Stefanni, R. Castilho, A. Brito, P. P. Ivanova, A. Levy, H. Cabral, G. Kilias, I. Doadrio, and V. C. Almada. 2008. Phylogenetic relationships of the north-eastern Atlantic and Mediterranean forms of Atherina (Pisces, Atherinidae). Molecular Phylogenetics and Evolution 48(2):782–788. Google Scholar

    535.

    Francisco, S. M., L. Congiu, S. Von Der Heyden, and V. C. Almada. 2011. Multilocus phylogenetic analysis of the genus Atherina (Pisces: Atherinidae). Molecular Phylogenetics and Evolution 61(1):71–78. Google Scholar

    536.

    Fraser, T. H. 1972. Some thoughts about the teleostean fish concept—the Paracanthopterygii. Japanese Journal of Ichthyology 19(4):232–242. Google Scholar

    537.

    Fraser, T. H. 2013. A new genus of cardinalfish (Apogonidae: Percomorpha), redescription of Archamia and resemblances and relationships with Kurtus (Kurtidae: Percomorpha). Zootaxa 3714(1):1–63.  https://doi.org/10.11646/zootaxa.3714.1.1 Google Scholar

    538.

    Freyhof, J., M. Özuluğ, and G. Saç. 2017. Neotype designation of Aphanius iconii, first reviser action to stabilise the usage of A. fontinalis, and A. meridionalis and comments on the family group names of fishes placed in Cyprinodontidae (Teleostei: Cyprinodontiformes). Zootaxa 4294(5):573–585.  https://doi.org/10.11646/zootaxa.4294.5.6 Google Scholar

    539.

    Fricke, R., J.-N. Chen, and W.-J. Chen. 2017. New case of lateral asymmetry in fishes: a new subfamily, genus and species of deep water clingfishes from Papua New Guinea, western Pacific Ocean. Comptes Rendus Biologies 340(1):47–62. Google Scholar

    540.

    Fricke, R., W. N. Eschmeyer, and J. D. Fong. 2023. Genera/species by family/subfamily. Eschmeyer's Catalog of Fishes [online database]. California Academy of Sciences. Accessed September 1, 2023.  https://researcharchive.calacademy.org/research/ichthyology/catalog/SpeciesByFamily.asp  Google Scholar

    541.

    Friedman, M. 2008. The evolutionary origin of flatfish asymmetry. Nature 454:209–212. Google Scholar

    542.

    Friedman, M. 2009. Ecomorphological selectivity among marine teleost fishes during the end-Cretaceous extinction. Proceedings of the National Academy of Sciences of the United States of America 106:5218–5223. Google Scholar

    543.

    Friedman, M. 2010. Explosive morphological diversification of spinyfinned teleost fishes in the aftermath of the end-Cretaceous extinction. Proceedings of the Royal Society B, Biological Sciences 277(1688):1675–1683. Google Scholar

    544.

    Friedman, M. 2012. Osteology of †Heteronectes chaneti (Acanthomorpha, Pleuronectiformes), an Eocene stem flatfish, with a discussion of flatfish sister-group relationships. Journal of Vertebrate Paleontology 32(4):735–756. Google Scholar

    545.

    Friedman, M., H. T. Beckett, R. A. Close, and Z. Johanson. 2016. The English chalk and London clay: two remarkable British bony fish Lagerstätten. In: Z. Johanson, P. M. Barrett, M. Richter, and M. Smith, eds. Arthur Smith Woodward: His Life and Influence on Modern Vetebrate Palaeontolology. London: Geological Society. pp. 165–200. (Special Publication 430.) Google Scholar

    546.

    Friedman, M., K. L. Feilich, H. T. Beckett, M. E. Alfaro, B. C. Faircloth, D. Černý, M. Miya, T. J. Near, and R. C. Harrington. 2019. A phylogenomic framework for pelagiarian fishes (Acanthomorpha: Percomorpha) highlights mosaic radiation in the open ocean. Proceedings of the Royal Society B, Biological Sciences 286(1910):20191502. Google Scholar

    547.

    Friedman, M., Z. Johanson, R. C. Harrington, T. J. Near, and M. R. Graham. 2013. An early fossil remora (Echeneoidea) reveals the evolutionary assembly of the adhesion disc. Proceedings of the Royal Society B, Biological Sciences 280(1766):20131200. Google Scholar

    548.

    Friedman, M., and G. D. Johnson. 2005. A new species of Mene (Perciformes: Menidae) from the Paleocene of South America, with notes on paleoenvironment and a brief review of menid fishes. Journal of Vertebrate Paleontology 25(4):770–783. Google Scholar

    549.

    Friedman, M., B. P. Keck, A. Dornburg, R. I. Eytan, C. H. Martin, C. D. Hulsey, P. C. Wainwright, and T. J. Near. 2013. Molecular and fossil evidence place the origin of cichlid fishes long after Gondwanan rifting. Proceedings of the Royal Society B, Biological Sciences 280(1770):20131733. Google Scholar

    550.

    Friedman, S. T., and M. M. Muñoz. 2023. A latitudinal gradient of deep-sea invasions for marine fishes. Nature Communications 14:773. Google Scholar

    551.

    Frost, G. A. 1928. Otoliths of fishes from the Tertiary formations of New Zealand, and from Balcombe Bay, Victoria. Transactions and Proceedings of the Royal Society of New Zealand 59:91–97. Google Scholar

    552.

    Gago, F. J. 1997. Character evolution and phylogeny of the cutlassfishes: an ontogenetic perspective (Scombroidei: Trichiuridae). Bulletin of Marine Science 60:161–191. Google Scholar

    553.

    Gago, F. J. 1998. Osteology and Phylogeny of the Cutlassfishes (Scombroidei: Trichiuridae). Los Angeles: Natural History Museum of Los Angeles. 79 pp. (Contributions in Science 476.) Google Scholar

    554.

    Gallo, V., and P. M. Coelho. 2008. First occurrence of an aulopiform fish in the Barremian of the Sergipe-Alagoas Basin, northeastern Brazil. In: G. Arratia, H.-P. Schultze, and M. V. H. Wilson, eds. Mesozoic Fishes 4: Homology and Phylogeny. Munich: Verlag Dr. Friedrich Pfeil. pp. 351–371. Google Scholar

    555.

    Gallo, V., and F. J. De Figueiredo. 2002. †Farinichthys gigas, a new albulid fish (Teleostei: Elopomorpha) from the Paleocene of the Pernambuco-Paraíba Basin, northeastern Brazil. Journal of Vertebrate Paleontology 22(4):747–758. Google Scholar

    556.

    Gallo, V., F. J. De Figueiredo, and S. A. Azevedo. 2009. Santanasalmo elegans gen. et sp nov., a basal euteleostean fish from the Lower Cretaceous of the Araripe Basin, northeastern Brazil. Cretaceous Research 30(6):1357–1366. Google Scholar

    557.

    Gardiner, B. G. 1973. Interrelationships of teleostomes. In: P. H. Greenwood, R. S. Miles, and C. Patterson, eds. Interrelationships of Fishes. London: Academic Press. pp. 105–135. Google Scholar

    558.

    Gardiner, B. G. 1984. The relationships of the palaeoniscid fishes: a review based on new specimens of Mimia and Moythomasia from Upper Devonian of Western Australia. Bulletin of the British Museum (Natural History), Geology 37(4):173–428. Google Scholar

    559.

    Gardiner, B. G., J. G. Maisey, and T. J. Littlewood. 1996. Interrelationships of basal neopterygians. In: M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson, eds. Interrelationships of Fishes. San Diego: Academic Press. pp. 117–146. Google Scholar

    560.

    Gardiner, B. G., and B. Schaeffer. 1989. Interrelationships of lower actinopterygian fishes. Zoological Journal of the Linnean Society 97(2):135–187. Google Scholar

    561.

    Gardiner, B. G., B. Schaeffer, and J. A. Masserie. 2005. A review of the lower actinopterygian phylogeny. Zoological Journal of the Linnean Society 144(4):511–525. Google Scholar

    562.

    Garman, S. 1895. The Cyprinodonts. Cambridge, MA: Printed for the Museum. 179 pp. (Memoirs of the Museum of Comparative Zoölogy at Harvard College 19, no. 1.) Google Scholar

    563.

    Garstang, W. 1931. A phyletic classification of Teleostei. Proceedings of the Leeds Philosophical and Literary Society, Scientific Section 2:240–260. Google Scholar

    564.

    Gaudant, J. 1978. Essai de révision taxonomique des “Pholidophorus” (Poissons Actinoptérygiens) du Jurassique supérieur de Cerin (Ain). Nouvelles Archives du Muséum d'Histoire naturelle de Lyon, fascicule 16:101–121. Google Scholar

    565.

    Gaudant, J. 1988. Les cyprinodontiformes (Poissons téléostéens) oligocènes de Ronzon, Le Puy-en-Velay (Haute-Loire): anatomie et signification paléoécologique. Géobios 21(6):773–785. Google Scholar

    566.

    Gaudant, J. 2012. An attempt at the palaeontological history of the European mudminnows (Pisces, Teleostei, Umbridae). Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 263:93–109. Google Scholar

    567.

    Gaudant, M. 1969. Sur quelques nouveaux poisson Bérycoïdes crétacés du Mont Liban. Notes et Mémoires sur le Moyen-Orient 10:273–284. Google Scholar

    568.

    Gaudant, M. 1978a. Contribution à la révision des Poissons crétacés du Jbel Tselfat (Rides prérifaines, Maroc). Première partie: Les Acanthoptérygiens. Notes et Mémoires du Service Géologique du Maroc 39(272):79–124. Google Scholar

    569.

    Gaudant, M. 1978b. Recherches sur l'anatomie et la systématique de Cténothrissiformes et des Pattersonichthyiformes (Poisson Téléostéens) du Cénomanien du Liban. Paris: Éditions du Muséum. 124 pp. (Mémoires du Muséum National d'Histoire Naturelle, Nouvelle Série, Série C, Sciences de la Terre 41.) Google Scholar

    570.

    Gaudant, M. 1979. Recherches sur les relations phylogénétiques de certains poissons Eurypterygii du Crétacé de la Mésogée occidentale. Comptes rendus hebdomadaires des séances de l'Académie des sciences Série D: sciences naturelles 288:1047–1050. Google Scholar

    571.

    Gayet, M. 1980a. Contribution a l'étude anatomique et systématique des poissons cénomaniens du Liban, anciennement placés dans les aeanthopterygiens. Paris: Éditions du Muséum. 149 pp. (Mémoires du Muséum National d'Histoire Naturelle, Nouvelle Série, Série C, Sciences de la Terre 44.) Google Scholar

    572.

    Gayet, M. 1980b. Recherches sur l'ichthyofaune cenomanienne des Monts de Judee: les “acanthopterygiens.” Annales de Paléontologie Vertébrés 66:75–128. Google Scholar

    573.

    Gayet, M. 1988a. Le plus ancien crâne de Siluriforme: Andinichthys bolivianensis nov. gen., nov. sp. (Andinichthyidae nov. fam.) du Maastrichtien de Tiupampa (Bolivie). Comptes rendus de l'Académie des sciences Série 2: Mécanique, physique, chimie, sciences de l'univers, sciences de la terre 307(7)”:833–836. Google Scholar

    574.

    Gayet, M. 1988b. Relations phylogénétiques de Barcarenichthys joneti Gayet du Cénomanien de Barcarena (Portugal) au sein des “Salmoniformes.” Comunicações dos Serviços Geológicos de Portugal 74:85–103. Google Scholar

    575.

    Gayet, M. 1989. Barcarenichthys joneti nov. gen., nov. sp., du Cénomanien marin de Barcarena (Portugal) et ses relations phylogénétiques au sein des “Salmoniformes.” Comptes rendus de l'Académie des sciences Série 2: Mécanique, physique, chimie, sciences de l'univers, sciences de la terre 308(1):137–140. Google Scholar

    576.

    Gayet, M. 1990. Nouveaux Siluriformes du Maastrichtien de Tiupampa (Bolivie). Comptes rendus de l'Académie des sciences Série 2: Mécanique, physique, chimie, sciences de l'univers, sciences de la terre 310(7):867–872. Google Scholar

    577.

    Gayet, M. 1993. Relations phylogénétiques des Gonorynchiformes (Ostariophysi). Belgian Journal of Zoology 123:165–192. Google Scholar

    578.

    Gayet, M. 1994. Fishes from the Lower Cretaceous (Hauterivian?) of Wadi-el-Malih (Israel). Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 194(1):73–94. Google Scholar

    579.

    Gayet, M., M. Jégu, J. Bocquentin, and F. R. Negri. 2003. New characoids from the Upper Cretaceous and Paleocene of Bolivia and the Mio-Pliocene of Brazil: phylogenetic position and paleobiogeographic implications. Journal of Vertebrate Paleontology 23(1):28–46. Google Scholar

    580.

    Gayet, M., and B. Lepicard. 1985. Salmoniforme nouveau du Maastrichtien supérieur des Petites Pyrénées (Haute-Garonne, France): Pyrenichthys jauzaci nov. gen. nov. sp. Bulletin du Muséum National d'Histoire Naturelle, 4e Série, Section C, Sciences de la Terre 7:131–141. Google Scholar

    581.

    Gayet, M., L. G. Marshall, T. Sempere, F. J. Meunier, H. Cappetta, and J. C. Rage. 2001. Middle Maastrichtian vertebrates (fishes, amphibians, dinosaurs and other reptiles, mammals) from Pajcha Pata (Bolivia). Biostratigraphic, palaeoecologic and palaeobiogeographic implications. Palaeogeography, Palaeoclimatology, Palaeoecology 169(1–2):39–68. Google Scholar

    582.

    Gayet, M., and F. J. Meunier. 1991. First discovery of Polypteridae (Pisces, Cladistia, Polypteriformes) outside of Africa = Première découverte de Polypteridae (Pisces, Cladistia, Polypteriformes) hors d'Afrique. Géobios 24(4):463–466. Google Scholar

    583.

    Gayet, M.. 1992. Polyptériformes (Pisces, Cladistia) du maastrichtien et du paléocène de Bolivie = Polyteriformes (Pisces, Cladistia) from Maastrichtian and Paleocene of Bolivia. Géobios 25(Suppl. 1):159–168. Google Scholar

    584.

    Gayet, M.. 1998. Maastrichtian to Early Late Paleocene freshwater Osteichthyes of Bolivia: additions and comments. In: L. R. Malabarba, R. E. Reis, R. P. Vari, Z. M. Lucena, and C. A. S. Lucena, eds. Phylogeny and Classification of Neotropical Fishes. Porto Alegre, Brazil: EDIPUCRS. pp. 85–110. Google Scholar

    585.

    Gayet, M.. 2000. Rectification of the nomenclature of the genus name Ellisella Gayet and Meunier 1991 (Teleostei, Ostarophysi, Gymnotiformes) in Humboldtichthys nom. nov. Cybium 24(1):104. Google Scholar

    586.

    Gayet, M.. 2003. Palaeontology and palaeobiogeography of catfishes. In: G. Arratia, B. G. Kapoor, M. Chardon, and R. Diogo, eds. Catfishes, Volume 2. Enfield, NH: Science Publishers. pp. 491–522. Google Scholar

    587.

    Gayet, M., F. J. Meunier, and F. Kirshbaum 1994. Ellisella kirschbaumi Gayet and Meunier, 1991, Gymnotiforme fossile de Bolivie et ses relations phylogénétiques au sein des formes actuelles. Cybium 18(3):273–306. Google Scholar

    588.

    Gayet, M., F. J. Meunier, and C. Werner. 2002. Diversification in Polypteriformes and special comparison with the Lepisosteiformes. Palaeontology 45(2):361–376. Google Scholar

    589.

    Geoffroy ST. Hilaire, E. 1809. Histoire naturelle des Poissons du Nil. In: France, Commission des sciences et arts d'Egypte. Description de l'Égypte ou recueil des observations et des recherches qui ont été faites en Égypte pendant l'expedition de l'Armée français, publié par les ordres de sa Majesté l'Empereur Napoléon le Grand. Histoire Naturelle 1, part 1. Paris: Imprimerie Impériale. pp. 1–52, plates 18–27. Google Scholar

    590.

    George, A. M., and V. G. Springer. 1980. Revision of the Clinid Fish Tribe Ophiclinini, Including Five New Species, and Definition of the Family Clinidae. Washington, DC: Smithsonian Institution Press. 31 pp. (Smithsonian Contributions to Zoology 307.) Google Scholar

    591.

    Ghedotti, M. J. 2000. Phylogenetic analysis and taxonomy of the poecilioid fishes (Teleostei: Cyprinodontiformes). Zoological Journal of the Linnean Society 130(1):1–53. Google Scholar

    592.

    Ghedotti, M. J., H. M. Dekay, A. J. Maile, W. L. Smith, and M. P. Davis. 2021. Anatomy and evolution of bioluminescent organs in the slimeheads (Teleostei: Trachichthyidae). Journal of Morphology 282(6):820–832. Google Scholar

    593.

    Ghedotti, M. J., J. N. Gruber, R. W. Barton, M. P. Davis, and W. L. Smith. 2018. Morphology and evolution of bioluminescent organs in the glowbellies (Percomorpha: Acropomatidae) with comments on the taxonomy and phylogeny of Acropomatiformes. Journal of Morphology 279(11):1640–1653. Google Scholar

    594.

    Ghezelayagh, A., R. C. Harrington, E. D. Burress, M. Campbell, J. Buckner, P. Chakrabarty, J. R. Glass, W. T. Mccraney, P. Unmack, C. E. Thacker, ET AL. 2022. Prolonged morphological expansion of spiny-rayed fishes following the end-Cretaceous. Nature Ecology and Evolution 6:1211–1220.  https://doi.org/10.1038/s41559-022-01801-3  Google Scholar

    595.

    Gierl, C., M. Dohrmann, P. Keith, W. F. Humphreys, H. R. Esmaeili, J. Vukić, R. Šanda, and B. Reichenbacher. 2022. An integrative phylogenetic approach for inferring relationships of fossil gobioids (Teleostei: Gobiiformes). PLOS ONE 17(7):e0271121.  https://doi.org/10.1371/journal.pone.0271121 Google Scholar

    596.

    Gierl, C., and B. Reichenbacher. 2015. A new fossil genus of Gobiiformes from the Miocene characterized by a mosaic set of characters. Copeia 103(4):792–805. Google Scholar

    597.

    Gierl, C.. 2017. Revision of so-called Pomatoschistus (Gobiiformes, Teleostei) from the Late Eocene and Early Oligocene. Palaeontologia Electronica 20.2.33A:1–17.  https://doi.org/10.26879/721  Google Scholar

    598.

    Gierl, C., B. Reichenbacher, J. Gaudant, D. Erpenbeck, and A. Pharisat. 2013. An extraordinary gobioid fish fossil from Southern France. PLOS ONE 8(5):e64117.  https://doi.org/10.1371/journal.pone.0064117 Google Scholar

    599.

    Gilbert, C. H. 1890. A preliminary report on the fishes collected by the steamer Albatross on the Pacific coast of North America during the year 1889, with descriptions of twelve new genera and ninety-two new species. Proceedings of the United States National Museum 13(879):49–126. Google Scholar

    600.

    Giles, S., K. Feilich, R. C. M. Warnock, S. E. Pierce, and M. Friedman. 2023. A Late Devonian actinopterygian suggests high lineage survivorship across the end-Devonian mass extinction. Nature Ecology and Evolution 7:10–19.  https://doi.org/10.1038/s41559-022-01919-4 Google Scholar

    601.

    Giles, S., G.-H. Xu, T. J. Near, and M. Friedman. 2017. Early members of ‘living fossil’ lineage imply later origin of modern ray-finned fishes. Nature 549:265–268. Google Scholar

    602.

    Gill, A. C. 1996. Comments on an intercalar path for the glossopharyngeal (Cranial IX) nerve as a synapomorphy of the Paracanthopterygii and on the phylogenetic position on the Gobiesocidae (Teleostei: Acanthomorpha). Copeia 1996(4):1022–1029. Google Scholar

    603.

    Gill, A. C. 2013. Classification and relationships of Assiculus and Assiculoides (Teleostei: Pseudochromidae). Zootaxa 3718(2):128–136.  https://doi.org/10.11646/zootaxa.3718.2.2 Google Scholar

    604.

    Gill, A. C., and J. M. Leis. 2019. Phylogenetic position of the fish genera Lobotes, Datnioides and Hapalogenys, with a reappraisal of acanthuriform composition and relationships based on adult and larval morphology. Zootaxa 4680(1):1–81.  https://doi.org/10.11646/zootaxa.4680.1.1 Google Scholar

    605.

    Gill, A. C., and R. D. Mooi. 1993. Monophyly of the Grammatidae and of the Notograptidae, with evidence for their phylogenetic positions among perciforms. Bulletin of Marine Science 52:327–350. Google Scholar

    606.

    Gill, A. C.. 2002. Phylogeny and systematics of fishes. In: P. J. B. Hart and J. D. Reynolds, eds. Handbook of Fish Biology and Fisheries. Malden, MA: Blackwell Science. pp. 15–42. Google Scholar

    607.

    Gill, A. C.. 2012. Thalasseleotrididae, new family of marine gobioid fishes from New Zealand and temperate Australia, with a revised definition of its sister taxon, the Gobiidae (Teleostei: Acanthomorpha). Zootaxa 3266(1):41–52.  https://doi.org/10.11646/zootaxa.3266.1.3 Google Scholar

    608.

    Gill, T. N. 1861a. Catalogue of the fishes of the eastern coast of North America, from Greenland to Georgia. Proceedings of the Academy of Natural Sciences of Philadelphia 13:1–63. Google Scholar

    609.

    Gill, T. N. 1861b. Observations on the genus Cottus, and descriptions of two new species (abridged from the forthcoming report of Capt. J. H. Simpson). Proceedings of the Boston Society of Natural History 8(1861–1862):40–42. Google Scholar

    610.

    Gill, T. N. 1861c. On a new type of aulostomatoids, found in Washington Territory. Proceedings of the Academy of Natural Sciences of Philadelphia 13:168–170. Google Scholar

    611.

    Gill, T. N. 1862. On a new genus of fishes allied to Aulorhynchus and on the affinities of the family Aulorhynchoidae, to which it belongs. Proceedings of the Academy of Natural Sciences of Philadelphia 14:233–235. Google Scholar

    612.

    Gill, T. N. 1872. Arrangement of the Families of Fishes, or Classes Pisces, Marsipobranchii, and Leptocardii. Washington, DC: Smithsonian Institution. 95 pp. (Smithsonian Miscellaneous Collections 11.) Google Scholar

    613.

    Gill, T. N. 1883. Deep-sea fishing fishes. Forest and Stream and Rod and Gun: The American Sportman's Journal 21:284. Google Scholar

    614.

    Gill, T. N. 1884a. Synopsis of the plectognath fishes. Proceedings of the United States National Museum 7(448): 411–427. Google Scholar

    615.

    Gill, T. N. 1884b. Three new families of fishes added to the deep-sea fauna in a year. American Naturalist 18(4):433. Google Scholar

    616.

    Gill, T. N. 1889. On the classification of the mail-cheeked fishes. Proceedings of the United States National Museum 11(756):567–592. Google Scholar

    617.

    Gill, T. N. 1893. Families and subfamiles of fishes. Memoirs of the National Academy of Sciences 6:127–138. Google Scholar

    618.

    Gill, T., and J. A. Ryder. 1883. On the anatomy and relations of the Eurypharyngidae. Proceedings of the United States National Museum 6(382):262–273. Google Scholar

    619.

    Gilliams, J. 1824. Description of a new species of fish of the Linnean genus Perca. Journal of the Academy of Natural Sciences of Philadelphia 4:80–82. Google Scholar

    620.

    Girard, C. 1854. Descriptions of new fishes, collected by Dr. A.L. Heermann, naturalist attached to the Survey of the Pacific Railroad Route, under Lieut. R.S. Williamson, U.S.A. Proceedings of the Academy of Natural Sciences of Philadelphia 7:129–140. Google Scholar

    621.

    Girard, C. 1858a. Fishes. In: U. S. Congress, Senate. Reports of Explorations and Surveys, to Ascertain the Most Practicable and Economical Route for a Railroad from the Mississippi River to the Pacific Ocean, Volume 10, Part 4. 33rd Congress, 2nd session, S. Rep. 33-78, serial 767. Washington, DC: Beverley Tucker. pp. 1–400. Google Scholar

    622.

    Girard, C. 1858b. Notice upon new genera and new species of marine and fresh-water fishes from western North America. Proceedings of the Academy of Natural Sciences of Philadelphia 9:200–202. Google Scholar

    623.

    Girard, M. G., M. P. Davis, C. C. Baldwin, A. Dettaï, R. P. Martin, and W. L. Smith. 2022. Molecular phylogeny of the threadfin fishes (Polynemidae) using ultraconserved elements. Journal of Fish Biology 100(3):793–810. Google Scholar

    624.

    Girard, M. G., M. P. Davis, and W. L. Smith. 2020. The phylogeny of carangiform fishes: morphological and genomic investigations of a new fish clade. Copeia 108(2):265–298. Google Scholar

    625.

    Girard, M. G., M. P. Davis, H. H. Tan, D. J. Wedd, P. Chakrabarty, W. B. Ludt, A. P. Summers, and W. L. Smith. 2022. Phylogenetics of archerfishes (Toxotidae) and evolution of the toxotid shooting apparatus. Integrative Organismal Biology 4(1):obac013.  https://doi.org/10.1093/iob/obac013 Google Scholar

    626.

    Girard, M. G., B. C. Mundy, A. Nonaka, and G. D. Johnson. 2023. Cusk-eel confusion: revisions of larval Lucio-brotula and Pycnocraspedum and re-descriptions of two bythitid larvae (Ophidiiformes). Ichthyological Research 70:474–489. Google Scholar

    627.

    Glass, J. R., R. C. Harrington, P. F. Cowman, B. C. Faircloth, and T. J. Near. 2023. Widespread sympatry in a species-rich clade of marine fishes (Carangoidei). Proceedings of the Royal Society B, Biological Sciences 290(2010):20230657. Google Scholar

    628.

    Gmelin, J. F. 1789. Caroli a Linné…Systema Naturae per regna tria naturae, secundum classes, ordines, genera, species, cum characteribus, differentiis, synonymis, locis. 13th ed., Volume 1, Part 3. Leipzig: Beer. 484 pp. Google Scholar

    629.

    Goatley, C. H. R., and L. Tornabene. 2022. Tempestichthys bettyae, a new genus and species of ocean sleeper (Gobiiformes, Thalasseleotrididae) from the central Coral Sea. Systematics and Biodiversity 20(1):1–15. Google Scholar

    630.

    Godkin, C. M., and R. Winterbottom. 1985. Phylogeny of the family Congrogadidae (Pisces, Perciformes) and its placement as a subfamily of the Pseudochromidae. Bulletin of Marine Science 36(3):633–671. Google Scholar

    631.

    Goode, G. B., and T. H. Bean. 1879. Description of Alepocephalus bairdii, a new species of fish from the deep-sea fauna of the western Atlantic. Proceedings of the United States National Museum 2(68):55–57. Google Scholar

    632.

    Goode, G. B.. 1883. Reports on the results of dredging under the supervision of Alexander Agassiz, on the east coast of the United States, during the summer of 1880, by the U. S. coast survey steamer “Blake,” Commander J. R. Bartlett, U.S.N., commanding. Bulletin of the Museum of Comparative Zoology 10:183–226. Google Scholar

    633.

    Goode, G. B.. 1896. Oceanic Ichthyology: A treatise on the deep-sea and pelagic fishes of the world, based chiefly upon the collections made by the steamers “Blake,” “Albatross,” and “Fishhawk” in the northwestern Atlantic, with an atlas containing 417 figures. 2 vols. Cambridge, MA: Printed for the Museum. Published in connection with the National Museum and the Smithsonian Institution. (Memoirs of the Museum of Comparitive Zoölogy at Harvard College 22.) Google Scholar

    634.

    Goodrich, E. S. 1909. Vertebrata Craniata (first fascicle: cyclostomes and fishes). Volume 9 of A Treatise on Zoology, edited by R. Lankester. London: Adam and Charles Black. 518 pp. Google Scholar

    635.

    Goodrich, E. S. 1928. Polypterus a palaeoniscid? Palaeobiologica 1:87–92. Google Scholar

    636.

    Goodrich, E. S. 1930. Studies on the Structure and Developtment of Vertebrates. London: Macmillan. 837 pp. Google Scholar

    637.

    Gosline, W. A. 1960. Contributions toward a classification of modern isospondylous fishes. Bulletin of the British Museum (Natural History), Zoology 6:327–365. Google Scholar

    638.

    Gosline, W. A. 1961. Some osteological features of modern lower teleostean fishes. Smithsonian Miscellaneous Collections 142(3):1–42. Google Scholar

    639.

    Gosline, W. A. 1963a. Considerations Regarding the Relationships of the Percopsiform, Cyprinodontiform, and Gadiform Fishes. Ann Arbor: University of Michigan Museum of Zoology. 38 pp. (Occasional Papers 629.) Google Scholar

    640.

    Gosline, W. A. 1963b. Notes on the osteology and systematic position of Hypoptychus dybowskii Steindachner and other elongate perciform fishes. Pacific Science 17(1):90–101. Google Scholar

    641.

    Gosline, W. A. 1965. Teleostean phylogeny. Copeia 1965(2):186–194. Google Scholar

    642.

    Gosline, W. A. 1968. The suborders of perciform fishes. Proceedings of the United States National Museum 124(3647):1–78. Google Scholar

    643.

    Gosline, W. A. 1969. The morphology and systematic position of the alepocephaloid fishes. Bulletin of the British Museum (Natural History), Zoology 18:187–218. Google Scholar

    644.

    Gosline, W. A. 1971. Functional Morphology and Classification of Teleostean Fishes. Honolulu: University Press of Hawaii. 208 pp. Google Scholar

    645.

    Gosline, W. A. 1983. The relationships of the mastacembelid and synbranchid fishes. Japanese Journal of Ichthyology 29(4):323–328. Google Scholar

    646.

    Gosline, W. A. 1985. Relationships among some relatively deep bodied percoid fish groups. Japanese Journal of Ichthyology 31(4):351–357. Google Scholar

    647.

    Gosline, W. A., N. B. Marshall, and G. W. Mead. 1966. Order Iniomi: characters and synopsis of the families. In: C. W. Mead, ed. Orders Iniomi and Lyomeri: Lizardfishes, Other Iniomi, Deepsea Gulpers. New Haven, CT: Sears Foundation for Marine Research, Yale University. pp. 1–18. (Memoir 1, Fishes of the Western North Atlantic 5.) Google Scholar

    648.

    Gottfried, M. D., R. E. Fordyce, and S. Rust. 2006. Megalampris keyesi, a giant moonfish (Teleostei, Lampridiformes) from the late oligocene of New Zealand. Journal of Vertebrate Paleontology 26(3):544–551. Google Scholar

    649.

    Gouiric-Cavalli, S., and G. Arratia. 2022. A new †Pachycormiformes (Actinopterygii) from the Upper Jurassic of Gondwana sheds light on the evolutionary history of the group. Journal of Systematic Palaeontology 19(21):1517–1550. Google Scholar

    650.

    Gradstein, F. M., and J. G. Ogg. 2020. The chronostratigraphic scale. In: F. M. Gradstein, J. G. Ogg, M. D. Schmitz, and G. M. Ogg, eds. Geologic Time Scale 2020, Volume 1. Amsterdam: Elsevier. pp. 21–32. Google Scholar

    651.

    Grand, A., R. Zaragüeta Bagils, L. M. Velez, and V. Ung. 2014. A cladistic re-analysis of the Gadiformes (Teleostei, Paracanthopterygii) using three-item analysis. Zootaxa 3889(4):525–552.  https://doi.org/10.11646/zootaxa.3889.4.3 Google Scholar

    652.

    Grande, L. 1982. A revision of the fossil genus Diplomystus, with comments on the interrelationships of clupeomorph fishes. American Museum Novitates 2728:1–38. Google Scholar

    653.

    Grande, L. 1984. Paleontology of the Green River Formation, with a Review of the Fish Fauna. Laramie, WY: Geological Survey of Wyoming. xvii, 333 pp. (Bulletin 63.) Google Scholar

    654.

    Grande, L. 1985. Recent and fossil clupeomorph fishes with material for revision of the subgroups of clupeoids. Bulletin of the American Museum of Natural History 181(art. 2):231–272. Google Scholar

    655.

    Grande, L. 1987. Redescription of †Hypsidoris farsonensis (Teleostei: Siluriformes), with a reassessment of its phylogenetic relationships. Journal of Vertebrate Paleontology 7(1):24–54. Google Scholar

    656.

    Grande, L. 1988. A well preserved paracanthopterygian fish (Teleostei) from freshwater Lower Paleocene deposits of Montana. Journal of Vertebrate Paleontology 8(2):117–130. Google Scholar

    657.

    Grande, L. 1999. The first Esox (Esocidae: Teleostei) from the Eocene Green River Formation, and a brief review of esocid fishes. Journal of Vertebrate Paleontology 19(2):271–292. Google Scholar

    658.

    Grande, L. 2010. An empirical and synthetic pattern study of gars (Lepisosteiformes) and closely related species, based mostly on skeletal anatomy: the resurrection of Holostei. Lawrence, KS: American Society of Ichthyologists and Herpetologists. 871 pp. (Special Publication 6.) Google Scholar

    659.

    Grande, L., and W. E. Bemis. 1991. Osteology and phylogenetic relationships of fossil and recent paddlefishes (Polyodontidae) with comments on the interrelationships of Acipenseriformes. Journal of Vertebrate Paleontology 11(1 Suppl.):1–121. (Society of Vertebrate Paleontology Memoir 1.) Google Scholar

    660.

    Grande, L.. 1996. Interrelationships of Acipenseriformes, with comments on “Chondrostei.” In: M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson, eds. Interrelationships of Fishes. San Diego: Academic Press. pp. 85–115. Google Scholar

    661.

    Grande, L.. 1998. A comprehensive phylogenetic study of amiid fishes (Amiidae) based on comparative skeletal anatomy. An empirical search for interconnected patterns of natural history. Journal of Vertebrate Paleontology 18(1 Suppl.):1–690. (Society of Vertebrate Paleontology Memoir 4.) Google Scholar

    662.

    Grande, L., and M. C. C. De Pinna. 1998. Description of a second species of the catfish †Hypsidoris and a reevaluation of the genus and the family †Hypsidoridae. Journal of Vertebrate Paleontology 18(3):451–474. Google Scholar

    663.

    Grande, L., and T. Grande. 1999. A new species of Notogoneus (Teleostei: Gonorynchidae) from the Upper Cretaceous Two Medicine Formation of Montana, and the poor Cretaceous record of freshwater fishes from North America. Journal of Vertebrate Paleontology 19(4):612–622. Google Scholar

    664.

    Grande, L., and E. J. Hilton. 2006. An exquisitely preserved skeleton representing a primitive sturgeon from the Upper Cretaceous Judith River Formation of Montana (Acipenseriformes: Acipenseridae: n. gen. and sp). Journal of Paleontology 80(4 Suppl.):1–39. (Paleontological Society Memoir 65.) Google Scholar

    665.

    Grande, T. 1994. Phylogeny and Paedomorphosis in an African Family of Freshwater Fishes (Gonorynchiformes: Kneriidae). Chicago: Field Museum of Natural History. 20 pp. (Fieldiana Zoology New Series 78.) Google Scholar

    666.

    Grande, T. 1996. The interrelationships of fossil and Recent gonorynchid fishes with comments on two Cretaceous taxa from Israel. In: G. Arratia and G. Viohl, eds. Mesozoic Fishes: Systematics and Paleoecology. Munich: Verlag Dr. Friedrich Pfeil. pp. 299–318. Google Scholar

    667.

    Grande, T. 1999. Revision of the genus Gonorynchus Scopoli, 1777 (Teleostei: Ostariophysi). Copeia 1999(2):453–469. Google Scholar

    668.

    Grande, T., W. C. Borden, and W. L. Smith. 2013. Limits and relationships of Paracanthopterygii: a molecular framework for evaluating past morphological hypotheses. In: G. Arratia, H.-P. Schultze, and M. V. H. Wilson, eds. Mesozoic Fishes 5: Global Diversity and Evolution. Munich: Verlag Dr. Fredrich Pfeil. pp. 385–418. Google Scholar

    669.

    Grande, T., W. C. Borden, M. V. H. Wilson, and L. Scarpitta. 2018. Phylogenetic relationships among fishes in the order Zeiformes based on molecular and morphological data. Copeia 106(1):20–48. Google Scholar

    670.

    Grande, T., and M. De Pinna. 2004. The evolution of the Weberian apparatus: a phylogenetic perspective. In: G. Arratia and A. Tintori, eds. Mesozoic Fishes 3: Systematics, Paleoenvironments and Biodiversity. Munich: Verlag Dr. Friedrich Pfeil. pp. 429–448. Google Scholar

    671.

    Grande, T., H. Laten, and J. A. López. 2004. Phylogenetic relationships of extant esocid species (Teleostei: Salmoniformes) based on mophological and molecular characters. Copeia 2004(4):743–757. Google Scholar

    672.

    Grande, T., and F. J. Poyato-Ariza. 1995. A cladistic analysis of fossil and living gonorynchiform ostariophysan fishes. Géobios 28(Suppl. 2):197–199. Google Scholar

    673.

    Grande, T.. 1999. Phylogenetic relationships of fossil and Recent gonorynchiform fishes (Teleostei: Ostariophysi). Zoological Journal of the Linnean Society 125(2):197–238. Google Scholar

    674.

    Gray, J. E. 1831. Description of twelve new genera of fish, discovered by Gen. Hardwicke, in India, the greater part in the British Museum. Zoological Miscellany 1831:7–9. Google Scholar

    675.

    Greenfield, D. W., R. Winterbottom, and B. B. Collette. 2008. Review of the toadfish genera (Teleostei: Batrachoididae). Proceedings of the California Academy of Sciences 59(15):665–710. Google Scholar

    676.

    Greenwood, P. H. 1960. Fossil denticipitid fishes from East Africa. Bulletin of the British Museum (Natural History), Geology 5(1):1–11. Google Scholar

    677.

    Greenwood, P. H. 1963. The swimbladder in African Notopteridae (Pisces) and its bearing on the taxonomy of the family. Bulletin of the British Museum (Natural History), Zoology 11(5):379–412. Google Scholar

    678.

    Greenwood, P. H. 1968. The osteology and relationships of the Denticipitidae: a family of clupeomorph fishes. Bulletin of the British Museum (Natural History), Zoology 16(6):215–273. Google Scholar

    679.

    Greenwood, P. H. 1973. Interrelationships of osteoglossomorphs. In: P. H. Greenwood, R. S. Miles, and C. Patterson, eds. Interrelationships of Fishes. London: Academic Press. pp. 307–332. Google Scholar

    680.

    Greenwood, P. H. 1977. Notes on the anatomy and classification of elopomorph fishes. Bulletin of the British Museum (Natural History), Zoology 32(4):65–102. Google Scholar

    681.

    Greenwood, P. H. 1995. A revised familial classification for certain cirrhitoid genera (Teleostei, Percoidei Cirrhitoidea), with comments on the group's monophyly and taxonomic ranking. Bulletin of the Natural History Museum (Zoology) 61(1):1–10. Google Scholar

    682.

    Greenwood, P. H., G. S. Myers, D. E. Rosen, and S. H. Weitzman. 1967. Named main divisions of teleostean fishes. Proceedings of the Biological Society of Washington 80:227–228. Google Scholar

    683.

    Greenwood, P. H., and D. E. Rosen. 1971. Notes on the structure and relationships of alepocephaloid fishes. American Museum Novitates 2473:1–41. Google Scholar

    684.

    Greenwood, P. H., D. E. Rosen, S. H. Weitzman, and G. S. Myers. 1966. Phyletic studies of teleostean fishes, with a provisional classification of living forms. Bulletin of the American Museum of Natural History 131(art.4):341–455. Google Scholar

    685.

    Gregorová, R. 2011. Fossil fish fauna (Teleostei, Selachii) from the Dynów marlstone (Rupelian, NP 23) of the Menilitic Formation at the locality of Litenčice (Czech Republic). Acta Musei Moraviae, Scientiae Geologicae 96:3–33. Google Scholar

    686.

    Gregory, W. K. 1907. The orders of teleostomus fishes. Annals of the New York Academy of Sciences 17(3):437–508. Google Scholar

    687.

    Gregory, W. K. 1933. Fish skulls: a study of the evolution of natural mechanisms. Transactions of the American Philosophical Society 23(2):75–481. Google Scholar

    688.

    Gregory, W. K., and G. M. Conrad. 1936. Pictoral phylogenies of deep sea Isospondyli and Iniomi. Copeia 1936(1):21–36. Google Scholar

    689.

    Guichenot, A. 1848. Fauna Chilena. Peces. In: C. Gay, ed. Historia fisica y politica de Chile. Zoología 2. Paris: Museo de historia natural de Santiago. pp. 137–370. Google Scholar

    690.

    Guinot, G., and L. Cavin. 2018. Body size evolution and habitat colonization across 100 million years (Late Jurassic–Paleocene) of the actinopterygian evolutionary history. Fish and Fisheries 19(4):577–597. Google Scholar

    691.

    Günther, A. 1859. Catalogue of the acanthopterygian fishes in the collection of the British Museum, Volume 1. London: British Museum. xxxi, 524 pp. Google Scholar

    692.

    Günther, A. 1861. Catalogue of the Acanthopterygian Fishes in the Collection of the British Museum, Volume 3. London: British Museum. 586 pp. Google Scholar

    693.

    Günther, A. 1863a. On new species of fishes from the Essequibo. Annals and Magazine of Natural History, Series 3, 12(72):441–443. Google Scholar

    694.

    Günther, A. 1863b. On new species of fishes from Victoria, South Australia. Annals and Magazine of Natural History, Series 3, 12(68):114–117. Google Scholar

    695.

    Günther, A. 1864a. Catalogue of the Fishes in the British Museum, Volume 5. London: British Museum. 455 pp. Google Scholar

    696.

    Günther, A. 1864b. Report of a collection of fishes made by Messrs. Dow, Godman, and Salvin in Guatemala. Proceedings of the Zoological Society of London 1864, pt. 1 (art. 3), pp. 144–154. Google Scholar

    697.

    Günther, A. 1868. Catalogue of the Fishes in the British Museum, Volume 7. London: British Museum. 512 pp. Google Scholar

    698.

    Günther, A. 1870. Catalogue of the Fishes in the British Museum, Volume 8. London: British Museum. 549 pp. Google Scholar

    699.

    Günther, A. 1878a. Preliminary notices of deep-sea fishes collected during the voyage of the H.M.S. “Challenger.” Annals and Magazine of Natural History, Series 5, 2(7):17–28. Google Scholar

    700.

    Günther, A. 1878b. Preliminary notices of deep-sea fishes collected during the voyage of the H.M.S. “Challenger.” Annals and Magazine of Natural History, Series 5, 2(9):248–251. Google Scholar

    701.

    Günther, A. 1880. An Introduction to the Study of Fishes. Edinburgh: Adam and Charles Black. 720 pp. Google Scholar

    702.

    Gushiken, S. 1988. Phylogenetic relationships of the perciform genera of the family Carangidae. Japanese Journal of Ichthyology 34(4):443–461. Google Scholar

    703.

    Haeckel, E. 1866. Generalle morphologie der organismen, Volume 2. Berlin: G. Reimer. clx, 462 pp. Google Scholar

    704.

    Haedrich, R. L. 1967. The stromateoid fishes: systematics and a classification. Bulletin of the Museum of Comparative Zoology 135(2):31–139. Google Scholar

    705.

    Hamilton, F. 1822. An Account of the Fishes Found in the River Ganges and Its Branches. Edinburgh: A. Constable and Company. 405 pp. Google Scholar

    706.

    Hamilton, H., N. Saarman, G. Short, A. B. Sellas, B. Moore, T. Hoang, C. L. Grace, M. Gomon, K. Crow, and W. B. Simison. 2017. Molecular phylogeny and patterns of diversification in syngnathid fishes. Molecular Phylogenetics and Evolution 107:388–403. Google Scholar

    707.

    Han, Z., C. Shou, M. Liu, and T. Gao. 2021. Large-scale phylogenomic analysis provides new insights into the phylogeny of the order Gadiformes and evolution of freshwater gadiform species Burbot (Lota lota) [preprint]. Preprints.org[online database].  https://doi.org/10.20944/preprints202106.0610.v1  Google Scholar

    708.

    Hao, C., Y. Liu, N. Wei, K. Arken, C. Shi, and C. Yue. 2023. The complete mitochondrial genomes of the Leuciscus baicalensis and Rutilus rutilus: a detailed genomic comparison among closely related species of the Leuciscinae subfamily. Gene 877:147535. Google Scholar

    709.

    Hao, S., K. Han, L. Meng, X. Huang, W. Cao, C. Shi, M. Zhang, Y. Wang, Q. Liu, Y. Zhang, H. Sun, I. Seim, X. Xu, X. Liu, and G. Fan. 2020. African Arowana genome provides insights on ancient teleost evolution. iScience 23(11):101662.  https://doi.org/10.1016/j.isci.2020.101662 Google Scholar

    710.

    Hardman, M. 2005. The phylogenetic relationships among non-diplomystid catfishes as inferred from mitochondrial cytochrome b sequences; the search for the ictalurid sister taxon (Otophysi: Siluriformes). Molecular Phylogenetics and Evolution 37(3):700–720. Google Scholar

    711.

    Harold, A. S. 1993. Phylogenetic relationships of the sternoptychid Argyropelecus (Teleostei: Stomiiformes). Copeia 1993(1):123–133. Google Scholar

    712.

    Harold, A. S. 1994. A taxonomic revision of the sternoptychid genus Polyipnus (Teleostei, Stromiiformes) with an analysis of phylogenetic relationships. Bulletin of Marine Science 54(2):428–534. Google Scholar

    713.

    Harold, A. S. 1998. Phylogenetic relationships of the Gonostomatidae (Teleostei: Stomiiformes). Bulletin of Marine Science 62(3):715–741. Google Scholar

    714.

    Harold, A. S., and S. H. Weitzman. 1996. Interrelationships of stomiiform fishes. In: M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson, eds. Interrelationships of Fishes. San Diego: Academic Press. pp. 333–353. Google Scholar

    715.

    Harrington, R. C., B. C. Faircloth, R. I. Eytan, W. L. Smith, T. J. Near, M. E. Alfaro, and M. Friedman. 2016. Phylogenomic analysis of carangimorph fishes reveals flatfish asymmetry arose in a blink of the evolutionary eye. BMC Evolutionary Biology 16:224.  https://doi.org/10.1186/s12862-016-0786-x Google Scholar

    716.

    Harrington, R. C., M. Friedman, M. Miya, T. J. Near, and M. A. Campbell. 2021. Phylogenomic resolution of the monotypic and enigmatic Amarsipus, the Bagless Glassfish (Teleostei, Amarsipidae). Zoologica Scripta 50(4):411–422. Google Scholar

    717.

    Harrington, R. C., M. A. Kolmann, J. J. Day, B. C. Faircloth, M. Friedman, and T. J. Near. 2023. Dispersal sweepstakes: biotic interchange propelled air-breathing fishes across the globe. Journal of Biogeography. Published ahead of print, 27 December.  https://doi.org/10.1111/jbi.14781  Google Scholar

    718.

    Hart, P. B., R. J. Arnold, F. Alda, C. P. Kenaley, T. W. Pietsch, D. Hutchinson, and P. Chakrabarty. 2022. Evolutionary relationships of anglerfishes (Lophiiformes) reconstructed using ultraconserved elements. Molecular Phylogenetics and Evolution 171:107459. Google Scholar

    719.

    Hart, P. B., M. L. Niemiller, E. D. Burress, J. W. Armbruster, W. B. Ludt, and P. Chakrabarty. 2020. Cave-adapted evolution in the North American amblyopsid fishes inferred using phylogenomics and geometric morphometrics. Evolution 74(5):936–949. Google Scholar

    720.

    Hartel, K. E., and M. L. J. Stiassny. 1986. The identification of larval Parasudis (Teleostei, Chlorophthalmidae); with notes on the anatomy and relationships of aulopiform fishes. Breviora 487:1–23. Google Scholar

    721.

    Harvey, V. L., J. N. Keating, and M. Buckley. 2021. Phylogenetic analyses of ray-finned fishes (Actinopterygii) using collagen type I protein sequences. Royal Society Open Science 8(8):201955.  https://doi.org/10.1098/rsos.201955 Google Scholar

    722.

    Hastings, P. A. 1993. Relationships of fishes of the perciform suborder Notothenioidei. In: R. G. Miller, ed. A History and Atlas of the Fishes of the Antarctic Ocean. Carson City, NV: Foresta Institute for Ocean and Mountain Studies. pp. 99–107. Google Scholar

    723.

    Hastings, P. A., and V. G. Springer. 2009. Systematics of the Blennioidei and the included families Dactyloscopidae, Chaenopsidae, Clinidae and Labrisomidae. In: R. A. Patzner, E. J. Goncalves, P. A. Hastings, and B. G. Kapoor, eds. The Biology of Blennies. Enfield, NH: Science Publishers. pp. 3–30. Google Scholar

    724.

    Hay, O. P. 1903. Some remarks on the fossil fishes of Mount Lebanon, Syria. American Naturalist 37(442):685–695. Google Scholar

    725.

    Hay, O. P. 1929. Second Bibliography and Catalogue of the Fossil Vertebrata of North America, Volume 1. Washington, DC: Carnegie Institution of Washington. 916 pp. Google Scholar

    726.

    Haÿ, V., M. I. Mennesson, P. Keith, and C. Lord. 2022. Revision of the genus Rhyacichthys using integrative taxonomy. Pacific Science 76(2):123–137. Google Scholar

    727.

    He, S. P., X. Gu, R. L. Mayden, W.-J. Chen, K. W. Conway, and Y. Y. Chen. 2008. Phylogenetic position of the enigmatic genus Psilorhynchus (Ostariophysi: Cypriniformes): evidence from the mitochondrial genome. Molecular Phylogenetics and Evolution 47(1):419–425. Google Scholar

    728.

    He, S. P., R. L. Mayden, X. Z. Wang, W. Wang, K. L. Tang, W.-J. Chen, and Y. Y. Chen. 2008. Molecular phylogenetics of the family Cyprinidae (Actinopterygii: Cypriniformes) as evidenced by sequence variation in the first intron of S7 ribosomal protein-coding gene: further evidence from a nuclear gene of the systematic chaos in the family. Molecular Phylogenetics and Evolution 46(3):818–829. Google Scholar

    729.

    Heckel, J. J. 1853. Berichtüber die vom Herrn Cavaliere Achille de Zigno hier angelangte Sammlung fossiler Fische. Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften, Mathematisch-Naturwissenschaftliche Classe 11:122–138. Google Scholar

    730.

    Hector, J. 1871. On a new species of fish, Coryphaenoides novae Zelandiae; native name, Okarari. Transactions and Proceedings of the New Zealand Institute 3(21):136. Google Scholar

    731.

    Heemstra, P. C., and T. Hecht. 1986. Dinopercidae, a new family for the percoid marine fish genera Dinoperca Boulenger and Centrarchops Fowler (Pisces: Perciformes). Grahamstown, South Africa: J. L. B. Smith Institute of Ichthyology. 21 pp. (Bulletin 51.) Google Scholar

    732.

    Helmstetter, A. J., A. S. T. Papadopulos, J. Igea, T. J. M. Van Dooren, A. M. Leroi, and V. Savolainen. 2016. Viviparity stimulates diversification in an order of fish. Nature Communications 7:11271. Google Scholar

    733.

    Hensel, K. 1970. Review of the classification and of the opinions on the evolution of Cyprinoidei (Eventognathi) with an annotated list of genera and subgenera described since 1921. Annotationes Zoologicae et Botanice 57:1–45. Google Scholar

    734.

    Hensley, D. A. 1997. An overview of the systematics and biogeography of the flatfishes. In: Proceedings of the Third International Symposium on Flatfish Ecology, Part I; 2–8 November 1996, Texel, The Netherlands. Special issue, Journal of Sea Research 37(3–4):187–194. Google Scholar

    735.

    Hensley, D. A., and E. H. Ahlstrom. 1984. Pleuronectiformes: relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 670–687. (Special Publication 1.) Google Scholar

    736.

    Heras, S., and M. I. Roldán. 2011. Phylogenetic inference in Odontesthes and Atherina (Teleostei: Atheriniformes) with insights into ecological adaptation. Comptes Rendus Biologies 334(4):273–281. Google Scholar

    737.

    Herler, J., S. Koblmüller, and C. Sturmbauer. 2009. Phylogenetic relationships of coral-associated gobies (Teleostei, Gobiidae) from the Red Sea based on mitochondrial DNA data. Marine Biology 156:725–739. Google Scholar

    738.

    Hermann, J. 1781. Schreiben an den Herausgeber über eine neues amerikanisches Fischgeschlecht, Sternoptyx diaphana, der durchsichtige Brust-Falten-Fisch. Naturforscher 16:8–36. Google Scholar

    739.

    Hernández-Guerrero, C., K. M. Cantalice, K. A. González-Rodríguez, and V. M. Bravo-Cuevas. 2021. The first record of a pterothrissin (Albuliformes, Albulidae) from the Muhi Quarry, mid-Cretaceous (Albian-Cenomanian) of Hidalgo, central Mexico. Journal of South American Earth Sciences 107:103032. Google Scholar

    740.

    Hertwig, S. T. 2008. Phylogeny of the Cyprinodontiformes (Teleostei, Atherinomorpha): the contribution of cranial soft tissue characters. Zoologica Scripta 37(2):141–174. Google Scholar

    741.

    Hidaka, K., Y. Tsukamoto, and Y. Iwatsuki. 2017. Nemoossis, a new genus for the eastern Atlantic long-fin bonefish Pterothrissus belloci Cadenat 1937 and a redescription of P. gissu Hilgendorf 1877 from the northwestern Pacific. Ichthyological Research 64:45–53. Google Scholar

    742.

    Hilgendorf, F. M. 1877. Pterothrissus, eine neue Clupeidengattung. Leopoldina, Amtliches Organ der Kaiserlich Leopoldinisch-Carolinisch-Deutschen Akademie der Naturforscher 13:127–128. Google Scholar

    743.

    Hilgendorf, F. M. 1878. Über das Vorkommen einer Brama-Art und einer neuen Fischgattung Centropholis aus der Nachbarschaft des Genus Brama in den japanischen Meeren. Sitzungsberichte der Gesellschaft Naturforschender Freunde zu Berlin 1878:1–2. Google Scholar

    744.

    Hilton, E. J. 2003. Comparative osteology and phylogenetic systematics of fossil and living bony-tongue fishes (Actinopterygii, Teleostei, Osteoglossomorpha). Zoological Journal of the Linnean Society 137(1):1–100. Google Scholar

    745.

    Hilton, E. J. 2009. Osteology of the Graveldiver Scytalina cerdale (Perciformes: Zoarcoidei: Scytalinidae). Journal of Morphology 270(12):1475–1491. Google Scholar

    746.

    Hilton, E. J. 2022. Origin and diversity of early teleostean fishes. In: P. C. Heemstra, E. Heemstra, D. A. Ebert, W. Hollerman, and J. E. Randall, eds. Coastal Fishes of the Western Indian Ocean, Volume 2. Makhanda, South Africa: South African Institute for Aquatic Biodiversity. pp. 4–19. Google Scholar

    747.

    Hilton, E. J., and R. Britz. 2010. The caudal skeleton of osteglossomorph fishes, revisited: comparisons, homologies, and characters. In: J. S. Nelson, H.-P. Schultze, and M. V. H. Wilson, eds. Origin and Phylogenetic Interrelationships of Teleosts. Munich: Verlag Dr. Friedrich Pfeil. pp. 219–237. Google Scholar

    748.

    Hilton, E. J., and P. L. Forey. 2009. Redescription of †Chondrosteus acipenseroides Egerton, 1858 (Acipenseriformes, †Chondrosteidae) from the Lower Lias of Lyme Regis (Dorset, England), with comments on the early evolution of sturgeons and paddlefishes. Journal of Systematic Palaeontology 7(4):427–453. Google Scholar

    749.

    Hilton, E. J., L. Grande, and W. E. Bemis. 2011. Skeletal anatomy of the Shortnose Sturgeon, Acipenser brevirostrum Lesueur, 1818, and the systematics of sturgeons (Acipenseriformes, Acipenseridae). Chicago: Field Museum of Natural History. 168 pp. (Fieldiana: Life and Earth Sciences 3.) Google Scholar

    750.

    Hilton, E. J., P. Konstantinidis, and A. Williston. 2021. Osteology of Parabrotula plagiophthalmus (Ophidiiformes: Bythitoidei: Bythitidae). Breviora 569:1–18. Google Scholar

    751.

    Hilton, E. J., and S. Lavoué. 2018. A review of the systematic biology of fossil and living bony-tongue fishes, Osteoglossomorpha (Actinopterygii: Teleostei). Neotropical Ichthyology 16(3):e180031. Google Scholar

    752.

    Hilton, E. J., D. E. Stevenson, and A. C. Matarese. 2019. Osteology of Ronquilus jordani (Zoarcoidei: Bathymasteridae), with a discussion of the developmental osteology and systematics of bathymasterid fishes. Acta Zoologica 100(4):389–407. Google Scholar

    753.

    Hinegardner, R., and D. E. Rosen. 1972. Cellular DNA content and the evolution of teleostean fishes. American Naturalist 106(951):621–644. Google Scholar

    754.

    Hirt, M. V., G. Arratia, W.-J. Chen, R. L. Mayden, K. L. Tang, R. M. Wood, and A. M. Simons. 2017. Effects of gene choice, base composition and rate heterogeneity on inference and estimates of divergence times in cypriniform fishes. Biological Journal of the Linnean Society 121(2):319–339. Google Scholar

    755.

    Hoese, D. F. 1984. Gobioidei: relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 588–591. (Special Publication 1.) Google Scholar

    756.

    Hoese, D. F., and A. C. Gill. 1993. Phylogenetic relationships of eleotridid fishes (Perciformes: Gobioidei). Bulletin of Marine Science 52:415–440. Google Scholar

    757.

    Hoffmann, M., and R. Britz. 2006. Ontogeny and homology of the neural complex of otophysan Ostariophysi. Zoological Journal of the Linnean Society 147(3):301–330. Google Scholar

    758.

    Hoganson, J., J. M. Erickson, and F. Holland. 2019. Chondrichthyan and Osteichthyan Paleofaunas of the Cretaceous (Late Maastrichtian) Fox Hills Formation of North Dakota, USA: Paleoecology, Paleogeography, and Extinction. Ithaca, NY: Paleontological Research Institution. 94 pp. (Bulletins of American Paleontology 398.) Google Scholar

    759.

    Holcroft, N. I. 2004. A molecular test of alternative hypotheses of tetraodontiform (Acanthomorpha: Tetraodontiformes) sister group relationships using data from the RAG1 gene. Molecular Phylogenetics and Evolution 32(3):749–760. Google Scholar

    760.

    Holcroft, N. I. 2005. A molecular analysis of the interrelationships of tetraodontiform fishes (Acanthomorpha: Tetraodontiformes). Molecular Phylogenetics and Evolution 34(3):525–544. Google Scholar

    761.

    Holcroft, N. I., and E. O. Wiley. 2008. Acanthuroid relationships revisited: a new nuclear gene-based analysis that incorporates tetraodontiform representatives. Ichthyological Research 55:274–283. Google Scholar

    762.

    Hora, S. L. 1925. Notes on fishes of the Indian Museum. XII. The systematic position of cyprinoid genus Psilorhynchus McClelland. Records of the Indian Museum 27:457–460. Google Scholar

    763.

    Horn, M. H. 1984. Stromateoidei: development and relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 620–628. (Special Publication 1.) Google Scholar

    764.

    Hoshino, K. 2001. Monophyly of the Citharidae (Pleuronectoidei: Pleuronectiformes: Teleostei) with considerations of pleuronectoid phylogeny. Ichthyological Research 48:391–404. Google Scholar

    765.

    Hotaling, S., M. L. Borowiec, L. S. F. Lins, T. Desvignes, and J. L. Kelley. 2021. The biogeographic history of eelpouts and related fishes: linking phylogeny, environmental change, and patterns of dispersal in a globally distributed fish group. Molecular Phylogenetics and Evolution 162:107211.  https://doi.org/10.1016/j.ympev.2021.107211 Google Scholar

    766.

    Howes, G. J. 1983. The cranial muscles of the loricarioid catfishes, their homologies and value as taxonomic characters (Teleostei: Siluroidei). Bulletin of the British Museum (Natural History), Zoology 45(6):309–345. Google Scholar

    767.

    Howes, G. J. 1989. Phylogenetic relationships of macrouroid and gadoid fishes based on cranial myology and arthrology. In: D. M. Cohen, ed. Papers on the Systematics of Gadiform Fishes. Los Angeles: Natural History Museum of Los Angeles County. pp. 113–128. Google Scholar

    768.

    Howes, G. J. 1990. The syncranial osteology of the southern eelcod family Muraenolepididae, with comments on its phylogenetic-relationships and on the biogeography of sub-antarctic gadoid fishes. Zoological Journal of the Linnean Society 100(1):73–100. Google Scholar

    769.

    Howes, G. J. 1991a. Anatomy, phylogeny and taxonomy of the gadoid fish genus Macruronus Günther, 1873, with a revised hypothesis of gadoid phylogeny. Bulletin of the British Museum (Natural History), Zoology 57(1):77–110. Google Scholar

    770.

    Howes, G. J. 1991b. Systematics and biogeography: an overview. In: I. J. Winfield and J. S. Nelson, eds. Cyprinid Fishes: Systematics, Biology, and Exploitation. London: Chapman and Hall. pp. 1–33. Google Scholar

    771.

    Howes, G. J. 1992. Notes on the anatomy and classification of ophidiiform fishes with particular reference to the abyssal genus Acanthonus Günther, 1878. Bulletin of the British Museum (Natural History), Zoology 58(2):95–131. Google Scholar

    772.

    Howes, G. J. 1993. Anatomy of the Melanonidae (Teleostei: Gadiformes), with comments on its phylogenetic relationships. Bulletin of the British Museum (Natural History), Zoology 59(1):11–31. Google Scholar

    773.

    Howes, G. J., and C. P. J. Sanford. 1987. The phylogenetic position of the Plecoglossidae (Teleostei, Salmoniformes), with comments on the Osmeridae and Osmeroidei. In: S. O. Kullander and B. Fernholm, eds. Proceedings V Congress of European Ichthyologists. Stockholm: Department of Vertebrate Zoology, Swedish Museum of Natural History. pp. 17–30. Google Scholar

    774.

    Huang, Y., K. Zhu, Y. Yang, L. Fang, Z. Liu, J. Ye, C. Jia, J. Chen, and H. Jiang. 2023. Comparative analysis of complete mitochondrial genome of Ariosoma meeki (Jordan and Snider, 1900), revealing gene rearrangement and the phylogenetic relationships of Anguilliformes. Biology 12(3):348. Google Scholar

    775.

    Hubbs, C. L. 1924. Studies of the Fishes of the Order Cyprinodontes. Ann Arbor: University of Michigan Museum of Zoology. 31 pp. (Miscellaneous Publications 13.) Google Scholar

    776.

    Hubbs, C. L. 1927. Notes of the blennioid fishes of western North America. Papers of the Michigan Academy of Science, Arts, and Letters 7:351–394. Google Scholar

    777.

    Hubbs, C. L. 1936. Fishes of the Yucatan Peninsula. Carnegie Institution of Washington Publications 457:157–287. Google Scholar

    778.

    Hubbs, C. L. 1945. Phylogenetic Position of the Citharidae, a Family of Flatfishes. Ann Arbor: University of Michigan Museum of Zoology. 38 pp. (Miscellaneous Publications 63.) Google Scholar

    779.

    Hubbs, C. L. 1952. A contribution to the classification of the blennioid fishes of the family Clinidae, with a partial revision of eastern Pacific forms. Stanford Ichthyological Bulletin 4(2):41–165. Google Scholar

    780.

    Hubert, N., C. Bonillo, and D. Paugy. 2005a. Does elision account for molecular saturation: case study based on mitochondrial ribosomal DNA among Characiform fishes (Teleostei: Ostariophysii). Molecular Phylogenetics and Evolution 35(1):300–308. Google Scholar

    781.

    Hubert, N.. 2005b. Early divergence among the Alestidae (Teleostei, Ostariophyses, Characiformes): mitochondrial evidences and congruence with morphological data. Comptes Rendus Biologies 328(5):477–491. Google Scholar

    782.

    Huddleston, R. W., and K. M. Savoie. 1983. Teleostean otoliths from the Late Cretaceous (Maestrichtian age) Severn Formation of Maryland. Proceedings of the Biological Society of Washington 96(4):658–663. Google Scholar

    783.

    Hughes, L. C., C. M. Nash, W. T. White, and M. W. Westneat. 2023. Concordance and discordance in the phylogenomics of the wrasses and parrotfishes (Teleostei: Labridae). Systematic Biology 72(3):530–543. Google Scholar

    784.

    Hughes, L. C., G. Ortí, Y. Huang, Y. Sun, C. C. Baldwin, A. W. Thompson, D. Arcila, R. Betancur-R, C. Li, L. Becker, ET AL. 2018. Comprehensive phylogeny of ray-finned fishes (Actinopterygii) based on transcriptomic and genomic data. Proceedings of the National Academy of Sciences of the United States of America 115(24):6249–6254. Google Scholar

    785.

    Huie, J. M., C. E. Thacker, and L. Tornabene. 2020. Co-evolution of cleaning and feeding morphology in western Atlantic and eastern Pacific gobies. Evolution 74(2):419–433. Google Scholar

    786.

    Hulet, W. H., and C. R. Robins. 1989. The evolutionary significance of the lepotocephalus larva. In: E. B. Böhlke, ed. Orders Anguilliformes and Saccopharyngiformes, Volume 2. New Haven, CT: Sears Foundation for Marine Research, Yale University. pp. 669–677. (Memoir 1, Fishes of the Western North Atlantic 9.) Google Scholar

    787.

    Humboldt, F. H. A. V., and A. Valenciennes. 1821. Recherches sur les poissons fluviatiles de l'Amérique Équinoxiale. In: Voyage de Humboldt et Bonpland, deuxième partie. Observations de Zoologie et d'Anatomie comparée, Volume 2. Paris: J. Smith. pp. 145–216. Google Scholar

    788.

    Hundt, P. J., S. P. Iglésias, A. S. Hoey, and A. M. Simons. 2014. A multilocus molecular phylogeny of combtooth blennies (Percomorpha: Blennioidei: Blenniidae): multiple invasions of intertidal habitats. Molecular Phylogenetics and Evolution 70:47–56. Google Scholar

    789.

    Hundt, P. J., and A. M. Simons. 2018. Extreme dentition does not prevent diet and tooth diversification within combtooth blennies (Ovalentaria: Blenniidae). Evolution 72(4):930–943. Google Scholar

    790.

    Hureau, J.-C. 1986. Relations phylogénétiques au sein des Notothenioidei. Oceanis 12(5):367–376. Google Scholar

    791.

    Hurley, I. A., R. L. Mueller, K. A. Dunn, E. J. Schmidt, M. Friedman, R. K. Ho, V. E. Prince, Z. H. Yang, M. G. Thomas, and M. I. Coates. 2007. A new time-scale for ray-finned fish evolution. Proceedings of the Royal Society B, Biological Sciences 274(1609):489–498. Google Scholar

    792.

    Ida, H. 1976. Removal of the family Hypoptychidae from the suborder Ammodytoidei, order Perciformes, to the suborder Gasterosteoidei, order Syngnathiformes. Japanese Journal of Ichthyology 23(1):33–42. Google Scholar

    793.

    Ilves, K. L., and E. B. Taylor. 2009. Molecular resolution of the systematics of a problematic group of fishes (Teleostei: Osmeridae) and evidence for morphological homoplasy. Molecular Phylogenetics and Evolution 50(1):163–178. Google Scholar

    794.

    Imamura, H. 1996. Phylogeny of the family Platycephalidae and related taxa (Pisces: Scorpaeniformes). Species Diversity 1(2):123–233. Google Scholar

    795.

    Imamura, H. 2000. An alternative hypothesis on the phylogenetic position of the family Dactylopteridae (Pisces: Teleostei), with a proposed new classification. Ichthyological Research 47:203–222. Google Scholar

    796.

    Imamura, H. 2004. Phylogenetic relationships and new classification of the superfamily Scorpaenoidea (Actinopterygii: Perciformes). Species Diversity 9(1):1–36. Google Scholar

    797.

    Imamura, H., and K. Matsuura. 2003. Redefinition and phylogenetic relationships of the family Pinguipedidae (Teleostei: Perciformes). Ichthyological Research 50:259–269. Google Scholar

    798.

    Imamura, H., and K. Odani. 2013. An overview of the phylogenetic relationships of the suborder Trachinoidei (Acanthomorpha: Perciformes). Ichthyological Research 60:1–15. Google Scholar

    799.

    Imamura, H., and G. Shinohara. 1998. Scorpaeniform fish phylogeny: an overview. Bulletin of the National Science Museum, Tokyo, Series A, 24:185–212. Google Scholar

    800.

    Imamura, H., S. M. Shirai, and M. Yabe. 2005. Phylogenetic position of the family Trichodontidae (Teleostei: Perciformes), with a revised classification of the perciform suborder Cottoidei. Ichthyological Research 52:264–274. Google Scholar

    801.

    Imamura, H., and M. Yabe. 2002. Demise of the Scorpaeniformes (Actinopterygii: Percomorpha): an alternative phylogenetic hypothesis. Bulletin of Fisheries Sciences, Hokkaido University 53(3):107–128. Google Scholar

    802.

    Inada, T. 1989. Current status of the systematics of Merlucciidae. In: D. M. Cohen, ed. Papers on the Systematics of Gadiform Fishes. Los Angeles: Natural History Museum of Los Angeles County. pp. 197–207. Google Scholar

    803.

    Inoue, J. G., Y. Kumazawa, M. Miya, and M. Nishida. 2009. The historical biogeography of the freshwater knifefishes using mitogenomic approaches: a Mesozoic origin of the Asian notopterids (Actinopterygii: Osteoglossomorpha). Molecular Phylogenetics and Evolution 51(3):486–499. Google Scholar

    804.

    Inoue, J. G., M. Miya, M. J. Miller, T. Sado, R. Hanel, K. Hatooka, J. Aoyama, Y. Minegishi, M. Nishida, and K. Tsukamoto. 2010. Deep-ocean origin of the freshwater eels. Biology Letters 6(3):363–366. Google Scholar

    805.

    Inoue, J. G., M. Miya, K. Tsukamoto, and M. Nishida. 2001. A mitogenomic perspective on the basal teleostean phylogeny: resolving higher-level relationships with longer DNA sequences. Molecular Phylogenetics and Evolution 20(2):275–285. Google Scholar

    806.

    Inoue, J. G.. 2003a. Basal actinopterygian relationships: a mitogenomic perspective on the phylogeny of the “ancient fish.” Molecular Phylogenetics and Evolution 26(1):110–120. Google Scholar

    807.

    Inoue, J. G.. 2003b. Evolution of the deep-sea gulper eel mitochondrial genomes: large-scale gene rearrangements originated within the eels. Molecular Biology and Evolution 20(11):1917–1924. Google Scholar

    808.

    Inoue, J. G.. 2004. Mitogenomic evidence for the monophyly of elopomorph fishes (Teleostei) and the evolutionary origin of the leptocephalus larva. Molecular Phylogenetics and Evolution 32(1):274–286. Google Scholar

    809.

    [IPA] INTERNATIONAL PHONETIC ASSOCIATION. 1999. Handbook of the International Phonetic Association: A Guide to the Use of the International Phonetic Alphabet. Cambridge: Cambridge University Press. 204 pp. Google Scholar

    810.

    Ishida, M. 1994. Phylogeny of the suborder Scorpaenoidei (Pisces: Scorpaeniformes). Bulletin of Nansei National Fisheries Research Institute 27:1–112. Google Scholar

    811.

    Ishiguro, N. B., M. Miya, and M. Nishida. 2003. Basal euteleostean relationships: a mitogenomic perspective on the phylogenetic reality of the “Protacanthopterygii.” Molecular Phylogenetics and Evolution 27(3):476–488. Google Scholar

    812.

    Ivantsoff, W., B. Said, and A. Williams. 1987. Systematic position of the family Dentatherinidae in relationship to Phallostethidae and Atherinidae. Copeia 1987(3):649–658. Google Scholar

    813.

    Iwami, T. 1985. Osteology and Relationships of the Family Channichthyidae. Tokyo: National Institute of Polar Research. 69 pp. (Memoirs of the National Institute of Polar Research 36.) Google Scholar

    814.

    Iwamoto, T. 1989. Phylogeny of grenadiers (suborder Macrouroidei). In: D. M. Cohen, ed. Papers on the Systematics of Gadiform Fishes. Los Angeles: Natural History Museum of Los Angeles County. pp. 159–173. (Science Series 32.) Google Scholar

    815.

    Jackson, K. L. 2003. Contributions to the systematics of cottoid fishes (Teleostei: Scorpaenifromes) [dissertation]. University of Alberta. 181 pp.  https://doi.org/10.7939/r3-4wnj-sn29 Google Scholar

    816.

    Jamieson, B. G. M. 1991. Fish Evolution and Systematics: Evidence from Spermatozoa. New York: Cambridge University Press. 319 pp. Google Scholar

    817.

    Jarvik, E. 1980. Basic Structure and Evolution of Vertebrates. London: Academic Press. 2 volumes. Google Scholar

    818.

    Ji, H.-S., J.-K. Kim, and B.-J. Kim. 2016. Molecular phylogeny of the families Pleuronectidae and Poecilopsettidae (Pisces, Pleuronectiformes) from Korea, with a proposal for a new classification. Ocean Science Journal 51:299–304. Google Scholar

    819.

    Jiangyong, Z. 1990. The new materials of Kuntulunia and its systematic position. Vertebrata PalAsiatica 28(2):128–139. Google Scholar

    820.

    Johansen, K. 1966. Air breathing in the teleost Symbranchus marmoratus. Comparative Biochemistry and Physiology 18(2):383–395. Google Scholar

    821.

    Johnson, G. D. 1980. The limits and relationships of the Lutjanidae and associated families. Berkeley: University of California Press. 114 pp. (Bulletin of the Scrippps Institution of Oceanography 24.) Google Scholar

    822.

    Johnson, G. D. 1983. Niphon spinosus: a primitive epinepheline serranid, with comments on the monophyly and intrarelationships of the Serranidae. Copeia 1983(3):777–787. Google Scholar

    823.

    Johnson, G. D. 1984. Percoidei: development and relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 464–498. (Special Publication 1.) Google Scholar

    824.

    Johnson, G. D. 1986. Scombroid phylogeny: an alternative hypothesis. Bulletin of Marine Science 39(1):1–41. Google Scholar

    825.

    Johnson, G. D. 1992. Monophyly of the euteleostean clades: Neoteleostei, Eurypterygii, and Ctenosquamata. Copeia 1992(1):8–25. Google Scholar

    826.

    Johnson, G. D. 1993. Percomorph phylogeny: progress and problems. Bulletin of Marine Science 52(1):3–28. Google Scholar

    827.

    Johnson, G. D. 2019. Revisions of anatomical descriptions of the pharyngeal jaw apparatus in moray eels of the family Muraenidae (Teleostei: Anguilliformes). Copeia 107(2):341–357. Google Scholar

    828.

    Johnson, G. D., and E. B. Brothers. 1993. Schindleria: a paedomorphic goby (Teleostei: Gobioidei). Bulletin of Marine Science 52(1):441–471. Google Scholar

    829.

    Johnson, G. D., H. Ida, J. Sakaue, T. Sado, T. Asahida, and M. Miya. 2012. A ‘living fossil’ eel (Anguilliformes: Protanguillidae, fam. nov.) from an undersea cave in Palau. Proceedings of the Royal Society B, Biological Sciences 279(1730):934–943. Google Scholar

    830.

    Johnson, G. D., and C. Patterson. 1993. Percomorph phylogeny: a survey of acanthomorphs and a new proposal. Bulletin of Marine Science 52(1):554–626. Google Scholar

    831.

    Johnson, G. D.. 1996. Relationships of lower euteleostean fishes. In: M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson, eds. Interrelationships of Fishes. San Diego: Academic Press. pp. 251–332. Google Scholar

    832.

    Johnson, G. D.. 1997. The gill-arches of gonorynchiform fishes. South African Journal of Science 93(11–12):594–600. Google Scholar

    833.

    Johnson, G. D., J. R. Paxton, T. T. Sutton, T. P. Satoh, T. Sado, M. Nishida, and M. Miya. 2009. Deep-sea mystery solved: astonishing larval transformations and extreme sexual dimorphism unite three fish families. Biology Letters 5(2):235–239. Google Scholar

    834.

    Johnson, G. D., and B. B. Washington. 1987. Larvae of the Moorish idol, Zanclus cornutus, including a comparison with other larval acanthuroids. Bulletin of Marine Science 40(3):494–511. Google Scholar

    835.

    Johnson, J. Y. 1862. Descriptions of some new genera and species of fishes obtained at Madeira. Proceedings of the Zoological Society of London 1862, pt. 2. (art. 2), pp. 167–180. Google Scholar

    836.

    Johnson, J. Y. 1863. Descriptions of five new species of fishes obtained at Madeira. Proceedings of the Zoological Society of London 1863, pt. 1 (art. 5), pp. 36–46, pl. 7. Google Scholar

    837.

    Johnson, J. Y. 1864. Descriptions of three new genera of marine fishes obtained at Madeira. Proceedings of the Zoological Society of London 1863, pt. 3 (art. 8), pp. 403–410. Google Scholar

    838.

    Johnson, R. K. 1982. Fishes of the Families Evermannellidae and Scopelarchidae: Systmatics, Morphology, Interrelationships, and Zoogeography. Chicago: Field Museum of Natural History. 252 pp. (Fieldiana Zoology New Series 12.) Google Scholar

    839.

    Jondeung, A., P. Sangthong, and R. Zardoya. 2007. The complete mitochondrial DNA sequence of the Mekong giant catfish (Pangasianodon gigas), and the phylogenetic relationships among Siluriformes. Gene 387:49–57. Google Scholar

    840.

    Jones, J. M. 1874. A new fish. Zoologist, Series 2, 9:3837–3838. Google Scholar

    841.

    Jones, W. J., and J. M. Quattro. 1999. Phylogenetic affinities of pygmy sunfishes (Elassoma) inferred from mitochondrial DNA sequences. Copeia 1999(2):470–474. Google Scholar

    842.

    Jordan, D. S. 1896. Notes on fishes, little known or new to science. Proceedings of the California Academy of Sciences, Series 2, 6:201–244. Google Scholar

    843.

    Jordan, D. S. 1905. A Guide to the Study of Fishes, Volume 2. New York: Henry Holt and Company. 599 pp. Google Scholar

    844.

    Jordan, D. S. 1923. A classification of fishes including families and genera as far as known. Stanford University Publications University Series Biological Sciences 3(2):77–243. Google Scholar

    845.

    Jordan, D. S., and B. W. Evermann. 1896. The fishes of North and Middle America: a descriptive catalogue of the species of fish-like vertebrates found in the waters of North America, north of the Isthmus of Panama. Part I. Bulletin of the United States National Museum 47:1–1240. Google Scholar

    846.

    Jordan, D. S.. 1898. The fishes of North and Middle America: a descriptive catalogue of the species of fish-like vertebrates found in the waters of North America, north of the Isthmus of Panama. Part III. Bulletin of the United States National Museum 47:2184–3136. Google Scholar

    847.

    Jordan, D. S., and H. W. Fowler. 1902. A review of the berycoid fishes of Japan. Proceedings of the United States National Museum 26(1306):1–21. Google Scholar

    848.

    Jordan, D. S., and C. H. Gilbert. 1882a. Descriptions of thirty-three new species of fishes from Mazatlan, Mexico. Proceedings of the United States National Museum 4(237):338–365. Google Scholar

    849.

    Jordan, D. S.. 1882b. Notes on fishes observed about Pensacola, Florida, and Galveston, Texas, with description of new species. Proceedings of the United States National Museum 5(282):241–307. Google Scholar

    850.

    Jordan, D. S., and C. L. Hubbs. 1919. Studies in Ichthyology: A Monographic Review of the Family of Atherinidae or Silversides. Stanford, CA: Stanford University Press. 87 pp. (Stanford Junior University Publications, University Series 40,) Google Scholar

    851.

    Jordan, D. S., and M. Sindo. 1902. A review of the pediculate fishes or anglers of Japan. Proceedings of the United States National Museum 24(1261):361–381. Google Scholar

    852.

    Jordan, D. S., and J. O. Snyder. 1901. A review of the cardinal fishes of Japan. Proceedings of the United States National Museum 23(1240):891–913. Google Scholar

    853.

    Jordan, D. S.. 1902. A review of the trachinoid fishes and their supposed allies found in the waters of Japan. Proceedings of the United States National Museum 24(1263):461–497. Google Scholar

    854.

    Jordan, D. S., and E. C. Starks. 1895. The fishes of Puget Sound. Proceedings of the California Academy of Sciences, Series 2, 5:785–855. Google Scholar

    855.

    Jordan, D. S.. 1901. A review of the atherine fishes of Japan. Proceedings of the United States National Museum 24(1250):199–206. Google Scholar

    856.

    Kai, Y., and F. Tashiro. 2019. Zenopsis filamentosa (Zeidae), a new mirror dory from the western Pacific Ocean, with redescription of Zenopsis nebulosa. Ichthyological Research 66:340–352. Google Scholar

    857.

    Kalumba, L. N., E. Abwe, F. D. B. Schedel, A. Chocha Manda, U. K. Schliewen, and E. J. W. M. N. Vreven. 2023. Two new shellear species (Gonorhynchiformes: Kneriidae), from the Luansa River (Upper Congo Basin): hidden diversity revealed by integrative taxonomy. Diversity 15(10):1044. Google Scholar

    858.

    Kappas, I., S. Vittas, C. N. Pantzartzi, E. Drosopoulou, and Z. G. Scouras. 2016. A time-calibrated mitogenome phylogeny of catfish (Teleostei: Siluriformes). PLOS ONE 11(12):e0166988.  https://doi.org/10.1371/journal.pone.0166988 Google Scholar

    859.

    Kartavtsev, Y. P., T. J. Park, J. S. Lee, K. A. Vinnikov, V. N. Ivankov, S. N. Sharina, and A. S. Ponomarev. 2008. Phylogenetic inferences introduced on cytochrome b gene sequences data for six flatfish species (Teleostei, Pleuronectidae) and species synonymy between representatives of genera Pseudopleuronectes and Hippoglossoides from far eastern seas. Russian Journal of Genetics 44(4):451–458. Google Scholar

    860.

    Kaufman, L., and K. F. Liem. 1982. Fishes of the suborder Labroidei (Pisces: Perciformes): phylogeny, ecology, and evolutionary significance. Breviora 472:1–19. Google Scholar

    861.

    Kaup, J. J. 1856. Catalogue of Apodal Fish, in the Collection of the British Museum. London: British Museum. 163 pp. Google Scholar

    862.

    Kawahara, R., M. Miya, K. Mabuchi, S. Lavoué, J. G. Inoue, T. P. Satoh, A. Kawaguchi, and M. Nishida. 2008. Interrelationships of the 11 gasterosteiform families (sticklebacks, pipefishes, and their relatives): a new perspective based on whole mitogenome sequences from 75 higher teleosts. Molecular Phylogenetics and Evolution 46(1):224–236. Google Scholar

    863.

    Kawahara, R., M. Miya, K. Mabuchi, T. J. Near, and M. Nishida. 2009. Stickleback phylogenies resolved: evidence from mitochondrial genomes and 11 nuclear genes. Molecular Phylogenetics and Evolution 50(2):401–404. Google Scholar

    864.

    Keivany, Y., and J. S. Nelson. 2004. Phylogenetic relationships of sticklebacks (Gasterosteidae), with emphasis on ninespine sticklebacks (Pungitius spp.). Behaviour 141(11–12):1485–1497. Google Scholar

    865.

    Kenaley, C. P., S. C. Devaney, and T. T. Fjeran. 2014. The complex evolutionary history of seeing red: molecular phylogeny and the evolution of an adaptive visual system in deep-sea dragonfishes (Stomiiformes: Stomiidae). Evolution 68(4):996–1013. Google Scholar

    866.

    Kendall, W. C., and L. Radcliffe. 1912. The shore fishes. Memoirs of the Museum of Comparative Zoology at Harvard College 35(3):75–171. Google Scholar

    867.

    Keskin, E., and H. H. Atar. 2013. DNA barcoding commercially important fish species of Turkey. Molecular Ecology Resources 13(5):788–797. Google Scholar

    868.

    Khajuria, C. K., and G. V. R. Prasad. 1998. Taphonomy of a late Cretaceous mammal-bearing microvertebrate assemblage from the Deccan inter-trappean beds of Naskal, peninsular India. Palaeogeography, Palaeoclimatology, Palaeoecology 137(1):153–172. Google Scholar

    869.

    Khalloufi, B., D. Ouarhache, and H. Lelièvre. 2010. New paleontological and geological data about Jbel Tselfat (Late Cretaceous of Morocco). Historical Biology 22(1–3):57–70. Google Scholar

    870.

    Kiernan, A. M. 1990. Systematics and zoogeography of the ronquils family Bathymasteridae (Teleostei:Perciformes) [dissertation]. University of Washington. 190 pp. Google Scholar

    871.

    Kikugawa, K., K. Katoh, S. Kuraku, H. Sakurai, O. Ishida, N. Iwabe, and T. Miyata. 2004. Basal jawed vertebrate phylogeny inferred from multiple nuclear DNA-coded genes. BMC Biology 2:3.  https://doi.org/10.1186/1741-7007-2-3 Google Scholar

    872.

    Kimura, K., H. Imamura, and T. Kawai. 2018. Comparative morphology and phylogenetic systematics of the families Cheilodactylidae and Latridae (Perciformes: Cirrhitoidea), and proposal of a new classification. Zootaxa 4536(1):1–72.  https://doi.org/10.11646/zootaxa.4536.1.1 Google Scholar

    873.

    Klausewitz, W. 1968. Fische aus dem Roten Meer. IX. Pseudochromis fridmani n. sp. aus dem Golf von Aqaba (Pisces, Osteichthyes, Pseudochromidae). Senckenbergiana Biologica 49(6):443–450. Google Scholar

    874.

    Klausewitz, W., and I. Eibl-Eibesfeldt. 1959. Neue Röhrenaale von den Maldiven und Nikobaren (Pisces, Apodes, Heterocongridae). Senckenbergiana Biologica 40(3–4):135–153. Google Scholar

    875.

    Kner, R. 1868. Über neue Fische aus dem Museum der Herren Johann Cäsar Godeffroy and Sohn in Hamburg. (IV. Folge). Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften. Mathematisch-Naturwissenschaftliche Classe 58:26–31. Google Scholar

    876.

    Knudsen, S. W., J. H. Choat, and K. D. Clements. 2019. The herbivorous fish family Kyphosidae (Teleostei: Perciformes) represents a recent radiation from higher latitudes. Journal of Biogeography 46(9):2067–2080. Google Scholar

    877.

    Knudsen, S. W., and K. D. Clements. 2016. World-wide species distributions in the family Kyphosidae (Teleostei: Perciformes). Molecular Phylogenetics and Evolution 101:252–266. Google Scholar

    878.

    Kobyliansky, S. G., N. V. Gordeeva, and A. N. Kotlyar. 2020. New findings of the rare species Rondeletia bicolor (Stephanoberycoidei) over the Mid-Atlantic Ridge and some peculiarities of the Rondeletiidae family's phylologeny. Journal of Ichthyology 60:13–21. Google Scholar

    879.

    Koeda, K., and H.-C. Ho. 2019. A new tapertail species (family Radiicephalidae; Lampridiformes) from Taiwan, the first confirmed western Pacific Ocean record of the family. Ichthyological Research 66:207–214. Google Scholar

    880.

    Konishi, Y., and M. Okiyama. 1997. Morphological development of four trachichthyoid larvae (Pisces: Beryciformes), with comments on trachichthyoid relationships. Bulletin of Marine Science 60:66–88. Google Scholar

    881.

    Kontula, T., S. V. Kirilchik, and R. Väinölä. 2003. Endemic diversification of the monophyletic cottoid fish species flock in Lake Baikal explored with mtDNA sequencing. Molecular Phylogenetics and Evolution 27(1):143–155. Google Scholar

    882.

    Kottelat, M. 1990. Synopsis of the endangered buntingi (Osteichthyes: Adrianichthyidae and Oryziidae) of Lake Poso, Central Sulawesi, Indonesia, with a new reproductive guild and descriptions of three new species. Ichthyological Exploration of Freshwaters 1:49–67. Google Scholar

    883.

    Kottelat, M. 2012. Conspectus cobitidum: an inventory of the loaches of the world (Teleostei: Cypriniformes: Cobitoidei). Raffles Bulletin of Zoology Suppl. 26:1–199. Google Scholar

    884.

    Kottelat, M., R. Britz, T. H. Hui, and K.-E. Witte. 2006. Paedocypris, a new genus of Southeast Asian cyprinid fish with a remarkable sexual dimorphism, comprises the world's smallest vertebrate. Proceedings of the Royal Society B, Biological Sciences 273(1589):895–899. Google Scholar

    885.

    Kriwet, J., and T. Hecht. 2008. A review of early gadiform evolution and diversification: first record of a rattail fish skull (Gadiformes, Macrouridae) from the Eocene of Antarctica, with otoliths preserved in situ. Naturwissenschaften 95:899–907. Google Scholar

    886.

    Kuang, T., L. Tornabene, J. Li, J. Jiang, P. Chakrabarty, J. S. Sparks, G. J. P. Naylor, and C. Li. 2018. Phylogenomic analysis on the exceptionally diverse fish clade Gobioidei (Actinopterygii: Gobiiformes) and data-filtering based on molecular clocklikeness. Molecular Phylogenetics and Evolution 128:192–202. Google Scholar

    887.

    Kuehne, L. M., and J. D. Olden. 2014. Ecology and conservation of mudminnow species worldwide. Fisheries 39(8):341–351. Google Scholar

    888.

    Kullander, S., M. Norén, M. M. Rahman, and A. R. Mollah. 2019. Chameleonfishes in Bangladesh: hipshot taxonomy, sibling species, elusive species, and limits of species delimitation (Teleostei: Badidae). Zootaxa 4586(2):301–337.  https://doi.org/10.11646/zootaxa.4586.2.7 Google Scholar

    889.

    Kullander, S., and R. Britz. 2002. Revision of the family Badidae (Teleostei: Perciformes), with description of a new genus and ten new species. Ichthyological Exploration of Freshwaters 13(4):295–372. Google Scholar

    890.

    Kwun, H. J. 2013. Phylogenetic study on the genera Eulophias and Zoarchias (Zoarcoidei: Perciformes) [dissertation]. Pukyong National University. 90 pp. Google Scholar

    891.

    Kwun, H. J., and J.-K. Kim. 2013. Molecular phylogeny and new classification of the genera Eulophias and Zoarchias (Pisces, Zoarcoidei). Molecular Phylogenetics and Evolution 69(3):787–795. Google Scholar

    892.

    Lacépède, B. G. E. 1801. Histoire naturelle des poissons, Volume 3. Paris: Plassan. 558 pp. Google Scholar

    893.

    Lacépède, B. G. E. 1802. Histoire naturelle des poissons, Volume 4. Paris: Plassan. 728 pp. Google Scholar

    894.

    Lacépède, B. G. E. 1803. Histoire naturelle des poissons, Volume 5. Paris: Plassan. 803 pp. Google Scholar

    895.

    Lambers, P. H. 1995. The monophyly of the Caturidae (Pisces; Actinopterygii) and the phylogeny of the Halecomorphi. Géobios 28(Suppl. 2):201–203. Google Scholar

    896.

    Landi, M., M. Dimech, M. Arculeo, G. Biondo, R. Martins, M. Carneiro, G. R. Carvalho, S. L. Brutto, and F. O. Costa. 2014. DNA barcoding for species assignment: the case of Mediterranean marine fishes. PLOS ONE 9(9):e106135.  https://doi.org/10.1371/journal.pone.0106135 Google Scholar

    897.

    Last, P. R. 2001. Leptoscopidae. In: K. E. Carpenter, and V. H. Niem, eds. Bony Fishes Part 4 (Labridae to Latimeriidae), Estuarine Crocodiles, Sea Turtles, Sea Snakes and Marine Mammals. Vol. 6 of FAO Species Identification Guide for Fishery Purposes: The Living Marine Resources of the Western Central Pacific. Rome: Food and Agriculture Organization of the United Nations. pp. 3517. Google Scholar

    898.

    Latimer, A. E., and S. Giles. 2018. A giant dapediid from the Late Triassic of Switzerland and insights into neopterygian phylogeny. Royal Society Open Science 5(8):180497.  https://doi.org/10.1098/rsos.180497 Google Scholar

    899.

    Lauder, G. V. 1980. Evolution of the feeding mechanism in primitive actionopterygian fishes: a functional anatomical analysis of Polypterus, Lepisosteus, and Amia. Journal of Morphology 163(3):283–317. Google Scholar

    900.

    Lauder, G. V. 1983. Functional design and evolution of the pharyngeal jaw apparatus in euteleostean fishes. Zoological Journal of the Linnean Society 77(1):1–38. Google Scholar

    901.

    Lauder, G. V., and K. F. Liem. 1982. Symposium summary: evolutionary patterns in actinopterygian fishes. American Zoologist 22(2):343–345. Google Scholar

    902.

    Lauder, G. V.. 1983. The evolution and interrelationships of the actinopterygian fishes. Bulletin of the Museum of Comparative Zoology 150(3):95–197. Google Scholar

    903.

    Lautredou, A.-C., H. Motomura, C. Gallut, C. Ozouf-Costaz, C. Cruaud, G. Lecointre, and A. Dettaï. 2013. New nuclear markers and exploration of the relationships among Serraniformes (Acanthomorpha, Teleostei): the importance of working at multiple scales. Molecular Phylogenetics and Evolution 67(1):140–155. Google Scholar

    904.

    Lavoué, S. 2015. Testing a time hypothesis in the biogeography of the arowana genus Scleropages (Osteoglossidae). Journal of Biogeography 42(12):2427–2439. Google Scholar

    905.

    Lavoué, S. 2016. Was Gondwanan breakup the cause of the intercontinental distribution of Osteoglossiformes? A time-calibrated phylogenetic test combining molecular, morphological, and paleontological evidence. Molecular Phylogenetics and Evolution 99:34–43. Google Scholar

    906.

    Lavoué, S., M. E. Arnegard, D. L. Rabosky, P. B. Mcintyre, D. Arcila, R. P. Vari, and M. Nishida. 2017. Trophic evolution in African citharinoid fishes (Teleostei: Characiformes) and the origin of intraordinal pterygophagy. Molecular Phylogenetics and Evolution 113:23–32. Google Scholar

    907.

    Lavoué, S., P. Konstantinidis, and W.-J. Chen. 2014. Progress in clupeiform systematics. In: K. Ganias, ed. Biology and Ecology of Anchovies and Sardines. Boca Raton, FL: CRC. pp. 3–42. Google Scholar

    908.

    Lavoué, S., M. Miya, M. E. Arnegard, P. B. Mcintyre, V. Mamonekene, and M. Nishida. 2011. Remarkable morphological stasis in an extant vertebrate despite tens of millions of years of divergence. Proceedings of the Royal Society B, Biological Sciences 278(1708):1003–1008. Google Scholar

    909.

    Lavoué, S., M. Miya, M. E. Arnegard, J. P. Sullivan, C. D. Hopkins, and M. Nishida. 2012. Comparable ages for the independent origins of electrogenesis in African and South American weakly electric fishes. PLOS ONE 7(5):e36287.  https://doi.org/10.1371/journal.pone.0036287 Google Scholar

    910.

    Lavoué, S., M. Miya, J. G. Inoue, K. Saitoh, N. B. Ishiguro, and M. Nishida. 2005. Molecular systematics of the gonorynchiform fishes (Teleostei) based on whole mitogenome sequences: implications for higher-level relationships within the Otocephala. Molecular Phylogenetics and Evolution 37(1):165–177. Google Scholar

    911.

    Lavoué, S., M. Miya, A. Kawaguchi, T. Yoshino, and M. Nishida. 2008. The phylogenetic position of an undescribed paedomorphic clupeiform taxon: mitogenomic evidence. Ichthyological Research 55:328–334. Google Scholar

    912.

    Lavoué, S., M. Miya, T. Moritz, and M. Nishida. 2012. A molecular timescale for the evolution of the African freshwater fish family Kneriidae (Teleostei: Gonorynchiformes). Ichthyological Research 59:104–112. Google Scholar

    913.

    Lavoué, S., M. Miya, P. Musikasinthorn, W.-J. Chen, and M. Nishida. 2013. Mitogenomic evidence for an Indo-West Pacific origin of the Clupeoidei (Teleostei: Clupeiformes). PLOS ONE 8(2):e56485.  https://doi.org/10.1371/journal.pone.0056485 Google Scholar

    914.

    Lavoué, S., M. Miya, and M. Nishida. 2010. Mitochondrial phylogenomics of anchovies (family Engraulidae) and recurrent origins of pronounced miniaturization in the order Clupeiformes. Molecular Phylogenetics and Evolution 56(1):480–485. Google Scholar

    915.

    Lavoué, S., M. Miya, J. Y. Poulsen, P. R. Møller, and M. Nishida. 2008. Monophyly, phylogenetic position and inter-familial relationships of the Alepocephaliformes (Teleostei) based on whole mitogenome sequences. Molecular Phylogenetics and Evolution 47(3):1111–1121. Google Scholar

    916.

    Lavoué, S., M. Miya, K. Saitoh, N. B. Ishiguro, and M. Nishida. 2007. Phylogenetic relationships among anchovies, sardines, herrings and their relatives (Clupeiformes), inferred from whole mitogenome sequences. Molecular Phylogenetics and Evolution 43(3):1096–1105. Google Scholar

    917.

    Lavoué, S., K. Nakayama, D. R. Jerry, Y. Yamanoue, N. Yagishita, N. Suzuki, M. Nishida, and M. Miya. 2014. Mitogenomic phylogeny of the Percichthyidae and Centrarchiformes (Percomorphaceae): comparison with recent nuclear gene-based studies and simultaneous analysis. Gene 549(1):46–57. Google Scholar

    918.

    Lavoué, S., and J. P. Sullivan. 2004. Simultaneous analysis of five molecular markers provides a well-supported phylogenetic hypothesis for the living bony-tongue fishes (Osteoglossomorpha: Teleostei). Molecular Phylogenetics and Evolution 33(1):171–185. Google Scholar

    919.

    Lê, H. L., G. Lecointre, and R. Perasso. 1993. A 28S rRNA-based phylogeny of gnathostomes: first steps in the analysis of conflict and congruence with morphologically based cladograms. Molecular Phylogenetics and Evolution 2(1):31–51. Google Scholar

    920.

    Lecointre, G. 1995. Molecular and morphological evidence for a Clupeomorpha-Ostariophysi sister-group relationship (Teleostei). Géobios 28(Suppl. 2):205–210. Google Scholar

    921.

    Lecointre, G., C. Bonillo, C. Ozouf-Costaz, and J.-C. Hureau. 1997. Molecular evidence for the origins of Antarctic fishes: paraphyly of the Bovichtidae and no indication for the monophyly of the Notothenioidei (Teleostei). Polar Biology 18:193–208. Google Scholar

    922.

    Lecointre, G., and G. Nelson. 1996. Clupeomorpha, sister-group of Ostariophysi. In: M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson, eds. Interrelationships of Fishes. San Diego: Academic Press. pp. 193–207. Google Scholar

    923.

    Leis, J. M. 1984. Tetraodontiformes: relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 459–463. (Special Publication 1.) Google Scholar

    924.

    Leslie, R. W., and W. S. Grant. 1994. Meristic and morphometric variation among anglerfish of the genus Lophius (Lophiiformes). Journal of Zoology 232(4):565–584. Google Scholar

    925.

    Lesueur, C. A. 1817. A short description of five (supposed) new species of the genus Muraena, discovered by Mr. Le Sueur, in the year 1816. Journal of the Academy of Natural Sciences of Philadelphia 1(5):81–83. Google Scholar

    926.

    Lesueur, C. A. 1818. Descriptions of several new species of North American fishes. Journal of the Academy of Natural Sciences of Philadelphia 1(2):222–235, 359–368. Google Scholar

    927.

    Levin, B. A., and A. S. Golubtsov. 2018. New insights into the molecular phylogeny and taxonomy of mormyrids (Osteoglossiformes, Actinopterygii) in northern East Africa. Journal of Zoological Systematics and Evolutionary Research 56(1):61–76. Google Scholar

    928.

    Li, B., A. Dettaï, C. Cruaud, A. Couloux, M. Desouttermeniger, and G. Lecointre. 2009. RNF213, a new nuclear marker for acanthomorph phylogeny. Molecular Phylogenetics and Evolution 50(1):345–363. Google Scholar

    929.

    Li, C. H., R. Betancur-R, W. L. Smith, and G. Ortí. 2011. Monophyly and interrelationships of snook and barramundi (Centropomidae sensu Greenwood) and five new markers for fish phylogenetics. Molecular Phylogenetics and Evolution 60(3):463–471. Google Scholar

    930.

    Li, C. H., G. Q. Lu, and G. Ortí. 2008. Optimal data partitioning and a test case for ray-finned fishes (Actinopterygii) based on ten nuclear loci. Systematic Biology 57(4):519–539. Google Scholar

    931.

    Li, C. H., and G. Ortí. 2007. Molecular phylogeny of Clupeiformes (Actinopterygii) inferred from nuclear and mitochondrial DNA sequences. Molecular Phylogenetics and Evolution 44(1):386–398. Google Scholar

    932.

    Li, C. H., G. Ortí, and J. L. Zhao. 2010. The phylogenetic placement of sinipercid fishes (“Perciformes”) revealed by 11 nuclear loci. Molecular Phylogenetics and Evolution 56(3):1096–1104. Google Scholar

    933.

    Li, G.-Q. 1996. A new species of Late Cretaceous osteoglossid (Teleostei) from the Oldman Formation of Alberta, Canada, and its phylogenetic relationships. In: G. Arratia and G. Viohl, eds. Mesozoic Fishes: Systematics and Paleoecology. Munich: Verlag Dr. Friedrich Pfeil. pp. 285–298. Google Scholar

    934.

    Li, G.-Q., L. Grande, and M. V. H. Wilson. 1997. The species of †Phareodus (Teleostei: Osteoglossidae) from the Eocene of North America and their phylogenetic relationships. Journal of Vertebrate Paleontology 17:487–505. Google Scholar

    935.

    Li, G.-Q., and M. V. H. Wilson. 1996. Phylogeny of Osteoglossomorpha. In: M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson, eds. Interrelationships of Fishes. San Diego: Academic Press. pp. 163–174. Google Scholar

    936.

    Li, G.-Q.1999. Early divergence of Hiodontiformes sensu stricto in East Asia and phylogeny of some Late Mesozoic teleosts from China. In: G. Arratia and H.-P. Schultze, eds. Mesozoic Fishes 2: Systematics and Fossil Record. Munich: Verlag Dr. Friedrich Pfeil. pp. 369–384. Google Scholar

    937.

    Li, G.-Q., M. V. H. Wilson, and L. Grande. 1997. Review of Eohiodon (Teleostei: Osteoglossomorpha) from western North America, with a phylogenetic reassessment of Hiodontidae. Journal of Paleontology 71(6):1109–1124. Google Scholar

    938.

    Li, H., Y. He, J. Jiang, Z. Liu, and C. Li. 2018. Molecular systematics and phylogenetic analysis of the Asian endemic freshwater sleepers (Gobiiformes: Odontobutidae). Molecular Phylogenetics and Evolution 121:1–11. Google Scholar

    939.

    Li, J., R. Xia, R. M. Mcdowall, J. A. López, G. C. Lei, and C. Z. Fu. 2010. Phylogenetic position of the enigmatic Lepidogalaxias salamandroides with comment on the orders of lower euteleostean fishes. Molecular Phylogenetics and Evolution 57(2):932–936. Google Scholar

    940.

    Li, X., P. Musikasinthorn, and Y. Kumazawa. 2006. Molecular phylogenetic analyses of snakeheads (Perciformes: Channidae) using mitochondrial DNA sequences. Ichthyological Research 53:148–159. Google Scholar

    941.

    Li, X., and W. Zhou. 2018. The species of Clupisoma from Yunnan, China (Teleostei: Siluriformes: Ailiidae), with a comment on the validity of the family Ailiidae. Zootaxa 4476(1):77–86.  https://doi.org/10.11646/zootaxa.4476.1.7  Google Scholar

    942.

    Lichtenstein, M. H. C. 1819. Ueber einige neue Arten von Fischen aus der Gattung Silurus. Zoologisches Magazin 1(3):57–63. Google Scholar

    943.

    Liddell, H. G., R. Scott, H. S. Jones, R. Mckenzie, and E. A. Barber. 1968. A Greek-English Lexicon, with a Supplement. Oxford: Clarendon Press. 2042 pp; 153 pp. Google Scholar

    944.

    Liem, K. F. 1970. Comparative Functional Anatomy of the Nandidae (Pisces: Teleostei). Chicago: Field Museum of Natural History. 166 pp. (Fieldiana: Zoology 56.) Google Scholar

    945.

    Liem, K. F., and P. H. Greenwood. 1981. A functional-approach to the phylogeny of the pharyngognath teleosts. American Zoologist 21(1):83–101. Google Scholar

    946.

    Lima, F. C. T., L. R. Malabarba, P. A. Buckup, J. F. Pezzi Da Silva, R. P. Vari, A. Harold, R. Benine, O. T. Oyakawa, C. S. Pavanelli, N. A. Menezes, ET AL. 2003. Family Characidae (Characins, tetras). In: R. E. Reis, S. O. Kullander, and C. J. Ferraris, eds. Check List of the Freshwater Fishes of South and Central America. Porto Alegre, Brazil: EDIPUCRS. pp. 104–169. Google Scholar

    947.

    Lin, H. C., and P. A. Hastings. 2013. Phylogeny and biogeography of a shallow water fish clade (Teleostei: Blenniiformes). BMC Evolutionary Biology 13:210.  https://doi.org/10.1186/1471-2148-13-210 Google Scholar

    948.

    Lin, Q., S. Fan, Y. Zhang, M. Xu, H. Zhang, Y. Yang, A. P. Lee, J. M. Woltering, V. Ravi, H. M. Gunter, ET AL. 2016. The seahorse genome and the evolution of its specialized morphology. Nature 540:395–399. Google Scholar

    949.

    Linnaeus, C. 1758. Systema Naturae per Regna Tria Naturae, Secundum Classes, Ordines, Genera, Species, cum Characteribus, Differentiis, Synonymis, Locis. Tomus I. Editio Decima, Reformata. Holmiae [Stockholm]: Impensis Direct. Laurentii Salvii. 824 pp. Google Scholar

    950.

    Linnaeus, C. 1761. Fauna Svecica, Sistens Animalia Sveciae Regni: Mammalia, Aves, Amphibia, Pisces, Insecta, Vermes: Distributa per Classes and Ordines, Genera and Species, Cum Differentiis Specierum, Synonymis Auctorum, Nominibus Incolarim, Locis Naturalium, Descriptionibus Insectorum. Altera Editio. Stockholmiae [Stockholm]: Sumtu and literis direct. Laurentii Salvii . 578 pp. Google Scholar

    951.

    Linnaeus, C. 1766. Systema Naturae per Regna Tria Naturae, Secundum Classes, Ordines, Genera, Species, cum Characteribus, Differentiis, Synonymis, Locis. Tomus 1. Edition Duodecima, Reformata Holmiae [Stockholm]: Impensis Direct. Laurentii Salvii. 532 pp. Google Scholar

    952.

    Little, A. G., S. C. Lougheed, and C. D. Moyes. 2010. Evolutionary affinity of billfishes (Xiphiidae and Istiophoridae) and flatfishes (Plueronectiformes): independent and trans-subordinal origins of endothermy in teleost fishes. Molecular Phylogenetics and Evolution 56(3):897–904. Google Scholar

    953.

    Liu, J., M.-M. Chang, M. V. H. Wilson, and A. M. Murray. 2015. A new family of Cypriniformes (Teleostei, Ostariophysi) based on a redescription of †Jianghanichthys hubeiensis (Lei, 1977) from the Eocene Yangxi Formation of China. Journal of Vertebrate Paleontology 35(6):e1004073. Google Scholar

    954.

    Liu, L., C. Li, Q. Liu, Z. Chen, and X. Fan. 2023. Genome-wide survey reveals the microsatellite characteristics and phylogenomic relationships of Dictyosoma burgeri (Zoarcales, Perciformes). Thalassas 39:609–619.  https://doi.org/10.1007/s41208-023-00598-7 Google Scholar

    955.

    Liu, S.-Q., R. L. Mayden, J.-B. Zhang, D. Yu, Q.-Y. Tang, X. Deng, and H.-Z. Liu. 2012. Phylogenetic relationships of the Cobitoidea (Teleostei: Cypriniformes) inferred from mitochondrial and nuclear genes with analyses of gene evolution. Gene 508(1):60–72. Google Scholar

    956.

    Liu, Z., X. Mu, H. Li, L. Gui, W. Zeng, and J. Zhang. 2016. Complete mitochondrial genome of the striped scat Selenotoca multifasciata (Perciformes: Scatophagidae). Mitochondrial DNA Part A 27(4):2691–2692. Google Scholar

    957.

    LIVY. 1919. History of Rome, Volume 1, Books 1 and 2. Translatedby B. O. Foster. Cambridge, MA: Harvard University Press. 447 pp. (Loeb Classical Library 114.) Google Scholar

    958.

    Lloris, D., J. Matallanas, and P. Oliver. 2005. Hakes of the World (Merlucciidae): An Annotated and Illustrated Catalogue of Hake Species Known to Date. Rome: Food and Agriculture Organization of the United Nations. 57 pp. (FAO Species Catalogue for Fishery Purposes 2.) Google Scholar

    959.

    Lockington, W. N. 1880. Description of a new genus and some new species of California fishes (Icosteus aenigmaticus and Osmerus attenuatus). Proceedings of the United States National Museum 3(123):63–68. Google Scholar

    960.

    Lofgren, D. L., J. A. Lillegraven, W. A. Clemens, P. D. Gingerich, and T. E. Williamson. 2004. Paleocene biochronology: the Puercan through Clarkforkian land mammal ages. In: M. O. Woodburne, ed. Late Cretaceous and Cenozoic Mammals of North America. New York: Columbia University Press. pp. 43–105. Google Scholar

    961.

    Longo, S. J., B. C. Faircloth, A. Meyer, M. W. Westneat, M. E. Alfaro, and P. C. Wainwright. 2017. Phylogenomic analysis of a rapid radiation of misfit fishes (Syngnathiformes) using ultraconserved elements. Molecular Phylogenetics and Evolution 113:33–48. Google Scholar

    962.

    López, J. A., P. Bentzen, and T. W. Pietsch. 2000. Phylogenetic relationships of esocoid fishes (Teleostei) based on partial cytochrome b and 16S mitochondrial DNA sequences. Copeia 2000(2):420–431. Google Scholar

    963.

    López, J. A., W.-J. Chen, and G. Ortí. 2004. Esociform phylogeny. Copeia 2004(3):449–464. Google Scholar

    964.

    López, J. A., M. W. Westneat, and R. Hanel. 2007. The phylogenetic affinities of the mysterious anguilliform genera Coloconger and Thalassenchelys as supported by mtDNA sequences. Copeia 2007(4):959–966. Google Scholar

    965.

    López-Arbarello, A. 2012. Phylogenetic interrelationships of ginglymodian fishes (Actinopterygii: Neopterygii). PLOS ONE 7(7):e39370.  https://doi.org/10.1371/journal.pone.0039370 Google Scholar

    966.

    López-Arbarello, A., T. Bürgin, H. Furrer, and R. Stockar. 2016. New holostean fishes (Actinopterygii: Neopterygii) from the Middle Triassic of the Monte San Giorgio (Canton Ticino, Switzerland). PeerJ 4:e2234.  https://doi.org/10.7717/peerj.2234 Google Scholar

    967.

    López-Arbarello, A.. 2019. Taxonomy and phylogeny of Eosemionotus Stolley, 1920 (Neopterygii: Ginglymodi) from the Middle Triassic of Europe. Palaeontologia Electronica 22.1.10:1–64.  https://doi.org/10.26879/904  Google Scholar

    968.

    López-Arbarello, A., E. E. Maxwell, and G. Schweigert. 2020. New halecomorph (Actinopterygii, Neopterygii) from the Nusplingen Lithographic Limestone (Upper Jurassic, Late Kimmeridgian), Germany. Journal of Vertebrate Paleontology 40(2):e1771348. Google Scholar

    969.

    López-Arbarello, A., and E. Sferco. 2018. Neopterygian phylogeny: the merger assay. Royal Society Open Science 5(3):172337.  https://doi.org/10.1098/rsos.172337 Google Scholar

    970.

    López-Arbarello, A., R. Stockar, and T. Bürgin. 2014. Phylogenetic relationships of the Triassic Archaeosemionotus Deecke (Halecomorphi, Ionoscopiformes) from the ‘Perledo Fauna.’ PLOS ONE 9(10):e108665.  https://doi.org/10.1371/journal.pone.0108665 Google Scholar

    971.

    Lovejoy, N. R. 2000. Reinterpreting recapitulation: systematics of needlefishes and their allies (Teleostei: Beloniformes). Evolution 54(4):1349–1362. Google Scholar

    972.

    Lovejoy, N. R., M. Iranpour, and B. B. Collette. 2004. Phylogeny and jaw ontogeny of beloniform fishes. Integrative and Comparative Biology 44(5):366–377. Google Scholar

    973.

    Løvtrup, S. 1977. The Phylogeny of Vertebrata. New York: John Wiley and Sons. 330 pp. Google Scholar

    974.

    Lowe, R. T. 1833. Description of Alepisaurus, a new genus of fishes. Proceedings of the Zoological Society of London 1833. Part 1, p. 104. Google Scholar

    975.

    Lowe, R. T. 1843. Notices of fishes newly observed or discovered in Madeira during the years 1840, 1841, and 1842. Proceedings of the Zoological Society of London 1843. Part 11, pp. 81–95. Google Scholar

    976.

    L-Recinos, M., K. M. Cantalice, C. Caballero-Viñas, and J. Alvarado-Ortega. 2023. A new Mesozoic teleost of the subfamily Albulinae (Albuliformes: Albulidae) highlights the proto-Gulf of Mexico in the early diversification of extant bonefishes. Journal of Systematic Palaeontology 21(1):2223797. Google Scholar

    977.

    Lü, Z., L. Gong, Y. Ren, Y. Chen, Z. Wang, L. Liu, H. Li, X. Chen, Z. Li, H. Luo, ET AL. 2021. Large-scale sequencing of flatfish genomes provides insights into the polyphyletic origin of their specialized body plan. Nature Genetics 53:742–751. Google Scholar

    978.

    Lü, Z., K. Zhu, H. Jiang, X. Lu, B. Liu, Y. Ye, L. Jiang, L. Liu, and L. Gong. 2019. Complete mitochondrial genome of Ophichthus brevicaudatus reveals novel gene order and phylogenetic relationships of Anguilliformes. International Journal of Biological Macromolecules 135:609–618. Google Scholar

    979.

    Lucena, C. A. S., and N. A. Menezes. 1998. A phylogenetic analysis of Roestes Günther and Gilbertolus Eigenmann, with a hypothesis on the relationships of the Cynodontidae and Acestorhynchidae (Teleostei: Ostariophysi: Characiformes). In: L. R. Malabarba, R. E. Reis, R. P. Vari, Z. M. Lucena, and C. A. S. Lucena, eds. Phylogeny and Classification of Neotropical Fishes. Porto Alegre, Brazil: EDIPUCRS. pp. 261–278. Google Scholar

    980.

    Lucentini, L., M. E. Puletti, C. Ricciolini, L. Gigliarelli, D. Fontaneto, L. Lanfaloni, F. Bilò, M. Natali, and F. Panara. 2011. Molecular and phenotypic evidence of a new species of genus Esox (Esocidae, Esociformes, Actinopterygii): the Southern Pike, Esox flaviae. PLOS ONE 6(12):e25218.  https://doi.org/10.1371/journal.pone.0025218 Google Scholar

    981.

    Ludt, W. B., C. P. Burridge, and P. Chakrabarty. 2019. A taxonomic revision of Cheilodactylidae and Latridae (Centrarchiformes: Cirrhitoidei) using morphological and genomic characters. Zootaxa 4585(1):121–141.  https://doi.org/10.11646/zootaxa.4585.1.7 Google Scholar

    982.

    Lund, R. 2000. The new actinopterygian order Guildayichthyiformes from the Lower Carboniferous of Montana (USA). Geodiversitas 22(6):171–206. Google Scholar

    983.

    Lundberg, J. G. 1993. African-South American freshwater fish clades and continental drift: problems with a paradigm. In: P. Goldblatt, ed. Biological Relationships between Africa and South America. New Haven, CT: Yale University Press. pp. 156–199. Google Scholar

    984.

    Lundberg, J. G. 2020a. Ostariophysi. In: K. De Queiroz, P. D. Cantino, and J. A. Gauthier, eds. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC. pp. 725–726. Google Scholar

    985.

    Lundberg, J. G. 2020b. Otophysi. In: K. De Queiroz, P. D. Cantino, and J. A. Gauthier, eds. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC. pp. 729–730. Google Scholar

    986.

    Lundberg, J. G. 2020c. Pan-Siluriformes. In: K. De Queiroz, P. D. Cantino, and J. A. Gauthier, eds. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC. pp. 733–734. Google Scholar

    987.

    Lundberg, J. G. 2020d. Siluriformes. In: K. De Queiroz, P. D. Cantino, and J. A. Gauthier, eds. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC. pp. 735–739. Google Scholar

    988.

    Lundberg, J. G., K. R. Luckenbill, K. K. Subhash Babu, and H. H. Ng. 2014. A tomographic osteology of the taxonomically puzzling catfish Kryptoglanis shajii (Siluriformes, Siluroidei, incertae sedis): description and a first phylogenetic interpretation. Proceedings of the Academy of Natural Sciences of Philadelphia 163:1–41. Google Scholar

    989.

    Lundberg, J. G., L. G. Marshall, J. Guerrero, B. Hor-Ton, M. C. Malabarba, and F. Wesselingh. 1998. The stage for Neotropical fish diversification: a history of tropical South American rivers. In: L. R. Malabarba, R. E. Reis, R. P. Vari, Z. M. Lucena, and C. A. S. Lucena, eds. Phylogeny and Classification of Neotropical Fishes. Porto Alegre, Brazil: EDIPUCRS. pp. 13–48. Google Scholar

    990.

    Lundberg, J. G., J. P. Sullivan, R. Rodiles-Hernández, and D. A. Hendrickson. 2007. Discovery of African roots for the Mesoamerican Chiapas catfish, Lacantunia enigmatica, requires an ancient intercontinental passage. Proceedings of the Academy of Natural Sciences of Philadelphia 156:39–53. Google Scholar

    991.

    Lundsten, L., S. B. Johnson, G. M. Cailliet, A. P. Devogelaere, and D. A. Clague. 2012. Morphological, molecular, and in situ behavioral observations of the rare deep-sea anglerfish Chaunacops coloratus (Garman, 1899), order Lophiiformes, in the eastern North Pacific. Deep Sea Research Part 1: Oceanographic Research Papers 68:46–53. Google Scholar

    992.

    Luo, T., Q. Yang, L. Wu, Y.-L. Wang, J.-J. Zhou, H.-Q. Deng, N. Xiao, and J. Zhou. 2023. Phylogenetic relationships of Nemacheilidae cavefish (Heminoemacheilus, Oreonectes, Yunnanilus, Paranemachilus, and Troglonectes) revealed by analysis of mitochondrial genome and seven nuclear genes. Zoological Research 44(4):693–697. Google Scholar

    993.

    Ma, K. Y., M. T. Craig, J. H. Choat, and L. Van Herwerden. 2016. The historical biogeography of groupers: clade diversification patterns and processes. Molecular Phylogenetics and Evolution 100:21–30. Google Scholar

    994.

    Mabee, P. M., E. A. Grey, G. Arratia, N. Bogutskaya, A. Boron, M. M. Coburn, K. W. Conway, S. P. He, A. Naseka, N. Rios, ET AL. 2011. Gill arch and hyoid arch diversity and cypriniform phylogeny: distributed integration of morphology and web-based tools. Zootaxa 2877(1):1–40.  https://doi.org/10.11646/zootaxa.2877.1.1 Google Scholar

    995.

    Mabuchi, K., T. H. Fraser, H. Song, Y. Azuma, and M. Nishida. 2014. Revision of the systematics of the cardinalfishes (Percomorpha: Apogonidae) based on molecular analyses and comparative reevaluation of morphological characters. Zootaxa 3846(2):151–203.  https://doi.org/10.11646/zootaxa.3846.2.1 Google Scholar

    996.

    Mabuchi, K., M. Miya, Y. Azuma, and M. Nishida. 2007. Independent evolution of the specialized pharyngeal jaw apparatus in cichlid and labrid fishes. BMC Evolutionary Biology 7:10.  https://doi.org/10.1186/1471-2148-7-10 Google Scholar

    997.

    Maddison, W. P., and D. R. Maddison. 2021. Mesquite: a modular system for evolutionary analysis [online analysis tool]. Version 3.70.  http://www.mesquiteproject.org  Google Scholar

    998.

    Maduna, S. N., A. Vivian-Smith, Ó. D. B. Jónsdóttir, A. K. D. Imsland, C. F. C. Klütsch, T. Nyman, H. G. Eiken, and S. B. Hagen. 2022. Mitogenomics of the suborder Cottoidei (Teleostei: Perciformes): improved assemblies, mitogenome features, phylogeny, and ecological implications. Genomics 114(2):110297. Google Scholar

    999.

    Mago-Leccia, F. 1978. Los peces de la familia Sternopygidae de Venezuela. Acta Científica Venezolana 29(Suppl. 1):1–51. Google Scholar

    1000.

    Mago-Leccia, F., and T. M. Zaret. 1978. The taxonomic status of Rhabdolichops troscheli (Kaup, 1856), and speculations on gymnotiform evolution. Environmental Biology of Fishes 3(4):379–384. Google Scholar

    1001.

    Maisey, J. G. 1986. Heads and tails: a chordate phylogeny. Cladistics 2(4):201–256. Google Scholar

    1002.

    Maisey, J. G. 1993. A new clupeomorph fish from the Santana Formation (Albian) of NE Brazil. American Museum Novitates 3076:1–15. Google Scholar

    1003.

    Makushok, V. M. 1958. The morphology and classification of the northern blennioid fishes (Stichaeidae, Blennioidei, Pisces). Trudy Zoologicheskogo Instituta, Akademiya Nauk SSSR 25:3–129. Google Scholar

    1004.

    Malabarba, M. C., and F. Di Dario. 2017. A new predatory herring-like fish (Teleostei: Clupeiformes) from the early Cretaceous of Brazil, and implications for relationships in the Clupeoidei. Zoological Journal of the Linnean Society 180(1):175–194. Google Scholar

    1005.

    Malabarba, M. C., and L. R. Malabarba. 2010. Biogeography of Characiformes: an evaluation of the available information of fossil and extant taxa. In: J. S. Nelson, H.-P. Schultze, and M. V. H. Wilson, eds. Origin and Phylogenetic Interrelationships of Teleosts. Munich: Verlag Dr. Friedrich Pfeil. pp. 317–336. Google Scholar

    1006.

    Malmstrøm, M., R. Britz, M. Matschiner, O. K. Tørresen, R. K. Hadiaty, H. Yaakob, H.-H. Tan, K. S. Jakobsen, W. Salzburger, and L. Rüber. 2018. The most developmentally truncated fishes show extensive Hox gene loss and miniaturized genomes. Genome Biology and Evolution 10(4):1088–1103. Google Scholar

    1007.

    Malmstrøm, M., M. Matschiner, O. K. Tørresen, K. S. Jakobsen, and S. Jentoft. 2017. Whole genome sequencing data and de novo draft assemblies for 66 teleost species. Scientific Data 4:160132. Google Scholar

    1008.

    Malmstrøm, M., M. Matschiner, O. K. Tørresen, B. Star, L. G. Snipen, T. F. Hansen, H. T. Baalsrud, A. J. Nederbragt, R. Hanel, W. Salzburger, ET AL. 2016. Evolution of the immune system influences speciation rates in teleost fishes. Nature Genetics 48:1204–1210. Google Scholar

    1009.

    Marić, S., D. Stanković, J. Wanzenböck, R. Šanda, T. Erős, P. Takács, A. Specziár, N. Sekulić, D. Bănăduc, M. Ćaleta, ET AL. 2017. Phylogeography and population genetics of the European mudminnow (Umbra krameri) with a time-calibrated phylogeny for the family Umbridae. Hydrobiologia 792:151–168. Google Scholar

    1010.

    Mar'ie, Z. A., and M. Allam. 2019. Molecular phylogenetic linkage for Nile and marine puffer fishes using mitochondrial DNA sequences of cytochrome b and 16S rRNA. Egyptian Journal of Aquatic Biology and Fisheries 23(5):67–80. Google Scholar

    1011.

    Markle, D. F. 1976. Preliminary studies on the systematics of deep-sea Alepocephaloidea (Pisces: Salmoniformes) [dissertation]. College of William and Mary. 240 pp. Google Scholar

    1012.

    Markle, D. F. 1989. Aspects of character homology and phylogeny of the Gadiformes. In: D. M. Cohen, ed. Papers on the Systematics of Gadiform Fishes. Los Angeles: Natural History Museum of Los Angeles County. pp. 59–88. Google Scholar

    1013.

    Marramà, G., and G. Carnevale. 2017. Morphology, relationships and palaeobiology of the Eocene barracudina †Holosteus esocinus (Aulopiformes: Paralepididae) from Monte Bolca, Italy. Zoological Journal of the Linnean Society 181(1):209–228. Google Scholar

    1014.

    Marramà, G.. 2018. Eoalosa janvieri gen. et sp. nov., a new clupeid fish (Teleostei, Clupeiformes) from the Eocene of Monte Bolca, Italy. Paläontologische Zeitschrift 92:107–120. Google Scholar

    1015.

    Marramà, G., B. Khalloufi, and G. Carnevale. 2023. Redescription of ‘Diplomystus’ solignaci Gaudant and Gaudant, 1971 from the Cretaceous of Tunisia, and a new hypothesis of double-armored herring relationships. Historical Biology 35(1):163–184. Google Scholar

    1016.

    Marshall, L. G., T. Sempere, and R. F. Butler. 1997. Chronostratigraphy of the mammal-bearing Paleocene of South America. Journal of South American Earth Sciences 10(1):49–70. Google Scholar

    1017.

    Marshall, N. B. 1962. Observations on the Heteromi: an order of teleost fishes. Bulletin of the British Museum (Natural History), Zoology 9(6):249–270. Google Scholar

    1018.

    Marshall, N. B. 1965. Systematic and biological studies of the Macrourid fishes (Anacanthini-Teleostii). Deep-Sea Research and Oceanographic Abstracts 12(3):299–322. Google Scholar

    1019.

    Marshall, N. B. 1966. The relationships of the anacanthine fishes, Macruronus, Lyconus, and Steindachneria. Copeia 1966(2):275–280. Google Scholar

    1020.

    Marshall, N. B., and D. M. Cohen. 1973. Order Anacanthini (Gadiformes): characters and synopsis of families. In: D. M. Cohen, J. W. Atz, R. H. Gibbs, F. H. Berry, E. A. Lachner, J. E. Böhlke, G. W. Mead, and D. Merriman, eds. Orders Heteromi (Notacanthiformes), Berycomorphi (Beryciformes), Xenoberyces (Stephanoberyciformes), and Anacanthini (Gadiformes): Halosuriforms, Killfishes, Squirrelfishes and other Beryciforms, Stephanoberyciforms, Grenadiers. New Haven, CT: Sears Foundation for Marine Research, Yale University. pp. 479–495. (Memoir 1, Fishes of the Western North Atlantic 6.) Google Scholar

    1021.

    Martens, E. 1862. Über einen neuen Polyodon (P. gladius) aus dem Yangtsekiang und über die sogenannten Glaspolypen. Monatsberichte der Deutschen Akademie der Wissenschaften zu Berlin 1861:476–480. Google Scholar

    1022.

    Martin, C. H., and P. C. Wainwright. 2011. Trophic novelty is linked to exceptional rates of morphological diversification in two adaptive radiations of Cyprinodon pupfish. Evolution 65(8):2197–2212. Google Scholar

    1023.

    Martin, J. M. 2015. Phylogeny, ontogeny and distribution of the ribbonfishes (Lampridiformes: Trachipteridae) [dissertation]. College of William and Mary. 217 pp. Google Scholar

    1024.

    Martin, R. P., E. E. Olson, M. G. Girard, W. L. Smith, and M. P. Davis. 2018. Light in the darkness: new perspective on lanternfish relationships and classification using genomic and morphological data. Molecular Phylogenetics and Evolution 121:71–85. Google Scholar

    1025.

    Matschiner, M., R. Hanel, and W. Salzburger. 2011. On the origin and trigger of the notothenioid adaptive radiation. PLOS ONE 6(4):e18911.  https://doi.org/10.1371/journal.pone.0018911 Google Scholar

    1026.

    Matsubara, K. 1955. Fish Morphology and Hierarchy. Tokyo: Ishizaki-Shoten. 3 volumes. [In Japanese.] Google Scholar

    1027.

    Matsunuma, M., and J. W. Johnson. 2023. Revision of the Indo-Pacific genus Centrogenys Richardson, 1842 (Centrogenyidae) with descriptions of two new species from Australia. Ichthyology and Herpetology 111(4):503–521. Google Scholar

    1028.

    Matsuura, K. 2015. Taxonomy and systematics of tetraodontiform fishes: a review focusing primarily on progress in the period from 1980 to 2014. Ichthyological Research 62:72–113. Google Scholar

    1029.

    Mattern, M. Y. 2004. Molecular phylogeny of the Gasterosteidae: the importance of using multiple genes. Molecular Phylogenetics and Evolution 30(2):366–377. Google Scholar

    1030.

    Mattern, M. Y. 2007. Phylogeny, systematics, and taxomony of sticklebacks. In: S. Östlund-Nilsson, I. Mayer, and F. A. Huntingford, eds. Biology of the Three-spined Stickleback. Boca Raton, FL: CRC Press. pp. 1–40. Google Scholar

    1031.

    Mattern, M. Y., and D. A. Mclennan. 2004. Total evidence phylogeny of Gasterosteidae: combining molecular, morphological and behavioral data. Cladistics 20(1):14–22. Google Scholar

    1032.

    Matthei, C., and X. Matthei. 1975. Spermatogenesis and spermatozoa of the Elopomorpha (teleost fish). In: B. A. Afzelius, ed. The Functional Anatomy of the Spermatozoon. Oxford: Pergamon. pp. 211–221. Google Scholar

    1033.

    Mattox, G. M. T., and M. Toledo-Piza. 2012. Phylogenetic study of the Characinae (Teleostei: Characiformes: Characidae). Zoological Journal of the Linnean Society 165(4):809–915. Google Scholar

    1034.

    Mayden, R. L., and W.-J. Chen. 2010. The world's smallest vertebrate species of the genus Paedocypris: a new family of freshwater fishes and the sister group to the world's most diverse clade of freshwater fishes (Teleostei: Cypriniformes). Molecular Phylogenetics and Evolution 57(1):152–175. Google Scholar

    1035.

    Mayden, R. L., W.-J. Chen, H. L. Bart, M. H. Doosey, A. M. Simons, K. L. Tang, R. M. Wood, M. K. Agnew, L. Yang, M. V. Hirt, ET AL. 2009. Reconstructing the phylogenetic relationships of the earth's most diverse clade of freshwater fishes—order Cypriniformes (Actinopterygii: Ostariophysi): a case study using multiple nuclear loci and the mitochondrial genome. Molecular Phylogenetics and Evolution 51(3):500–514. Google Scholar

    1036.

    Mayden, R. L., K. L. Tang, R. M. Wood, W.-J. Chen, M. K. Agnew, K. W. Conway, L. Yang, A. M. Simons, H. L. Bart, P. M. Harris, ET AL. 2008. Inferring the tree of life of the order Cypriniformes, the earth's most diverse clade of freshwater fishes: implications of varied taxon and character sampling. Journal of Systematics and Evolution 46(2):424–438. Google Scholar

    1037.

    Mayrinck, D. 2011. Phylogenetic relationships of otophysan (Actinopterygii, Teleostei), notably the Characiformes, including fossil members [dissertation]. French Ichtyological Society–Cybium. 298 pp. Google Scholar

    1038.

    Mayrinck, D., P. M. Brito, and O. Otero. 2010. A new albuliform (Teleostei: Elopomorpha) from the Lower Cretaceous Santana Formation, Araripe Basin, northeastern Brazil. Cretaceous Research 31(2):227–236. Google Scholar

    1039.

    Mayrinck, D.. 2015a. Anatomical review of †Salminops ibericus, a Teleostei incertae sedis from the Cenomanian of Portugal, anciently assigned to Characiformes and possibly related to crossognathiform fishes. Cretaceous Research 56:66–75. Google Scholar

    1040.

    Mayrinck, D.. 2015b. Review of the osteology of the fossil fish formerly attributed to the genus †Chanoides and implications for the definition of otophysan bony characters. Journal of Systematic Palaeontology 13(5):397–420. Google Scholar

    1041.

    Mcallister, D. E. 1968. The Evolution of Branchiostegals and Associated Opercular, Gular, and Hyoid Bones, and the Classification of Teleostome Fishes, Living and Fossil. Ottawa: National Museum of Canada. 239 pp. (Bulletin 221.) Google Scholar

    1042.

    Mcbride, R. S., C. R. Rocha, R. Ruiz-Carus, and B. W. Bowen. 2010. A new species of ladyfish, of the genus Elops (Elopiformes: Elopidae), from the western Atlantic Ocean. Zootaxa 2346(1):29–41.  https://doi.org/10.11646/zootaxa.2346.1.3 Google Scholar

    1043.

    Mccraney, W. T. 2019. Phylogeny and divergence times of gobiarian fishes [dissertation]. University of California, Los Angeles. 111 pp. Google Scholar

    1044.

    Mccraney, W. T., C. E. Thacker, and M. E. Alfaro. 2020. Supermatrix phylogeny resolves goby lineages and reveals unstable root of Gobiaria. Molecular Phylogenetics and Evolution 151:106862.  https://doi.org/10.1016/j.ympev.2020.106862 Google Scholar

    1045.

    Mcdougall, I. A. N., and C. S. Feibel. 1999. Numerical age control for the Miocene-Pliocene succession at Lothagam, a hominoid-bearing sequence in the northern Kenya Rift. Journal of the Geological Society 156:731–745. Google Scholar

    1046.

    Mcdowall, R. M. 1969. Relationships of galaxioid fishes with a further discussion of salmoniform classification. Copeia 1969(4):796–824. Google Scholar

    1047.

    Mcdowall, R. M. 1973. Relationships and taxonomy of the New Zealand torrent fish, Cheimarrichthys fosteri Haast (Pisces: Mugiloididae). Journal of the Royal Society of New Zealand 3(2):199–217. Google Scholar

    1048.

    Mcdowall, R. M. 1984. Southern hemisphere freshwater salmoniforms: development and relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 150–153. (Special Publication 1.) Google Scholar

    1049.

    Mcdowall, R. M. 2000. Biogeography of the New Zealand torrentfish, Cheimarrichthys fosteri (Teleostei: Pinguipedidae): a distribution driven mostly by ecology and behaviour. Environmental Biology of Fishes 58:119–131. Google Scholar

    1050.

    Mcdowall, R. M., and C. P. Burridge. 2011. Osteology and relationships of the southern freshwater lower euteleostean fishes. Zoosystematics and Evolution 87(1):7–185. Google Scholar

    1051.

    Mckay, R. J. 1992. Sillaginid Fishes of the World (Order Gadiformes). Vol. 14 of FAO Species Catalogue. Rome: Food and Agriculture Organization of the United Nations. 87 pp. (FAO Fisheries Synopsis 125.) Google Scholar

    1052.

    Mckay, R. J. 1997. Pearl Perches of the World (Family Glaucosomatidae). Vol. 17 of FAO Species Catalogue. Rome: Food and Agriculture Organization of the United Nations. 26 pp. (FAO Fisheries Synopsis 125.) Google Scholar

    1053.

    Mclennan, D. A. 1993. Phylogenetic relationships in the Gasterosteidae: an updated tree based on behavioral characters with a discussion of homoplasy. Copeia 1993(2):318–326. Google Scholar

    1054.

    Mclennan, D. A., and M. Y. Mattern. 2001. The phylogeny of the Gasterosteidae: combining behavioral and morphological data sets. Cladistics 17(1):11–27. Google Scholar

    1055.

    Mcmahan, C. D., D. J. Elías, Y. Li, O. Domínguez-Domínguez, S. Rodriguez-Machado, A. Morales-Cabrera, D. Velásquez-Ramírez, K. R. Piller, P. Chakrabarty, and W. A. Matamoros. 2021. Molecular systematics of the Awaous banana complex (river gobies; Teleostei: Oxudercidae). Journal of Fish Biology 99(3):970–979. Google Scholar

    1056.

    Mecklenburg, C. W. 2003a. Family Anarhichadidae Bonaparte 1846: wolffishes. California Academy of Sciences Annotated Checklists of Fishes 10:1–6. Google Scholar

    1057.

    Mecklenburg, C. W. 2003b. Family Bathymasteridae Jordan and Gilbert 1883: ronquils. California Academy of Sciences Annotated Checklists of Fishes 7:1–4. Google Scholar

    1058.

    Mecklenburg, C. W. 2003c. Family Cryptacanthodidae Gill 1861: wrymouths. California Academy of Sciences Annotated Checklists of Fishes 8:1–4. Google Scholar

    1059.

    Mecklenburg, C. W. 2003d. Family Ptilichthyidae Jordan and Gilbert 1883: quillfishes. California Academy of Sciences Annotated Checklists of Fishes 12:1–3. Google Scholar

    1060.

    Mecklenburg, C. W. 2003e. Family Scytalinidae Jordan and Evermann 1898: graveldivers. California Academy of Sciences Annotated Checklists of Fishes 11:1–3. Google Scholar

    1061.

    Mecklenburg, C. W. 2003f. Family Zaproridae Jordan and Evermann 1898: prowfishes. California Academy of Sciences Annotated Checklists of Fishes 13:1–3. Google Scholar

    1062.

    Mecklenburg, C. W., and B. Sheiko. 2004. Family Stichaeidae Gill 1864: pricklebacks. California Academy of Sciences Annotated Checklists of Fishes 35:1–36. Google Scholar

    1063.

    Mees, G. F. 1961. Description of a new fish of the family Galaxiidae from Western Australia. Journal of the Royal Society of Western Australia 44(pt. 2):33–38. Google Scholar

    1064.

    Meinema, E., and J. H. Huber. 2023. Review of Pantanodontidae (Cyprinodontiformes) and its fossil and extant genera Pantanodon Myers, 1955 and †Paralebias Gaudant, 2013 with descriptions of two new recent genera and five new species. In: J. H. Huber, ed. Killi-Data Series 2023 [pdf]. Killi-Data Online [online database]. pp. 21–73. Accessed 27 November 2023. https://www.killi-data.org  Google Scholar

    1065.

    Melo, B. F., M. C. C. De Pinna, L. H. Rapp Py-Daniel, J. Zuanon, C. C. Conde-Saldaña, F. F. Roxo, and C. Oliveira. 2022. Paleogene emergence and evolutionary history of the Amazonian fossorial fish genus Tarumania (Teleostei: Tarumaniidae). Frontiers in Ecology and Evolution 10:924860. Google Scholar

    1066.

    Melo, B. F., B. L. Sidlauskas, K. Hoekzema, R. P. Vari, C. B. Dillman, and C. Oliveira. 2018. Molecular phylogenetics of Neotropical detritivorous fishes of the family Curimatidae (Teleostei: Characiformes). Molecular Phylogenetics and Evolution 127:800–812. Google Scholar

    1067.

    Melo, B. F., B. L. Sidlauskas, T. J. Near, F. F. Roxo, A. Ghezelayagh, L. E. Ochoa, M. L. J. Stiassny, J. Arroyave, J. Chang, B. C. Faircloth, ET AL. 2022. Accelerated diversification explains the exceptional species richness of tropical characoid fishes. Systematic Biology 71(1):78–92. Google Scholar

    1068.

    Melo, M. R. S. 2009. Taxonomic and phylogenetic revision of the family Chiasmodontidae (Perciformes: Acanthomorpha) [dissertation]. Auburn University. pp. 540. Google Scholar

    1069.

    Meunier, F. J., and M. Gayet. 1996. A new polypteriform from the Late Cretaceous and the Middle Paleocene of South America. In: G. Arratia and G. Viohl, eds. Mesozoic Fishes: Systematics and Paleoecology. Munich: Verlag Dr. Friedrich Pfeil. pp. 95–105. Google Scholar

    1070.

    Meyer, H. W. 2003. The Fossils of Florissant. Washington, DC: Smithsonian Institution Press. 258 pp. Google Scholar

    1071.

    Mickle, K. E., R. Lund, and E. D. Grogan. 2009. Three new palaeoniscoid fishes from the Bear Gulch Limestone (Serpukhovian, Mississippian) of Montana (USA) and the relationships of lower actinopterygians. Geodiversitas 31(3):623–668. Google Scholar

    1072.

    Milec, L. J. M., M. P. M. Vanhove, F. M. Bukinga, E. L. R. De Keyzer, V. L. Kapepula, P. M. Masilya, N. S. Mulimbwa, C. E. Wagner, and J. A. M. Raeymaekers. 2022. Complete mitochondrial genomes and updated divergence time of the two freshwater clupeids endemic to Lake Tanganyika (Africa) suggest intralacustrine speciation. BMC Ecology and Evolution 22:127.  https://doi.org/10.1186/s12862-022-02085-8 Google Scholar

    1073.

    Miller, E. C., C. M. Martinez, S. T. Friedman, P. C. Wainwright, S. A. Price, and L. Tornabene. 2022. Alternating regimes of shallow and deep-sea diversification explain a species-richness paradox in marine fishes. Proceedings of the National Academy of Sciences 119:e2123544119. Google Scholar

    1074.

    Miller, P. J. 1973. The osteology and adaptive features of Rhyacichthys aspro (Teleostei: Gobioidei) and the classification of gobioid fishes. Journal of Zoology 171(3):397–434. Google Scholar

    1075.

    Miller, P. J. 1986. Reproductive biology and systematic problems in gobioid fishes. In: T. Uyeno, R. Arai, T. Taniuchi, and K. Matsuura, eds. Indo-Pacific Fish Biology: Proceedings of the Second International Conference on Indo-Pacific Fishes. Tokyo: Ichthyological Society of Japan. pp. 640–647. Google Scholar

    1076.

    Miller, P. J. 1992. The sperm duct gland: a visceral synapomorphy for gobioid fishes. Copeia 1992(1):253–256. Google Scholar

    1077.

    Miller, R. R. 1947. A new genus and species of deepsea fish of the family Myctophidae from the Philippine Islands. Proceedings of the United States National Museum 97(3211):81–90. Google Scholar

    1078.

    Milliken, D. M., and E. D. Houde. 1984. A new species of Bregmacerotidae (Pisces), Bregmaceros cantori, from the western Atlantic Ocean. Bulletin of Marine Science 35:11–19. Google Scholar

    1079.

    Miranda Ribeiro, A. DE. 1915. Fauna Brasiliense. Peixes. Tomo 5. Part 2. Eleutherobranchios aspirophoros (Physoclisti). Arquivos do Museu Nacional de Rio de Janeiro 17(1):679. Google Scholar

    1080.

    Mirande, J. M. 2009. Weighted parsimony phylogeny of the family Characidae (Teleostei: Characiformes). Cladistics 25(6):574–613. Google Scholar

    1081.

    Mirande, J. M. 2010. Phylogeny of the family Characidae (Teleostei: Characiformes): from characters to taxonomy. Neotropical Ichthyology 8:385–568. Google Scholar

    1082.

    Mirande, J. M. 2017. Combined phylogeny of ray-finned fishes (Actinopterygii) and the use of morphological characters in large-scale analyses. Cladistics 33(4):333–350. Google Scholar

    1083.

    Mitchill, S. L. 1814. Report, in Part, of Samuel L. Mitchill, M. D., Professor of Natural History, &c, on the Fishes of New-York. New York: D. Carlisle. 28 pp. Google Scholar

    1084.

    Mitchill, S. L. 1815. The fishes of New-York, described and arranged. Transactions of the Literary and Philosophical Society of New-York 1(5):.355–492. Google Scholar

    1085.

    Mitchill, S. L. 1818a. Description of three species of fish. Journal of the Academy of Natural Sciences of Philadelphia 1:407–412. Google Scholar

    1086.

    Mitchill, S. L. 1818b. Memoir on ichthyology. The fishes of New York, described and arranged. In a supplement to the Memoir... (continued). American Monthly Magazine and Critical Review 2(5):321–328. Google Scholar

    1087.

    Miya, M., M. Friedman, T. P. Satoh, H. Takeshima, T. Sado, W. Iwasaki, Y. Yamanoue, M. Nakatani, K. Mabuchi, J. G. Inoue, ET AL. 2013. Evolutionary origin of the Scombridae (tunas and mackerels): members of a Paleogene adaptive radiation with 14 other pelagic fish families. PLOS ONE 8(9):e73535.  https://doi.org/10.1371/journal.pone.0073535 Google Scholar

    1088.

    Miya, M., N. I. Holcroft, T. P. Satoh, M. Yamaguchi, M. Nishida, and E. O. Wiley. 2007. Mitochondrial genome and a nuclear gene indicate a novel phylogenetic position of deep-sea tube-eye fish (Stylephoridae). Ichthyological Research 54:323–332. Google Scholar

    1089.

    Miya, M., A. Kawaguchi, and M. Nishida. 2001. Mitogenomic exploration of higher teleostean phylogenies: a case study for moderate-scale evolutionary genomics with 38 newly determined complete mitochondrial DNA sequences. Molecular Biology and Evolution 18(11):1993–2009. Google Scholar

    1090.

    Miya, M., and J. Nielsen. 1991. A new species of the deepsea fish genus Parabrotula (Parabrotulidae) from Sagami Bay with notes on its ecology. Japanese Journal of Ichthyology 38(1):1–5. Google Scholar

    1091.

    Miya, M., and M. Nishida. 2015. The mitogenomic contributions to molecular phylogenetics and evolution of fishes: a 15-year retrospect. Ichthyological Research 62:29–71. Google Scholar

    1092.

    Miya, M., T. W. Pietsch, J. W. Orr, R. J. Arnold, T. P. Satoh, A. M. Shedlock, H.-C. Ho, M. Shimazaki, M. Yabe, and M. Nishida. 2010. Evolutionary history of anglerfishes (Teleostei: Lophiiformes): a mitogenomic perspective. BMC Evolutionary Biology 10:58.  https://doi.org/10.1186/1471-2148-10-58 Google Scholar

    1093.

    Miya, M., T. R. Satoh, and M. Nishida. 2005. The phylogenetic position of toadfishes (order Batrachoidiformes) in the higher ray-finned fish as inferred from partitioned Bayesian analysis of 102 whole mitochondrial genome sequences. Biological Journal of the Linnean Society 85(3):289–306. Google Scholar

    1094.

    Miya, M., M. Takahashi, H. Endo, N. B. Ishiguro, J. G. Inoue, T. Mukai, T. P. Satoh, M. Yamaguchi, A. Kawaguchi, K. Mabuchi, ET AL. 2003. Major patterns of higher teleostean phylogenies: a new perspective based on 100 complete mitochondrial DNA sequences. Molecular Phylogenetics and Evolution 26(1): 121–138. Google Scholar

    1095.

    Mo, T. 1991. Anatomy, Relationships and Systematics of the Bagridae (Teleostei: Siluroidei) with a Hypothesis of Siluroid Phylogeny. Koenigstein: Koeltz Scientific Books. 216 pp. Google Scholar

    1096.

    Mok, H.-K., and S.-C. Shen. 1983. Osteology and Phylogeny of the Squamipinnes. Taipei: Taiwan Museum. 87 pp. (Special Publication 1.) Google Scholar

    1097.

    Molina, G. I. 1782. Saggio sulla Storia Naturale del Chili. Bologna: S. Tommaso d'Aquino. 367 pp. Google Scholar

    1098.

    Møller, P. R., A. D. Jordan, P. Gravlund, and J. F. Steffensen. 2002. Phylogenetic position of the cryopelagic codfish genus Arctogadus Drjagin, 1932 based on partial mitochondrial cytochrome b sequences. Polar Biology 25:342–349. Google Scholar

    1099.

    Møller, P. R., S. W. Knudsen, W. Schwarzhans, and J. G. Nielsen. 2016. A new classification of viviparous brotulas (Bythitidae)—with family status for Dinematichthyidae—based on molecular, morphological and fossil data. Molecular Phylogenetics and Evolution 100:391–408. Google Scholar

    1100.

    Monsch, K. A., and A. F. Bannikov. 2011. New taxonomic synopses and revision of the scombroid fishes (Scombroidei, Perciformes), including billfishes, from the Cenozoic of territories of the former USSR. Earth and Environmental Science Transactions of the Royal Society of Edinburgh 102(4):253–300. Google Scholar

    1101.

    Mooi, R. D. 1990. Egg surface morphology of pseudochromoids (Perciformes: Percoidei), with comments on its phylogenetic implications. Copeia 1990(2):455–475. Google Scholar

    1102.

    Mooi, R. D., and G. D. Johnson. 1997. Dismantling the Trachinoidei: evidence of a scorpaenoid relationship for the Champsodontidae. Ichthyological Research 44:143–176. Google Scholar

    1103.

    Moore, J. A. 1993a. The phylogeny and evolution of the Trachichthyiformes (Teleostei: Percomorpha), with comments on the intrarelationships of the Acanthomorpha [dissertation]. Yale University. pp. 271. Google Scholar

    1104.

    Moore, J. A. 1993b. Phylogeny of the Trachichthyiformes (Teleostei: Percomorpha). Bulletin of Marine Science 52:114–136. Google Scholar

    1105.

    Moore, J. A., and T. J. Near. 2020a. Actinopteri. In: K. De Queiroz, P. D. Cantino, and J. A. Gauthier, eds. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC. pp. 705–708. Google Scholar

    1106.

    Moore, J. A.2020b. Actinopterygii. In: K. De Queiroz, P. D. Cantino, and J. A. Gauthier, eds. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC. pp. 695–699. Google Scholar

    1107.

    Moore, J. A.2020c. Neopterygii. In: K. De Queiroz, P. D. Cantino, and J. A. Gauthier, eds. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC. pp. 711–714. Google Scholar

    1108.

    Moore, J. A.2020d. Osteichthyes. In: K. De Queiroz, P. D. Cantino, and J. A. Gauthier, eds. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC. pp. 685–689. Google Scholar

    1109.

    Moore, J. A.2020e. Pan-Teleostei. In: K. De Queiroz, P. D. Cantino, and J. A. Gauthier, eds. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC. pp. 715–717. Google Scholar

    1110.

    Moore, J. A.2020f. Teleostei. In: K. De Queiroz, P. D. Cantino, and J. A. Gauthier, eds. Phylonyms: A Companion to the PhyloCode. Boca Raton, FL: CRC. pp. 719–723. Google Scholar

    1111.

    Moreira, D. A., P. A. Buckup, C. Furtado, A. L. Val, R. Schama, and T. E. Parente. 2017. Reducing the information gap on Loricarioidei (Siluriformes) mitochondrial genomics. BMC Genomics 18:345.  https://doi.org/10.1186/s12864-017-3709-3 Google Scholar

    1112.

    Moritz, T., and R. Britz. 2019. Revision of the extant Polypteridae (Actinopterygii: Cladistia). Ichthyological Exploration of Freshwaters 29(2):97–192. Google Scholar

    1113.

    Mthethwa, S., A. E. Bester-Van Der Merwe, and R. Roodt-Wilding. 2023a. Addressing the complex phylogenetic relationship of the Gempylidae fishes using mitogenome data. Ecology and Evolution 13:e10217. Google Scholar

    1114.

    Mthethwa, S.. 2023b. The complete mitochondrial genome of the South African snoek Thyrsites atun (Euphrasén, 1791) (Perciformes, Gempylidae). Mitochondrial DNA Part B 8:288–291. Google Scholar

    1115.

    Mu, X., Y. Yang, J. Sun, L. Yi, M. Xu, C. Shao, K. H. Chu, W. Li, C. Liu, D. Gu, ET AL. 2022. FishPIE: a universal phylogenetically informative exon markers set for ray-finned fishes. iScience 25(9):105025. Google Scholar

    1116.

    Müller, J. 1845a. Über den Bau und die Grenzen der Ganoiden und über das natürliche System der Fische. Bericht der Akademie der Wissenschaften Berlin 1844:416–422. Google Scholar

    1117.

    Müller, J. 1845a1845b. Über den Bau und die Grenzen der Ganoiden und über das natürliche System der Fische. Archiv für Naturgeschichte 11:91–141. Google Scholar

    1118.

    Müller, J., and F. H. Troschel. 1848. Fische. In: M. R. Schomburgk, ed. Reisen in Britisch-Guiana in den Jahren 1840–1844, vol. 3. Leipzig: J. J. Weber. pp. 618–644. Google Scholar

    1119.

    Muller, T. A. 2023. Taxonomic revision of the pirate perches (Percopsiformes: Aphredoderidae) reveals three new species and elevates two subspecies [dissertation]. University of Minnesota. 52 pp. ProQuest (order no. 30630998).  https://www.proquest.com/  Google Scholar

    1120.

    Muñoz, M. 2010. Reproduction in Scorpaeniformes. In: K. S. Cole, ed. Reproduction and Sexuality in Marine Fishes: Patterns and Processes. Berkeley: University of California Press. pp. 64–89. Google Scholar

    1121.

    Munroe, T. A. 2015. Systematic diversity of the Pleuronectiformes. In: R. N. Gibson, R. Nash, A. Geffen, and H. Van Der Veer, eds. Flatfishes: Biology and Exploitation. 2nd ed. Chichester, UK: Wiley Blackwell. pp. 13–51. Google Scholar

    1122.

    Murphy, W. J., and G. E. Collier. 1997. A molecular phylogeny for aplocheiloid fishes (Atherinomorpha, Cyprinodontiformes): the role of vicariance and the origins of annualism. Molecular Biology and Evolution 14:790–799. Google Scholar

    1123.

    Murray, A. M. 1996. A new Paleocene genus and species of percopsid, †Massamorichthys wilsoni (Paracanthopterygii) from Joffre Bridge, Alberta, Canada. Journal of Vertebrate Paleontology 16(4):642–652. Google Scholar

    1124.

    Murray, A. M. 2003. A new Eocene citharinoid fish (Ostariophysi: Characiformes) from Tanzania. Journal of Vertebrate Paleontology 23(3):501–507. Google Scholar

    1125.

    Murray, A. M. 2016. Mid-Cretaceous acanthomorph fishes with the description of a new species from the Turonian of Lac des Bois, Northwest Territories, Canada. Vertebrate Anatomy Morphology Palaeontology 1:101–115. Google Scholar

    1126.

    Murray, A. M. 2019. Redescription of Barbus megacephalus Günther, 1876 and Thynnichthys amblyostoma von der Marck, 1876 (Cypriniformes: Cyprinidae) from probable Eocene deposits of Southeast Asia, and an assessment of their taxonomic positions. Journal of Systematic Palaeontology 17(17):1433–1455. Google Scholar

    1127.

    Murray, A. M. 2020. Early Cenozoic cyprinoids (Ostariophysi: Cypriniformes: Cyprinidae and Danionidae) from Sumatra, Indonesia. Journal of Vertebrate Paleontology 40(1):e1762627. Google Scholar

    1128.

    Murray, A. M. 2022. Re-description and phylogenetic relationships of †Protosyngnathus sumatrensis (Teleostei: Syngnathoidei), a freshwater pipefish from the Eocene of Sumatra, Indonesia. Journal of Systematic Palaeontology 20(1):2113832. Google Scholar

    1129.

    Murray, A. M., D. B. Brinkman, M. G. Newbrey, and A. G. Neuman. 2020. Earliest North American articulated freshwater acanthomorph fish (Teleostei: Percopsiformes) from Upper Cretaceous deposits of Alberta, Canada. Geological Magazine 157(7):1087–1096. Google Scholar

    1130.

    Murray, A. M., and S. L. Cumbaa. 2013. Early Turonian acanthomorphs from Lac Des Bois, Northwest Territories, Canada. Journal of Vertebrate Paleontology 33(2):293–300. Google Scholar

    1131.

    Murray, A. M., L. E. Nelson, and D. B. Brinkman. 2023. A new sturgeon from the Upper Cretaceous Horseshoe Canyon Formation in central Alberta, Canada. Journal of Vertebrate Paleontology 43(1):e2232846. Google Scholar

    1132.

    Murray, A. M., and J. G. M. Thewissen. 2008. Eocene actinopterygian fishes from Pakistan, with the description of a new genus and species of channid (Channiformes). Journal of Vertebrate Paleontology 28(1):41–52. Google Scholar

    1133.

    Murray, A. M., and M. V. H. Wilson. 1996. A new Paleocene genus and species of percopsiform (Teleostei: Paracanthopterygii) from the Paskapoo Formation, Smoky Tower, Alberta. Canadian Journal of Earth Science 33(3):429–438. Google Scholar

    1134.

    Murray, A. M.. 1999. Contributions of fossils to the phylogenetic relationships of the percopsiform fishes (Teleostei: Paracanthopterygii): order restored. In: G. Arratia and H.-P. Schultze, eds. Mesozoic Fishes 2: Systematics and Fossil Record. Munich: Verlag Dr. Friedrich Pfeil. pp. 397–411. Google Scholar

    1135.

    Murray, A. M.. 2013. Two new paraclupeid fishes (Clupeomorpha: Ellimmichthyiformes) from the Upper Cretaceous of Morocco. In: G. Arratia, H.-P. Schultze, and M. V. H. Wilson, eds. Mesozoic Fishes 5: Global Diversity and Evolution. Munich: Verlag Dr. Friedrich Pfeil. pp. 267–290. Google Scholar

    1136.

    Murray, A. M.. 2014. Four new basal acanthomorph fishes from the Late Cretaceous of Morocco. Journal of Vertebrate Paleontology 34(1):34–48. Google Scholar

    1137.

    Murray, A. M., D. K. Zelenitsky, D. B. Brinkman, and A. G. Neuman. 2018. Two new Palaeocene osteoglossomorphs from Canada, with a reassessment of the relationships of the genus †Joffrichthys, and analysis of diversity from articulated versus microfossil material. Zoological Journal of the Linnean Society 183(4):907–944. Google Scholar

    1138.

    Musilova, Z., F. Cortesi, M. Matschiner, W. I. L. Davies, J. S. Patel, S. M. Stieb, F. De Busserolles, M. Malmstrøm, O. K. Tørresen, C. J. Brown, ET AL. 2019. Vision using multiple distinct rod opsins in deepsea fishes. Science 364(6440):588–592. Google Scholar

    1139.

    Myers, G. S., and C. B. Wade. 1946. New fishes of the families Dactyloscopidae, Microdesmidae, and Antennariidae from the west coast of Mexico and the Galapagos Islands, with a brief account of the use of rotenone fish poisons in ichthyological collecting. Allan Hancock Pacific Expeditions 9, no. 6. Los Angeles: University of Southern California Press. pp. 151–179. (Allan Hancock Foundation Publications of the University of Southern California Series 1.) Google Scholar

    1140.

    Nakae, M., and K. Sasaki. 2010. Lateral line system and its innervation in Tetraodontiformes with outgroup comparisons: descriptions and phylogenetic implications. Journal of Morphology 271(5):559–579. Google Scholar

    1141.

    Nakatani, M., M. Miya, K. Mabuchi, K. Saitoh, and M. Nishida. 2011. Evolutionary history of Otophysi (Teleostei), a major clade of the modern freshwater fishes: Pangaean origin and Mesozoic radiation. BMC Evolutionary Biology 11:177.  https://doi.org/10.1186/1471-2148-11-177 Google Scholar

    1142.

    Nam, K.-S., and M. V. Nazarkin. 2018. Fossil prowfish, Zaprora koreana, sp. nov. (Pisces, Zaproridae), from the Neogene of South Korea. Journal of Vertebrate Paleontology 38(5):e1514616.  https://doi.org/10.1080/02724634.2018.1514616 Google Scholar

    1143.

    Nazarkin, M. V. 1997. A new genus and species of greenling (Hexagrammidae) from the Miocene of Sakhalin Island. Journal of Ichthyology 37(1):5–13. Google Scholar

    1144.

    Nazarkin, M. V. 2002. Gunnels (Perciformes, Pholidae) from the Miocene of Sakhalin Island. Journal of Ichthyology 42(4):279–288. Google Scholar

    1145.

    Nazarkin, M. V., G. Carnevale, and A. F. Bannikov. 2013. A new greenling (Teleostei, Cottoidei) from the Miocene of Sakhalin Island, Russia. Journal of Vertebrate Paleontology 33(4):794–803. Google Scholar

    1146.

    Near, T. J., D. I. Bolnick, and P. C. Wainwright. 2004. Investigating phylogenetic relationships of sunfishes and black basses (Actinopterygii: Centrarchidae) using DNA sequences from mitochondrial and nuclear genes. Molecular Phylogenetics and Evolution 32(1):344–357. Google Scholar

    1147.

    Near, T. J., A. Dornburg, R. I. Eytan, B. P. Keck, W. L. Smith, K. L. Kuhn, J. A. Moore, S. A. Price, F. T. Burbrink, M. Friedman, and P. C. Wainwright. 2013. Phylogeny and tempo of diversification in the superradiation of spiny-rayed fishes. Proceedings of the National Academy of Sciences of the United States of America 110(31):12738–12743. Google Scholar

    1148.

    Near, T. J., A. Dornburg, and M. Friedman. 2014. Phylogenetic relationships and timing of diversification in gonorynchiform fishes inferred using nuclear gene DNA sequences (Teleostei: Ostariophysi). Molecular Phylogenetics and Evolution 80:297–307. Google Scholar

    1149.

    Near, T. J., A. Dornburg, R. C. Harrington, C. Oliveira, T. W. Pietsch, C. E. Thacker, T. P. Satoh, E. Katayama, P. C. Wainwright, ET AL. 2015. Identification of the notothenioid sister lineage illuminates the biogeographic history of an Antarctic adaptive radiation. BMC Evolutionary Biology 15:109.  https://doi.org/10.1186/s12862-015-0362-9 Google Scholar

    1150.

    Near, T. J., A. Dornburg, K. L. Kuhn, J. T. Eastman, J. N. Pennington, T. Patarnello, L. Zane, D. A. Fernández, and C. D. Jones. 2012. Ancient climate change, antifreeze, and the evolutionary diversification of Antarctic fishes. Proceedings of the National Academy of Sciences of the United States of America 109(9):3434–3439. Google Scholar

    1151.

    Near, T. J., A. Dornburg, M. Tokita, D. Suzuki, M. C. Brandley, and M. Friedman. 2014. Boom and bust: ancient and recent diversification in bichirs (Polypteridae: Actinopterygii), a relictual lineage of ray-finned fishes. Evolution 68(4):1014–1026. Google Scholar

    1152.

    Near, T. J., R. I. Eytan, A. Dornburg, K. L. Kuhn, J. A. Moore, M. P. Davis, P. C. Wainwright, M. Friedman, and W. L. Smith. 2012. Resolution of ray-finned fish phylogeny and timing of diversification. Proceedings of the National Academy of Sciences of the United States of America 109(34):13698–13703. Google Scholar

    1153.

    Near, T. J., and D. Kim. 2021. Phylogeny and time scale of diversification in the fossil-rich sunfishes and black basses (Teleostei: Percomorpha: Centrarchidae). Molecular Phylogenetics and Evolution 161:107156.  https://doi.org/10.1016/j.ympev.2021.107156 Google Scholar

    1154.

    Near, T. J., D. J. Macguigan, E. Parker, C. D. Struthers, C. D. Jones, and A. Dornburg. 2018. Phylogenetic analysis of Antarctic notothenioids illuminates the utility of RADseq for resolving Cenozoic adaptive radiations. Molecular Phylogenetics and Evolution 129:268–279. Google Scholar

    1155.

    Near, T. J., J. J. Pesavento, and C.-H. C. Cheng. 2004. Phylogenetic investigations of Antarctic notothenioid fishes (Perciformes: Notothenioidei) using complete gene sequences of the mitochondrial encoded 16S rRNA. Molecular Phylogenetics and Evolution 32(3):881–891. Google Scholar

    1156.

    Near, T. J., M. Sandel, K. L. Kuhn, P. J. Unmack, P. C. Wainwright, and W. L. Smith. 2012. Nuclear gene-inferred phylogenies resolve the relationships of the enigmatic pygmy sunfishes, Elassoma (Teleostei: Percomorpha). Molecular Phylogenetics and Evolution 63(2):388–395. Google Scholar

    1157.

    Near, T. J., and C. E. Thacker. 2023. Phylogenetic classification of living and fossil ray-finned fishes (Actinopterygii) [online dataset]. Dryad.  https://doi.org/10.5061/dryad.80gb5mkwp Google Scholar

    1158.

    Neilson, M. E., and C. A. Stepien. 2009. Escape from the Ponto-Caspian: evolution and biogeography of an endemic goby species flock (Benthophilinae: Gobiidae: Teleostei). Molecular Phylogenetics and Evolution 52(1):84–102. Google Scholar

    1159.

    Nelson, G. J. 1967. Gill arches of teleostean fishes of the family Clupeidae. Copeia 1967(2):389–399. Google Scholar

    1160.

    Nelson, G. J. 1968. Gill arches of teleosteanfishes of the division Osteoglossomorpha. Journal of the Linnean Society, Zoology 47(312):261–277. Google Scholar

    1161.

    Nelson, G. J. 1969a. Gill arches and the phylogeny of fishes, with notes on the classification of vertebrates. Bulletin of the American Museum of Natural History 141:475–552. Google Scholar

    1162.

    Nelson, G. J. 1969b. Infraorbital bones and their bearing on the phylogeny and geography of osteoglossomorph fishes. American Museum Novitates 2394:1–37. Google Scholar

    1163.

    Nelson, G. J. 1969c. Origin and diversification of teleostean fishes. Annals of the New York Academy of Sciences 167(1):18–30. Google Scholar

    1164.

    Nelson, G. J. 1970a. Gill arches of some teleostean fishes of the families Salangidae and Argentinidae. Japanese Journal of Ichthyology 17(2):61–66. Google Scholar

    1165.

    Nelson, G. J. 1970b. The hyobranchial apparatus of teleostean fishes of the families Engraulidae and Chirocentridae. American Museum Novitates 2410:1–30. Google Scholar

    1166.

    Nelson, G. J. 1972. Cephalic sensory canals, pitlines, and the classification of esocoid fishes, with notes on galaxiids and other teleosts. American Museum Novitates 2492:1–49. Google Scholar

    1167.

    Nelson, G. J. 1973. Relationships of clupeomorphs with remarks on the structure of the lower jaw in fishes. In: P. H. Greenwood, R. S. Miles, and C. Patterson, eds. Interrelationships of Fishes. London: Academic Press. pp. 333–349. Google Scholar

    1168.

    Nelson, G. J. 1989. Phylogeny of major fish groups. In: B. Fernholm, K. Bremer, and H. Jôrnvall, eds. The Hierarchy of Life: Molecules and Morphology in Phylogenetic Analysis. Amsterdam: Elsevier. pp. 325–336. Google Scholar

    1169.

    Nelson, J. S. 1976. Fishes of the World, 1st ed. New York: Wiley. 416 pp. Google Scholar

    1170.

    Nelson, J. S. 1978. Review of The Biology of Sticklebacks, by R. J. Wootton. Copeia 1978(3):552–554. Google Scholar

    1171.

    Nelson, J. S. 1984. Fishes of the World, 2nd ed. New York: Wiley. 523 pp. Google Scholar

    1172.

    Nelson, J. S. 1986. Some characters of Trichonotidae, with emphasis to those distinguishing it from Creediidae (Perciformes, Trachinoidei). Japanese Journal of Ichthyology 33(1):1–6. Google Scholar

    1173.

    Nelson, J. S. 1994. Fishes of the World, 3rd ed. New York: Wiley. 600 pp. Google Scholar

    1174.

    Nelson, J. S. 2006. Fishes of the World, 4th ed. Hoboken, NJ: John Wiley. 601 pp. Google Scholar

    1175.

    Nelson, J. S., T. C. Grande, and M. V. H. Wilson. 2016. Fishes of the World, 5th ed. Hoboken, NJ: John Wiley and Sons. 707 pp. Google Scholar

    1176.

    Newbrey, M. G., and T. Konishi. 2015. A new lizardfish (Teleostei, Aulopiformes) from the Late Cretaceous Bearpaw Formation of Alberta, Canada, with a revised diagnosis of Apateodus (Aulopiformes, Ichthyotringoidei). Journal of Vertebrate Paleontology 35(3):e918042.  https://doi.org/10.1080/02724634.2014.918042 Google Scholar

    1177.

    Newbrey, M. G., A. M. Murray, M. V. H. Wilson, D. B. Brinkman, and A. G. Neuman. 2013. A new species of the paracanthopterygian Xenyllion (Sphenocephaliformes) from the Mowry Formation (Cenomanian) of Utah, USA. In: G. Arratia, H.-P. Schultze, and M. V. H. Wilson, eds. Mesozoic Fishes 5: Global Diversity and Evolution. Munich: Verlag Dr. Fredrich Pfeil. pp. 363–384. Google Scholar

    1178.

    Niebuhr, C. 1775. Descriptiones animalium avium, amphibiorum, piscium, insectorum, vermium; quae in itinere orientali observavit Petrus Forskål. Post mortem auctoris edidit Carsten Niebuhr. Hauniæ: ex officina Mölleri. 164 pp. Google Scholar

    1179.

    Nielsen, J. G. 1968. Re-description and reassignment of Parabrotula and Leucobrotula (Pisces, Zoarcidae). Videnskabelige Meddelelser fra Dansk Naturhistorisk Forening, Kjøbenhavnsk 131:225–250. Google Scholar

    1180.

    Nielsen, J. G., J. Badcock, and N. R. Merrett. 1990. New data elucidating the taxonomy and ecology of the Parabrotulidae (Pisces: Zoarcoidei). Journal of Fish Biology 37(3):437–448. Google Scholar

    1181.

    Nielsen, J. G., D. M. Cohen, D. F. Markle, and C. R. Robins. 1999. Ophidiiform Fishes of the World (Order Ophidiiformes). Vol. 18 of FAO Species Catalogue. Rome: Food and Agriculture Organization of the United Nations. 178 pp. (FAO Fisheries Synopsis 125.) Google Scholar

    1182.

    Niemiller, M. L., B. M. Fitzpatrick, P. Shah, L. Schmitz, and T. J. Near. 2013. Evidence for repeated loss of selective constraint in rhodopsin of amblyopsid cavefishes (Teleostei: Amblyopsidae). Evolution 67:732–748. Google Scholar

    1183.

    Nolf, D. 1972. Les otolithes du Calcaire Grossier à Fercourt (Éocène du Bassin de Paris). Bulletin de la Société Belge de Géologie, de Paléontologie et d'Hydrologie 81:139–157. Google Scholar

    1184.

    Nolf, D. 1988. Les Otolithes de Téléostéens Éoeènes d'Aquitaine (Sud-Ouest de la France) et Leur Intérêt Stratigraphique. Brussels: Académie Royale de Belgique. 147 pp. (Mémoires de la Classe des Sciences 19.) Google Scholar

    1185.

    Nolf, D., and H. Lapierre. 1979. Otolithes de Poissons nouveaux ou peu connus du Calcaire Grossier et de la formation d'Auvers (Éocène du Bassin parisien). Bulletin du Muséum National d'Histoire Naturelle, Section C, Sciences de la Terre, Paléontologie, Géologie, Minéralogie, Series 4, 1(2):79–125. Google Scholar

    1186.

    Nolf, D., R. S. Rana, and H. Singh. 2006. Fish otoliths from the Ypresian (early Eocene) of Vastan, Gujarat, India. Bulletin de l'Institut Royal des Sciences Naturelles de Belgique Sciences de la Terre 76:105–118. Google Scholar

    1187.

    Nolf, D., and E. Steurbaut. 1989. Evidence from otoliths for establishing relationships within gadiforms. In: D. M. Cohen, ed. Papers on the Systematics of Gadiform Fishes. Los Angeles: Natural History Museum of Los Angeles County. pp. 89–111. Google Scholar

    1188.

    Nolf, D., and G. Stringer. 1996. Cretaceous fish otoliths—a synthesis of the North American record. In: G. Arratia and G. Viohl, eds. Mesozoic Fishes: Systematics and Paleoecology. Munich: Verlag Dr. Friedrich Pfeil. pp. 433–459. Google Scholar

    1189.

    Norman, J. R. 1934. A Systematic Monograph of the Flatfishes (Heterosomata). Vol. I of Psettodidæ, Bothidæ, Pleruonectidæ. London: Printed by order of the Trustees of the British Museum. 459 pp. Google Scholar

    1190.

    Norman, J. R. 1938a. Coast fishes. Part III. The Antarctic zone. Discovery Reports 18:1–104. Google Scholar

    1191.

    Norman, J. R. 1938b. On the affinities of the Chilean fish, Normanichthys crockeri Clark. Copeia 1938(1):29–32. Google Scholar

    1192.

    Normark, B. B., A. R. Mccune, and R. G. Harrison. 1991. Phylogenetic relationships of neopterygian fishes inferred from mitochondrial DNA sequences. Molecular Biology and Evolution 8:819–834. Google Scholar

    1193.

    Nybelin, O. 1971. On the caudal Skeleton in Elops with remarks on other teleostean fishes. Acta Regiae Societatis Scientiarum et Litterarum Gothoburgensis Zoologica 7:1–52. Google Scholar

    1194.

    Nyegaard, M., E. Sawai, N. Gemmell, J. Gillum, N. R. Loneragan, Y. Yamanoue, and A. L. Stewart. 2018. Hiding in broad daylight: molecular and morphological data reveal a new ocean sunfish species (Tetraodontiformes: Molidae) that has eluded recognition. Zoological Journal of the Linnean Society 182(3):631–658. Google Scholar

    1195.

    Obermiller, L. E., and E. Pfeiler. 2003. Phylogenetic relationships of elopomorph fishes inferred from mitochondrial ribosomal DNA sequences. Molecular Phylogenetics and Evolution 26(2):202–214. Google Scholar

    1196.

    Oelschläger, H. 1983. Vergleichende und funktionelle anatomie der Allotriognathi (=Lampridiformes), ein beitrag zur evolutionsmorphologie der knochenfische. Abhandlungen der Senckenbergischen Naturforschenden Gesellschaft 541:1–127. Google Scholar

    1197.

    Oh, D.-J., J.-C. Lee, Y.-M. Ham, and Y.-H. Jung. 2021. The mitochondrial genome of Stereolepis doederleini (Pempheriformes: Polyprionidae) and mitogenomic phylogeny of Pempheriformes. Genetics and Molecular Biology 44:e20200166.  https://doi.org/10.1590/1678-4685-GMB-2020-0166 Google Scholar

    1198.

    Ohashi, S. 2018. Morphology of a unique ophidiid, Hypopleuron caninum Radcliffe 1913 (Ophidiiformes, Ophidiidae, Neobythitinae), suggesting a close relationship with the family Carapidae. Zootaxa 4521(4):499–538.  https://doi.org/10.11646/zootaxa.4521.4.2 Google Scholar

    1199.

    Okamura, O. 1989. Relationships of the suborder Macrouroidei and related groups, with comments on Merlucciidae and Steindachneria. In: D. M. Cohen, ed. Papers on the Systematics of Gadiform Fishes. Los Angeles: Natural History Museum of Los Angeles County. pp. 129–142. Google Scholar

    1200.

    Okiyama, M. 1984. Myctophiformes: relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 254–259. (Special Publication 1.) Google Scholar

    1201.

    Oliveira, C., G. S. Avelino, K. T. Abe, T. C. Mariguela, R. C. Benine, G. Ortí, R. P. Vari, and R. M. Corrêa E Castro. 2011. Phylogenetic relationships within the speciose family Characidae (Teleostei: Ostariophysi: Characiformes) based on multilocus analysis and extensive ingroup sampling. BMC Evolutionary Biology 11:275.  https://doi.org/10.1186/1471-2148-11-275  Google Scholar

    1202.

    Olney, J. E., G. D. Johnson, and C. C. Baldwin. 1993. Phylogeny of lampridiform fishes. Bulletin of Marine Science 52:137–169. Google Scholar

    1203.

    Olsen, P. E. 1984. The skull and pectoral girdle of the parasemionotid fish Watsonulus eugnathoides from the early Triassic Sakamena group of Madagascar, with comments on the relationships of the holostean fishes. Essays Presented to Dr. Bobb Schaeffer, Journal of Vertebrate Paleontology 4(3):481–499. Google Scholar

    1204.

    Olsen, P. E., and A. R. Mccune. 1991. Morphology of the Semionotus elegans species group from the early Jurassic part of the Newark supergroup of eastern North America with comments on the family Semiontidae (Neopterygii). Journal of Vertebrate Paleontology 11(3):269–292. Google Scholar

    1205.

    Orr, J. W. 1995. Phylogenetic relationships of gasterosteiform fishes (Teleostei: Acanthomorpha) [dissertation]. University of Washington. 813 pp. Google Scholar

    1206.

    Orrell, T. M., B. B. Collette, and G. D. Johnson. 2006. Molecular data support separate scombroid and xiphioid clades. Bulletin of Marine Science 79:505–519. Google Scholar

    1207.

    Ortí, G. 1997. Radiation of characiform fishes: evidence from mitochondrial and DNA sequences. In: T. D. Kocher and C. A. Stepien, eds. Molecular Systematics of Fishes. San Diego: Academic Press. pp. 219–314. Google Scholar

    1208.

    Ortí, G., and A. Meyer. 1996. Molecular evolution of ependymin and the phylogenetic resolution of early divergences among euteleost fishes. Molecular Biology and Evolution 13:556–573. Google Scholar

    1209.

    Ortí, G.. 1997. The radiation of characiform fishes and the limits of resolution of mitochondrial ribosomal DNA sequences. Systematic Biology 46(1):75–100. Google Scholar

    1210.

    Ørvig, T. O. R. 1978. Microstructure and growth of the dermal skeleton in fossil actinopterygian fishes: Nephrotus and Colobodus, with remarks on the dentition in other forms. Zoologica Scripta 7(1–4):297–326. Google Scholar

    1211.

    Osinov, A., and V. Lebedev. 2004. Salmonid fishes (Salmonidae, Salmoniformes): the systematic position in the superorder Protacanthopterygii, the main stages of evolution, and molecular dating. Journal of Ichthyology 44:690–715. Google Scholar

    1212.

    Östlund-Nilsson, S., and G. E. Nilsson. 2004. Breathing with a mouth full of eggs: respiratory consequences of mouthbrooding in cardinalfish. Proceedings of the Royal Society B, Biological Sciences 271(1543):1015–1022. Google Scholar

    1213.

    Otero, O. 2004. Anatomy, systematics and phylogeny of both Recent and fossil latid fishes (Teleostei, Perciformes, Latidae). Zoological Journal of the Linnean Society 141(1):81–133. Google Scholar

    1214.

    Otero, O., and M. Gayet. 1995. Étude phylogénétique des Aipichthyides; poissons téléostéens de la Téthys cénomanienne. Géobios 28(Suppl. 2):221–224. Google Scholar

    1215.

    Otero, O.. 1996. Anatomy and phylogeny of the Aipichthyoidea nov. of the Cenomanian Tethys and their place in the Acanthomorpha (Teleostei). Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 202:313–344. Google Scholar

    1216.

    Otero, O., A. Pinton, H. T. Mackaye, A. Likius, P. Vignaud, and M. Brunet. 2009. First description of a Pliocene ichthyofauna from Central Africa (site KL2, Kolle area, Eastern Djurab, Chad): what do we learn? Journal of African Earth Sciences 54:62–74. Google Scholar

    1217.

    O'toole, B. 2002. Phylogeny of the species of the superfamily Echeneoidea (Perciformes: Carangoidei: Echeneidae, Rachycentridae, and Coryphaenidae), with an interpretation of echeneid hitchhiking behaviour. Canadian Journal of Zoology 80(4):596–623. Google Scholar

    1218.

    Pallas, P. S. 1769. Spicilegia zoologica: quibus novae imprimis et obscurae animalium species iconibus, descriptionibus atque commentariis illustrantur, Volume 1 (fasc. 7). Berolini: Prostant apud Gottl. August. Lange. 42 pp. Google Scholar

    1219.

    Pallas, P. S. 1810. Labraces, novum genus piscium, oceani orientalis. Mémoires de l'Académie Impériale des Sciences de St. Pétersbourg 2:382–398. Google Scholar

    1220.

    Pan, Q., R. Feron, E. Jouanno, H. Darras, A. Herpin, B. Koop, E. Rondeau, F. W. Goetz, W. A. Larson, L. Bernatchez, ET AL. 2021. The rise and fall of the ancient northern pike master sex-determining gene. eLife 10:e62858.  https://doi.org/10.7554/eLife.62858  Google Scholar

    1221.

    Parenti, L. R. 1981. A phylogenetic and biogeographic analysis of cyprinodontiform fishes (Teleostei: Atherinomorpha). Bulletin of the American Museum of Natural History 168:335–557. Google Scholar

    1222.

    Parenti, L. R. 1984. On the relationships of phallostethid fishes (Atherinomorpha): with notes on the anatomy of Phallostethus dunckeri Regan, 1913. American Museum Novitates 2779:1–12. Google Scholar

    1223.

    Parenti, L. R. 1986. The phylogenetic significance of bone types in euteleost fishes. Zoological Journal of the Linnean Society 87(1):37–51. Google Scholar

    1224.

    Parenti, L. R. 1987. Phylogenetic aspects of tooth and jaw structure of the Medaka, Oryzias latipes, and other beloniform fishes. Journal of Zoology 211(3):561–572. Google Scholar

    1225.

    Parenti, L. R. 1989. A phylogenetic revision of the phallostethid fishes (Atherinomorpha, Phallostethidae). Proceedings of the California Academy of Sciences, 4th series, 46(11):243–277. Google Scholar

    1226.

    Parenti, L. R. 1993. Relationships of atherinomorph fishes (Teleostei). Bulletin of Marine Science 52:170–196. Google Scholar

    1227.

    Parenti, L. R. 2005. The phylogeny of atherinomorphs: evolution of a novel fish reproductive system. In: M. C. Uribe, and H. J. Grier, eds. Viviparous Fishes. Homestead, FL: New Life. pp. 13–30. Google Scholar

    1228.

    Parenti, L. R. 2008. A phylogenetic analysis and taxonomic revision of ricefishes, Oryzias and relatives (Beloniformes, Adrianichthyidae). Zoological Journal of the Linnean Society 154(3):494–610. Google Scholar

    1229.

    Parenti, L. R., and H. J. Grier. 2004. Evolution and phylogeny of gonad morphology in bony fishes. Integrative and Comparative Biology 44(5):333–348. Google Scholar

    1230.

    Parenti, P., and J. E. Randall. 2020. An annotated checklist of the fishes of the family Serranidae of the world with description of two new related families of fishes. FishTaxa 15:1–170. Google Scholar

    1231.

    Parey, E., A. Louis, J. Montfort, O. Bouchez, C. Roques, C. Iampietro, J. Lluch, A. Castinel, C. Donnadieu, T. Desvignes, ET AL. 2023. Genome structures resolve the early diversification of teleost fishes. Science 379(6632):572–575. Google Scholar

    1232.

    Parin, N. V., I. B. Shakhovskoy, K. E. Bemis, and B. B. Collette. 2019. Family Exocoetidae. In: B. B. Collette, K. E. Bemis, N. V. Parin, and I. B. Shakhovskoy. Order Beloniformes: Needlefishes, Sauries, Halfbeaks, and Flyingfishes. B. B. Collette and T. J. Near, eds. New Haven, CT: Peabody Museum of Natural History, Yale University. pp. 149–239. (Memoir 1, Fishes of the Western North Atlantic 10.) Google Scholar

    1233.

    Parker, A., and I. Kornfield. 1995. Molecular perspective on evolution and zoogeography of cyprinodontid killifishes (Teleostei: Atherinomorpha). Copeia 1995(1):8–21. Google Scholar

    1234.

    Parker, E., A. Dornburg, C. D. Struthers, C. D. Jones, and T. J. Near. 2022. Phylogenomic species delimitation dramatically reduces species diversity in an Antarctic adaptive radiation. Systematic Biology 71(1):58–77. Google Scholar

    1235.

    Parker, E., and T. J. Near. 2022. Phylogeny reconciles classification in Antarctic plunderfishes. Ichthyology and Herpetology 110(4):662–674. Google Scholar

    1236.

    Parr, A. E. 1951. Preliminary revision of the Alepocephalidae, with the introduction of a new family, Searsidae. American Museum Novitates 1531:1–21. Google Scholar

    1237.

    Pastana, M. N. L., G. D. Johnson, and A. Datovo. 2022. Comprehensive phenotypic phylogenetic analysis supports the monophyly of stromateiform fishes (Teleostei: Percomorphacea). Zoological Journal of the Linnean Society 195(3):841–963. Google Scholar

    1238.

    Patterson, C. 1964. A review of Mesozoic acanthopterygian fishes, with special reference to those of the English Chalk. Philosophical Transactions of the Royal Society of London Series B, Biological Sciences 247(739):213–482. Google Scholar

    1239.

    Patterson, C. 1967. New Cretaceous berycoid fishes from the Lebanon. Bulletin of The British Museum (Natural History) Geology 14(3):67–109. Google Scholar

    1240.

    Patterson, C. 1968. The caudal skeleton in Mesozoic acanthopterygian fishes. Bulletin of the British Museum (Natural History), Geology 17(2):47–102. Google Scholar

    1241.

    Patterson, C. 1970. A clupeomorph fish from the Gault (Lower Cretaceous). Zoological Journal of the Linnean Society 49(3):161–182. Google Scholar

    1242.

    Patterson, C. 1973. Interrelationships of holosteans. In: P. H. Greenwood, R. S. Miles, and C. Patterson, eds. Interrelationships of Fishes. London: Academic Press. pp. 233–305. Google Scholar

    1243.

    Patterson, C. 1975. Braincase of pholidophorid and leptolepid fishes, with a review of the actinopterygian braincase. Philosophical Transactions of the Royal Society of London Series B, Biological Sciences 269(899):275–579. Google Scholar

    1244.

    Patterson, C. 1977. The contribution of paleontology to teleostean phylogeny. In: P. C. Hecht, P. C. Goody, and B. M. Hecht, eds. Major Patterns in Vertebrate Evolution. New York: Plenum Press. pp. 579–643. Google Scholar

    1245.

    Patterson, C. 1982. Morphology and interrelationships of primitive actinopterygian fishes. American Zoologist 22(2):241–259. Google Scholar

    1246.

    Patterson, C. 1984a. Chanoides, a marine Eocene otophysan fish (Teleostei: Ostariophysi). Journal of Vertebrate Paleontology 4(3):430–456. Google Scholar

    1247.

    Patterson, C. 1984b. Family Chanidae and other teleostean fishes as living fossils. In: N. Eldredge, and S. M. Stanley, eds. Living Fossils. New York: Springer. pp. 132–139. Google Scholar

    1248.

    Patterson, C. 1993. An overview of the early fossil record of acanthomorphs. Bulletin of Marine Science 52:29–59. Google Scholar

    1249.

    Patterson, C. 1994. Bony fishes. In: D. R. Prothero and R. M. Schoch, eds. Major Features of Vertebrate Evolution. Knoxville, TN: Paleontological Society. pp. 57–84. Google Scholar

    1250.

    Patterson, C. 1997. Peter Humphry Greenwood. 21 April 1927–3 March 1995. Biographical Memoirs of Fellows of the Royal Society 43:195–213. Google Scholar

    1251.

    Patterson, C., and G. D. Johnson. 1995. The Intermuscular Bones and Ligaments of Teleostean Fishes. Washington, DC: Smithsonian Institution Press. 83 pp. (Smithsonian Contributions to Zoology 559.) Google Scholar

    1252.

    Patterson, C., and D. E. Rosen. 1977. Review of ichthyodectiform and other Mesozoic teleost fishes and the theory and practice of classifying fossils. Bulletin of the American Museum of Natural History 158:85–172. Google Scholar

    1253.

    Patterson, C.. 1989. The Paracanthopterygii revisited: order and disorder. In: D. M. Cohen, ed. Papers on the Systematics of Gadiform Fishes. Los Angeles: Natural History Museum of Los Angeles County. pp. 5–36. Google Scholar

    1254.

    Paucă, M. 1929. Vorläufige Mitteilung über eine fossile Fischfauna aus den Oligozänschiefern von Sŭslanesti (Muscel). Bulletin de la Section Scientifique Académie Roumaine 12(4–5):112–120. Google Scholar

    1255.

    Paxton, J. R., E. H. Ahlstrom, and H. G. Moser. 1984. Myctophidae: relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 239–244. (Special Publication 1.) Google Scholar

    1256.

    Paxton, J. R., and P. A. Hulley. 1999a. Myctophidae. Lanternfishes. In: K. E. Carpenter and V. H. Niem, eds. Batoid Fishes, Chimaeras and Bony Fishes Part 1 (Elopidae to Linophrynidae), Vol. 3 of FAO Species Identification Guide for Fishery Purposes: The Living Marine Resources of the Western Central Pacific. Rome: Food and Agriculture Organization of the United Nations. pp. 1957–1965. Google Scholar

    1257.

    Paxton, J. R.1999b. Neoscopelidae. Neoscopelids. In: K. E. Carpenter and V. H. Niem, eds. Batoid Fishes, Chimaeras and Bony Fishes Part 1 (Elopidae to Linophrynidae), Vol. 3 of FAO Species Identification Guide for Fishery Purposes: The Living Marine Resources of the Western Central Pacific. Rome: Food and Agriculture Organization of the United Nations. pp. 1955–1956. Google Scholar

    1258.

    Paxton, J. R., G. D. Johnson, and T. Trnski. 2001. Larvae and juveniles of the deepsea “whalefishes” Barbourisia and Rondeletia (Stephanoberyciformes: Barbourisiidae, Rondeletiidae), with comments on family relationships. Records of the Australian Museum 53(3):407–425. Google Scholar

    1259.

    [PBDB] The Paleobiology Database. 2023. Cottus otiakensis [online dataset]. Accessed 15 November 2022.  https://paleobiodb.org  Google Scholar

    1260.

    Peng, Z. G., S. P. He, J. Wang, W. Wang, and R. Diogo. 2006. Mitochondrial molecular clocks and the origin of the major otocephalan clades (Pisces: Teleostei): a new insight. Gene 370:113–124. Google Scholar

    1261.

    Pérez-Rodríguez, R., O. Domínguez-Domínguez, A. F. Mar-Silva, I. Doadrio, and G. Pérez-Ponce De León. 2016. The historical biogeography of the southern group of the sucker genus Moxostoma (Teleostei: Catostomidae) and the colonization of central Mexico. Zoological Journal of the Linnean Society 177(3):633–647. Google Scholar

    1262.

    Peskin, B., K. Henke, N. Cumplido, S. Treaster, M. P. Harris, M. Bagnat, and G. Arratia. 2020. Notochordal signals establish phylogenetic identity of the teleost spine. Current Biology 30(14):2805–2814. Google Scholar

    1263.

    Peters, W. 1852. Diagnosen von neuen Flussfischen aus Mossambique. Monatsberichte der Königlichen Preussischen Akademie der Wissenschaften zu Berlin 1852:275–276, 681–685. Google Scholar

    1264.

    Peters, W. 1876. Über eine merkwürdige von Hrn. Professor Dr. Buchholz entdeckte neue Gattung von Süsswasserfischen, Pantodon Buchholzi, welche zugleich eine neue, den Malacopterygii abdominales angehörige Gruppe von Fischen, Pantodontes, repräsentirt. Monatsberichte der Königlichen Preussischen Akademie der Wissenschaften zu Berlin 1876:195–200. Google Scholar

    1265.

    Peterson, R. D., J. P. Sullivan, C. D. Hopkins, A. Santaquiteria, C. B. Dillman, S. Pirro, R. Betancur-R, D. Arcila, L. C. Hughes, and G. Ortí. 2022. Phylogenomics of bony-tongue fishes (Osteoglossomorpha) shed light on the craniofacial evolution and biogeography of the weakly electric clade (Mormyridae). Systematic Biology 71(5):1032–1044. Google Scholar

    1266.

    Pfaff, C., R. Zorzin, and J. Kriwet. 2016. Evolution of the locomotory system in eels (Teleostei: Elopomorpha). BMC Evolutionary Biology 16:159. https://doi.org/10.1186/s12862-016-0728-7 Google Scholar

    1267.

    Pfeiffer, W. 1977. The distribution of fright reaction and alarm substance cells in fishes. Copeia 1977(4):653–665. Google Scholar

    1268.

    Pickett, B. D., E. M. Wallace, P. G. Ridge, and J. S. K. Kauwe. 2020. Lingering taxonomic challenges hinder conservation and management of global bonefishes. Fisheries 45(7):347–358. Google Scholar

    1269.

    Pietsch, T. W. 1978. Evolutionary relationships of the sea moths (Teleostei: Pegasidae) with a classification of the gasterosteiform families. Copeia 1978(3):517–529. Google Scholar

    1270.

    Pietsch, T. W. 1981. The osteology and relationships of the anglerfish genus Tetrabrachium with comments on lophiiform classification. Fishery Bulletin 79(3):387–419. Google Scholar

    1271.

    Pietsch, T. W. 1984. Lophiiformes: development and relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 320–325. (Special Publication 1.) Google Scholar

    1272.

    Pietsch, T. W. 1989. Phylogenetic relationships of trachinoid fishes of the family Uranoscopidae. Copeia 1989(2):253–303. Google Scholar

    1273.

    Pietsch, T. W. 2005. Dimorphism, parasitism, and sex revisited: modes of reproduction among deep-sea ceratioid anglerfishes (Teleostei: Lophiiformes). Ichthyological Research 52:207–236. Google Scholar

    1274.

    Pietsch, T. W. 2009. Oceanic Anglerfishes: Extraordinary Diversity in the Deep Sea. Berkeley: University of California Press. 557 pp. Google Scholar

    1275.

    Pietsch, T. W., and R. J. Arnold. 2020. Frogfishes: Biodiversity, Zoogeography, and Behavioral Ecology. Baltimore: John Hopkins University Press. 601 pp. Google Scholar

    1276.

    Pietsch, T. W., and G. Carnevale. 2011. A new genus and species of anglerfish (Teleostei: Lophiiformes: Lophiidae) from the Eocene of Monte Bolca, Italy. Copeia 2011(1):64–71. Google Scholar

    1277.

    Pietsch, T. W., and D. B. Grobecker. 1987. Frogfishes of the World. Stanford, CA: Stanford University Press. 420 pp. Google Scholar

    1278.

    Pietsch, T. W., and J. W. Orr. 2007. Phylogenetic relationships of deep-sea anglerfishes of the suborder Ceratioidei (Teleostei: Lophiiformes) based on morphology. Copeia 2007(1):1–34. Google Scholar

    1279.

    Pietsch, T. W., and C. P. Zabetian. 1990. Osteology and interrelationships of the sand lances (Teleostei: Ammodytidae). Copeia 1990(1):78–100. Google Scholar

    1280.

    Piller, K. R., E. Parker, A. R. Lemmon, and E. M. Lemmon. 2022. Investigating the utility of anchored hybrid enrichment data to investigate the relationships among the killifishes (Actinopterygii: Cyprinodontiformes), a globally distributed group of fishes. Molecular Phylogenetics and Evolution 173:107482.  https://doi.org/10.1016/j.ympev.2022.107482 Google Scholar

    1281.

    Pinion, A. K., R. Britz, K. M. Kubicek, D. S. Siegel, and K. W. Conway. 2023. The larval attachment organ of the bowfin Amia ocellicauda Richardson, 1836 (Amiiformes: Amiidae) and its phylogenetic significance. Journal of Fish Biology 103(6):1300–1311. Published ahead of print, 18 August 2023. https://doi.org/10.1111/jfb.15528  Google Scholar

    1282.

    Pohl, M., F. Milvertz, A. Meyer, and M. Vences. 2015. Multigene phylogeny of cyprinodontiform fishes suggests continental radiations and a rogue taxon position of Pantanodon. Vertebrate Zoology 65(1):37–44. Google Scholar

    1283.

    Poll, M., and D. J. Stewart. 1975. Un Mochocidae et un Kneriidae nouveaux de la rivière Luongo (Zambia), affluent du bassin du Congo (Pisces). Revue de Zoologie Africaine 89:151–158. Google Scholar

    1284.

    Poly, W. J. 2004a. Family Aphredoderidae Bonaparte 1846: pirate perches. California Academy of Sciences Annotated Checklists of Fishes 24:1–5. Google Scholar

    1285.

    Poly, W. J. 2004b. Family Percopsidae Bonaparte 1846: trout perches and sand rollers. California Academy of Sciences Annotated Checklists of Fishes 23:1–5. Google Scholar

    1286.

    Poly, W. J., and G. S. Proudlove. 2004. Family Amblyopsidae Bonaparte 1846: cavefishes. California Academy of Sciences Annotated Checklists of Fishes 25:1–7. Google Scholar

    1287.

    Posner, M., and R. J. Lavenberg. 1999. Kasatkia seigeli: a new species of stichaeid (Perciformes: Stichaeidae) from California. Copeia 1999(4):1035–1040. Google Scholar

    1288.

    Postel, E. 1959. Liste commentée des poissons signalés dans l'Atlantique tropico-oriental nord, du Cap Spartel au Cap Roxo, suivie d'un bref aperçu sur leur répartition bathymétrique et géographique. Bulletin de la Société Scientifique de Bretagne 34:129–179. Google Scholar

    1289.

    Poulsen, J. Y. 2019. New observations and ontogenetic transformation of photogenic tissues in the tubeshoulder Sagamichthys schnakenbecki (Platytroctidae, Alepocephaliformes). Journal of Fish Biology 94(1):62–76. Google Scholar

    1290.

    Poulsen, J. Y., I. Byrkjedal, E. Willassen, D. Rees, H. Takeshima, T. P. Satoh, G. Shinohara, M. Nishida, and M. Miya. 2013. Mitogenomic sequences and evidence from unique gene rearrangements corroborate evolutionary relationships of Myctophiformes (Neoteleostei). BMC Evolutionary Biology 13:111.  https://doi.org/10.1186/1471-2148-13-111 Google Scholar

    1291.

    Poulsen, J. Y., M. J. Miller, T. Sado, R. Hanel, K. Tsukamoto, and M. Miya. 2018. Resolving deep-sea pelagic saccopharyngiform eel mysteries: identification of Neocyema and Monognathidae leptocephali and establishment of a new fish family “Neocyematidae” based on larvae, adults and mitogenomic gene orders. PLOS ONE 13(7):e0199982.  https://doi.org/10.1371/journal.pone.0199982 Google Scholar

    1292.

    Poulsen, J. Y., P. R. Møller, S. Lavoué, S. W. Knudsen, M. Nishida, and M. Miya. 2009. Higher and lower-level relationships of the deep-sea fish order Alepocephaliformes (Teleostei: Otocephala) inferred from whole mitogenome sequences. Biological Journal of the Linnean Society 98(4):923–936. Google Scholar

    1293.

    Poyato-Ariza, F. J. 1994. A New Early Cretaceous Gonorynchiform Fish (Teleostei: Ostariophysi) from Las Hoyas (Cuenca, Spain). Lawrence, KS: Museum of Natural History University of Kansas. 37 pp. (Occasional Papers of the Museum of Natural History University of Kansas 164.) Google Scholar

    1294.

    Poyato-Ariza, F. J. 1995. Ichthyemidion, a new genus for the elopiform fish “Anaethalion” vidali, from the Early Cretaceous of Spain: phylogenetic comments. Comptes rendus de l'Académie des sciences Série 2A: sciences de la terre et des planètes 320':133–139. Google Scholar

    1295.

    Poyato-Ariza, F. J. 1996a. A revision of Rubiesichthys gregalis WENZ 1984 (Ostariophysi, Gonorynchiformes), from the Early Cretaceous of Spain. In: G. Arratia and G. Viohl, eds. Mesozoic Fishes: Systematics and Paleoecology. Munich: Verlag Dr. Friedrich Pfeil. pp. 319–328. Google Scholar

    1296.

    Poyato-Ariza, F. J. 1996b. A Revision of the Ostariophysan Fish Family Chanidae, with Special Reference to the Mesozoic Forms. Munich: Verlag Dr. Friedrich Pfeil. 52 pp. (Palaeo Ichthyologica 6.) Google Scholar

    1297.

    Poyato-Ariza, F. J. 2015. Studies on pycnodont fishes (I): evaluation of their phylogenetic position among actinopterygians. Rivista Italiana di Paleontologia e Stratigrafia 121(3):329–343. Google Scholar

    1298.

    Poyato-Ariza, F. J., T. Grande, and R. Diogo. 2010. Gonorynchiform interrelationships: historic overview, analysis, and revised systematics of the group. In: T. Grande, F. J. Poyato-Ariza, and R. Diogo, eds. Gonorynchiformes and Ostariophysan Relationships: A Comprehensive Review. Enfield, NH: Science Publishers. pp. 227–337. Google Scholar

    1299.

    Prashad, B., and D. D. Mukerji. 1929. The fish of the Indawgyi Lake and the streams of the Myitkyina District (Upper Burma). Records of the Indian Museum 31(3):161–223. Google Scholar

    1300.

    Přikryl, T. 2014. A new species of the sleeper goby (Gobioidei, Eleotridae) from the České Středohoří Mountains (Czech Republic, Oligocene) and analysis of the validity of the family Pirskeniidae. Paläontologische Zeitschrift 88:187–196. Google Scholar

    1301.

    Přikryl, T., and G. Carnevale. 2017. An Oligocene toadfish (Teleostei, Percomorpha) from Moravia, Czech Republic: the earliest skeletal record for the order Batrachoidiformes. Bulletin of Geosciences 92(1):123–130. Google Scholar

    1302.

    Přikryl, T.. 2019. An Oligocene tubeshoulder (Teleostei, Alepocephaliformes) from the Central Paratethys (Czech Republic): the first skeletal record for the family Platytroctidae. Journal of Vertebrate Paleontology 39(6):e1719123.  https://doi.org/10.1080/02724634.2019.1719123 Google Scholar

    1303.

    Přikryl, T., I. Kania, and W. Krzemiński. 2016. Synopsis of fossil fish fauna from the Hermanowa locality (Rupelian; Central Paratethys; Poland): current state of knowledge. Swiss Journal of Geosciences 109:429–443. Google Scholar

    1304.

    Prokofiev, A. M. 2002a. A remarkable new genus of Carangidae from the Upper Paleocene of Turkmenistan (Osteichthyes: Perciformes). Zoosystematica Rossica 11(1):219–228. Google Scholar

    1305.

    Prokofiev, A. M. 2002b. A new genus of cutlassfish from the Upper Paleocene of Turkmenistan (Scombroidei: Trichiuroidea). Zoosystematica Rossica 11(1):229–233. Google Scholar

    1306.

    Prokofiev, A. M. 2006a. Fossil myctophid fishes (Myctophiformes: Myctophoidei) from Russia and adjacent regions. Journal of Ichthyology 46(Suppl. 1):S38–S83. Google Scholar

    1307.

    Prokofiev, A. M. 2006b. A new genus of cardinalfishes (Perciformes: Apogonidae) from the South China Sea, with a discussion of the relationships between the families Apogonidae and Kurtidae. Journal of Ichthyology 46(4):279–291. Google Scholar

    1308.

    Quast, J. C. 1965. Osteological characteristics and affinities of the hexagrammid fishes, with a synopsis. Proceedings of the California Academy of Sciences, Series 4, 31(21):563–600. Google Scholar

    1309.

    Quoy, J. R. C., and J. P. Gaimard. 1824. Description des poissons. In: L. C. D. De Freycinet, ed. Voyage autour du Monde, entrepris par ordre du roi,…exécuté sur les corvettes de S. M. L'Uranie et La Physicienne, pendant les années 1817, 1818, 1819 et 1820. Paris: Pillet Ainé. pp. 192–401. Google Scholar

    1310.

    Quoy, J. R. C.1825. Description des Poissons. In: L. C. D. De Freycinet, ed. Voyage autour du Monde, entrepris par ordre du roi,…exécuté sur les corvettes de S. M. L'Uranie et La Physicienne, pendant les années 1817, 1818, 1819 et 1820. Paris: Pillet Ainé. pp. 329–616. Google Scholar

    1311.

    Rabosky, D. L., J. Chang, P. O. Title, P. F. Cowman, L. Sallan, M. Friedman, K. Kaschner, C. Garilao, T. J. Near, ET AL. 2018. An inverse latitudinal gradient in speciation rate for marine fishes. Nature 559:392–395. Google Scholar

    1312.

    Rabosky, D. L.2019. Data from: An inverse latitudinal gradient in speciation rate for marine fishes [online dataset]. Dryad.  https://doi.org/10.5061/dryad.fc71cp4  Google Scholar

    1313.

    Rabosky, D. L., F. Santini, J. Eastman, S. A. Smith, B. Sidlauskas, J. Chang, and M. E. Alfaro. 2013. Rates of speciation and morphological evolution are correlated across the largest vertebrate radiation. Nature Communications 4:1958.  https://doi.org/10.1038/ncomms2958 Google Scholar

    1314.

    Radchenko, O. A. 2015. The system of the suborder Zoarcoidei (Pisces, Perciformes) as inferred from molecular genetic data. Russian Journal of Genetics 51(11):1096–1112. Google Scholar

    1315.

    Radchenko, O. A. 2016. Timeline of the evolution of eelpouts from the suborder Zoarcoidei (Perciformes) based on DNA variability. Journal of Ichthyology 56(4):556–568. Google Scholar

    1316.

    Radchenko, O. A. 2017. Molecular Systematics and Phylogeny of Zoarcoid Fishes. Moscow: GEOS. 383 pp. [In Russian.] Google Scholar

    1317.

    Radchenko, O. A., I. A. Chereshnev, and A. V. Petrovskaya. 2010. Relationships and position of the genus Neozoarces of the subfamily neozoarcinae in the system of the suborder Zoarcoidei (Pisces, Perciformes) by molecular-genetic data. Journal of Ichthyology 50(3):246–251. Google Scholar

    1318.

    Radchenko, O. A.. 2012. Position of neck banded blenny Leptostichaeus pumilus (Perciformes: Zoarcoidei) in the system of the suborder Zoarcoidei as inferred from molecular genetic data. Journal of Ichthyology 52(9):592–598. Google Scholar

    1319.

    Radchenko, O. A.. 2014. Genetic differentiation of species and taxonomic structure of the superfamily Stichaeoidea (Perciformes: Zoarcoidei). Russian Journal of Marine Biology 40(6):473–485. Google Scholar

    1320.

    Radchenko, O. A., I. A. Chereshnev, A. V. Petrovskaya, A. A. Balanov, and S. V. Turanov. 2014. Position of the genus Azygopterus (Stichaeidae, Perciformes) in the system of the suborder Zoarcoidei as inferred from sequence variation of mitochondrial and nuclear genes. Russian Journal of Genetics 50(3):280–287. Google Scholar

    1321.

    Radovčič, J. 1975. Some new upper Cretaceous teleosts from Yugoslavia with special reference to localities, geology and palaeoenvironment. Palaeontolgia Jugoslavica 17:7–55. Google Scholar

    1322.

    Rafinesque, C. S. 1810a. Caratteri di alcuni nuovi generi e nuove specie di animali e piante della Sicilia: con varie osservazioni sopra i medesim. Palermo: Sanfilippo. 70 pp. Google Scholar

    1323.

    Rafinesque, C. S. 1810b. Indice d'ittiologia siciliana: ossia, catalogo metodico dei nomi Latini, Italiani, e Siciliani dei pesci, che si rinvengono in Sicilia disposti secondo un metodo naturale e seguito da un appendice che contiene la descrizione de alcuni nuovi pesci Siciliani. Messina: Giovanni del Nobolo. 70 pp. Google Scholar

    1324.

    Rafinesque, C. S. 1815. Analyse de la nature, ou tableau d' l'univers et des corps organisés. Palerme: Aux dépens d' l'auteur. 224 pp. Google Scholar

    1325.

    Raju, S. N. 1974. Three new species of the genus Monognathus and the leptocephali of the order Saccopharyngiformes. Fishery Bulletin 72(2):547–562. Google Scholar

    1326.

    Randall, J. E., and B. C. Victor. 2015. Descriptions of thirty-four new species of the fish genus Pempheris (Perciformes: Pempheridae), with a key to the species of the western Indian Ocean. Journal of the Ocean Science Foundation 18:1–77. Google Scholar

    1327.

    Randall, Z. S., and L. M. Page. 2015. On the paraphyly of Homaloptera (Teleostei: Balitoridae) and description of a new genus of hillstream loaches from the Western Ghats of India. Zootaxa 3926(1):57–86.  https://doi.org/10.11646/zootaxa.3926.1.2 Google Scholar

    1328.

    Rasquin, P. 1958. Ovarian morphology and early embryology of the pediculate fishes Antennarius and Histrio. Bulletin of the American Museum of Natural History 114:327–372. Google Scholar

    1329.

    Reed, D. L., K. E. Carpenter, and M. J. Degravelle. 2002. Molecular systematics of the Jacks (Perciformes: Carangidae) based on mitochondrial cytochrome b sequences using parsimony, likelihood, and Bayesian approaches. Molecular Phylogenetics and Evolution 23(3):513–524. Google Scholar

    1330.

    Regan, C. T. 1903. On the classification of the fishes of the suborder Plectognathi; with notes and descriptions of new species from specimens in the British Museum Collection. Proceedings of the Zoological Society of London 1902, vol. 2, pp. 284–303. Google Scholar

    1331.

    Regan, C. T. 1904a. A monograph of the fishes of the family Loricariidae. Transactions of the Zoological Society of London 17(pt. 3):191–350. Google Scholar

    1332.

    Regan, C. T. 1904b. The phylogeny of Teleostomi. Annals and Magazine of Natural History, Series 7, 13(7):329–349. Google Scholar

    1333.

    Regan, C. T. 1907a. Descriptions of the teleostean fish Velifer hypselopterus, and of a new species of the genus Velifer. Proceedings of the Zoological Society of London 1907, pt. 3 (art. 4), pp. 633–634. Google Scholar

    1334.

    Regan, C. T. 1907b. On the anatomy, classification and systematic position of the Teleostean fishes of the sub-order Allotriognathi. Proceedings of the Zoological Society of London 1907, pt. 3 (art. 4), pp. 634–643. Google Scholar

    1335.

    Regan, C. T. 1908. The systematic position of Stylophorus cauda-tus. Annals and Magazine of Natural History, Series 8, 2(11):447–449. Google Scholar

    1336.

    Regan, C. T. 1909a. The classification of teleostean fishes. Annals and Magazine of Natural History, Series 8, 3(13):75–86. Google Scholar

    1337.

    Regan, C. T. 1909b. On the anatomy and classification of the scombroid fishes. Annals and Magazine of Natural History, Series 8, 3(13):66–75. Google Scholar

    1338.

    Regan, C. T. 1910a. The anatomy and classification of the teleostean fishes of the order Zeomorphi. Annals and Magazine of Natural History, Series 8, 6(35):481–484. Google Scholar

    1339.

    Regan, C. T. 1910b. The origin and evolution of the teleostean fishes of the order Heterosomata. Annals and Magazine of Natural History, Series 8, 6(35):484–496. Google Scholar

    1340.

    Regan, C. T. 1911a. The anatomy and classification of the teleostean fishes of the order Iniomi. Annals and Magazine of Natural History, Series 8, 7(37):120–133. Google Scholar

    1341.

    Regan, C. T. 1911b. The anatomy and classification of the teleostean fishes of the order Salmopercae. Annals and Magazine of Natural History, Series 8, 7(39):294–296. Google Scholar

    1342.

    Regan, C. T. 1911c. The anatomy and classification of the teleostean fishes of the orders Berycomorphi and Xenoberyces. Annals and Magazine of Natural History, Series 8, 7(37):1–9. Google Scholar

    1343.

    Regan, C. T. 1911d. The classification of the teleostean fishes of the order Ostariophysi.—1. Cyprinoidea. Annals and Magazine of Natural History, Series 8, 8(43):13–32. Google Scholar

    1344.

    Regan, C. T. 1911e. The classification of the teleostean fishes of the order Ostariophysi.—2. Siluroidea. Annals and Magazine of Natural History, Series 8, 8(47):553–577. Google Scholar

    1345.

    Regan, C. T. 1911f. The classification of the teleostean fishes of the order Synentognathi. Annals and Magazine of Natural History, Series 8, 7(40):327–335. Google Scholar

    1346.

    Regan, C. T. 1911g. The osteology and classification of the teleostean fishes of the order Microcyprini. Annals and Magazine of Natural History, Series 8, 7(40):320–327. Google Scholar

    1347.

    Regan, C. T. 1912a. The anatomy and classification of the teleostean fishes of the order Lyomeri. Annals and Magazine of Natural History, Series 8, 10(57):347–349. Google Scholar

    1348.

    Regan, C. T. 1912b. The classification of the blennioid fishes. Annals and Magazine of Natural History, Series 8, 10(57):265–280.n bvc Google Scholar

    1349.

    Regan, C. T. 1912c. The classification of the teleostean fishes of the order Pediculati. Annals and Magazine of Natural History, Series 8, 9(51):277–289. Google Scholar

    1350.

    Regan, C. T. 1912d. Description of two new eels from West Africa, belonging to a new genus and family. Annals and Magazine of Natural History, Series 8, 10(57): 323–324. Google Scholar

    1351.

    Regan, C. T. 1912e. The osteology and classification of the teleostean fishes of the order Apodes. Annals and Magazine of Natural History, Series 8, 10(58):377–387. Google Scholar

    1352.

    Regan, C. T. 1912f. A revision of the Pæciliid fishes of the genera Rivulus, Pterolebias, and Cynolebias. Annals and Magazine of Natural History, Series 8, 10(59):494–508. Google Scholar

    1353.

    Regan, C. T. 1913a. The Antarctic fishes of the Scottish National Antarctic Expedition. Transactions of the Royal Society of Edinburgh 49:229–292. Google Scholar

    1354.

    Regan, C. T. 1913b. The classification of the percoid fishes. Annals and Magazine of Natural History, Series 8, 12(67):111–145. Google Scholar

    1355.

    Regan, C. T. 1913c. The osteology and classification of the teleostean fishes of the order Scleroparei. Annals and Magazine of Natural History, Series 8, 11(62):169–184. Google Scholar

    1356.

    Regan, C. T. 1914a. Description of a new cyprinodont fish of the genus Mollienisia from Yucatan. Annals and Magazine of Natural History, Series 8, 13(75):338. Google Scholar

    1357.

    Regan, C. T. 1914b. Fishes. British Antarctic (“Terra Nova”) expedition, 1910. Natural history report. Zoology 1(1):1–54. Google Scholar

    1358.

    Regan, C. T. 1923a. The classification of the stomiatoid fishes. Annals and Magazine of Natural History, Series 9, 11(65):612–614. Google Scholar

    1359.

    Regan, C. T. 1923b. The skeleton of Lepidosteus, with remards on the origin and evolution of the lower neopterygian fishes. Proceedings of the Zoological Society of London 1923 (art.24), pp. 445–461. Google Scholar

    1360.

    Regan, C. T. 1924. The morphology of a rare oceanic fish, Stylephorus chordatus, Shaw; based on specimens collected in the Atlantic by the “Dana” Expeditions, 1920–1922. Proceedings of the Royal Society of London B, Biological Sciences 96(674):193–207. Google Scholar

    1361.

    Regan, C. T. 1925. Dwarfed males parasitic on the females in oceanic angler-fishes (Pediculati Ceratioidea). Proceedings of the Royal Society of London B, Biological Sciences 97(684):386–400. Google Scholar

    1362.

    Regan, C. T. 1926. The Pediculate Fishes of the Suborder Ceratioidea. Copenhagen: Gyldendalske Boghandel. 45 pp. (Danish “Dana”—Expeditions in the North Atlantic and the Gulf of Panama [1920–1922] Oceanographical Reports 2.) Google Scholar

    1363.

    Regan, C. T. 1929. Fishes. In: Encyclopaedia Britannica. 14th ed. Volume 9. London: Encyclopaedia Britannica Company. pp. 305–329. Google Scholar

    1364.

    Reichenbacher, B., T. Přikryl, A. F. Cerwenka, P. Keith, C. Gierl, and M. Dohrmann. 2020. Freshwater gobies 30 million years ago: new insights into character evolution and phylogenetic relationships of †Pirskeniidae (Gobioidei, Teleostei). PLOS ONE 15(8):e0237366.  https://doi.org/10.1371/journal.pone.0237366 Google Scholar

    1365.

    Reinhardt, J. C. H. 1837. Ichthyologiske bidrag til den Grönlandske fauna. Indledning, indeholdende tillaeg og forandringer I den fabriciske fortegnelse paa Grönlandske hvirveldyr. Det Kongelige Danske Videnskabernes Selskabs Naturvidenskabelige og Mathematiske Afhandlinger 7:83–196. Google Scholar

    1366.

    Reis, R. E. 1998. Systematics, biogeography, and the fossil record of the Callichthyidae: a review of the available data. In: L. R. Malabarba, R. E. Reis, R. P. Vari, Z. M. Lucena, and C. A. S. Lucena, eds. Phylogeny and Classification of Neotropical Fishes. Porto Alegre, Brazil: EDIPUCRS. pp. 351–362. Google Scholar

    1367.

    Reist, J. D. 1987. Comparative morphometry and phenetics of the genera of esocoid Fishes (Salmoniformes). Zoological Journal of the Linnean Society 89(3):275–294. Google Scholar

    1368.

    Ren, Y., and G.-H. XU. 2021. A new species of Pteronisculus from the Middle Triassic (Anisian) of Luoping, Yunnan, China, and phylogenetic relationships of early actinopterygian fishes. Vertebrata PalAsiatica 59(3):169–199. Google Scholar

    1369.

    Reznick, D. N., A. I. Furness, R. W. Meredith, and M. S. Springer. 2017. The origin and biogeographic diversification of fishes in the family Poeciliidae. PLOS ONE 12(3):e0172546.  https://doi.org/10.1371/journal.pone.0172546 Google Scholar

    1370.

    Ribeiro, A. C., F. J. Poyato-Ariza, F. A. Bockmann, and M. R. De Carvalho. 2018. Phylogenetic relationships of Chanidae (Teleostei: Gonorynchiformes) as impacted by Dastilbe moraesi, from the Sanfranciscana basin, Early CretaceousofBrazil.NeotropicalIchthyology16:e180059. https://doi.org/10.1590/1982-0224-20180059 Google Scholar

    1371.

    Ribeiro, E., A. M. Davis, R. A. Rivero-Vega, G. Ortí, and R. Betancur-R. 2018. Post-Cretaceous bursts of evolution along the benthic-pelagic axis in marine fishes. Proceedings of the Royal Society B, Biological Sciences 285(1893):20182010. Google Scholar

    1372.

    Rice, A. N., and A. H. Bass. 2009. Novel vocal repertoire and paired swimbladders of the three-spined toadfish, Batrachomoeus trispinosus: insights into the diversity of the Batrachoididae. Journal of Experimental Biology 212(pt. 9):1377–1391. Google Scholar

    1373.

    Richardson, J. 1843. Contributions to the ichthyology of Australia (continued). Annals and Magazine of Natural History, Series 1, 11 (67–68, 71–72):22–28, 169–182, 352–359, 422–428. Google Scholar

    1374.

    Richardson, J. 1844. Ichthyology of the voyage of H. M. S. Erebus and Terror, under the command of Captain Sir James Clark Ross, R. N., F. R. S. In: J. Richardson and J. E. Gray, eds. The Zoology of the Voyage of H. M. S. Erebus and Terror, under the Command of Captain Sir J. C. Ross, R. N., F. R. S., during the Years 1839 to 1843, Volume 2. London: Longman, Brown, Green, and Longmans. pp. 1–16. Google Scholar

    1375.

    Richardson, J. 1845. Ichthyology of the voyage of H. M. S. Erebus and Terror, under the command of Captain Sir James Clark Ross, R. N., F. R. S. In: J. Richardson and J. E. Gray, eds. The Zoology of the Voyage of H. M. S. Erebus and Terror, under the Command of Captain Sir J. C. Ross, R. N., F. R. S., during the Years 1839 to 1843, Volume 2. London: Longman, Brown, Green, and Longmans. pp. 17–52. Google Scholar

    1376.

    Richardson, J. 1846. Ichthyology of the voyage of H. M. S. Erebus and Terror, under the command of Captain Sir James Clark Ross, R. N., F. R. S. In: J. Richardson and J. E. Gray, eds. The Zoology of the Voyage of H. M. S. Erebus and Terror, under the Command of Captain Sir J. C. Ross, R. N., F. R. S., during the Years 1839 to 1843, Volume 2. London: Longman, Brown, Green, and Longmans. pp. 53–74. Google Scholar

    1377.

    Ridewood, W. G. 1904. On the cranial osteology of the fishes of the families Mormyridae, Notopteridae and Hyodontidae. Zoological Journal of the Linnean Society 29(190):188–217. Google Scholar

    1378.

    Ridewood, W. G. 1905. On the cranial osteology of the fishes of the families Osteoglossidae, Pantodontidae, and Phractolaemidae. Zoological Journal of the Linnean Society 29(191):252–282. Google Scholar

    1379.

    Rincon-Sandoval, M., E. Duarte-Ribeiro, A. M. Davis, A. Santaquiteria, L. C. Hughes, C. C. Baldwin, L. Soto-Torres, A. Acero P, H. J. Walker, K. E. Carpenter, ET AL. 2020. Evolutionary determinism and convergence associated with water-column transitions in marine fishes. Proceedings of the National Academy of Sciences 117(52):33396–33403. Google Scholar

    1380.

    Risso, A. 1810. Ichthyologie de Nice, ou Histoire Naturelle des Poissons du Département des Alpes Maritimes. Paris: F. Schoell. 388 pp. Google Scholar

    1381.

    Risso, A. 1820. Mémoire sur un nouveau genre de poisson nommé Alépocéphale vivant dans les grandes profondeurs de la mer de Nice. Memorie della Reale Accademia delle Scienze di Torino 25:270–272. Google Scholar

    1382.

    Rivera-Rivera, C. J., and J. I. Montoya-Burgos. 2017. Trunk dental tissue evolved independently from underlying dermal bony plates but is associated with surface bones in living odontode-bearing catfish. Proceedings of the Royal Society B, Biological Sciences 284(1865):20171831. Google Scholar

    1383.

    Rivera-Rivera, C. J.. 2018. Back to the roots: reducing evolutionary rate heterogeneity among sequences gives support for the early morphological hypothesis of the root of Siluriformes (Teleostei: Ostariophysi). Molecular Phylogenetics and Evolution 127:272–279. Google Scholar

    1384.

    Roa-Varón, A., R. B. Dikow, G. Carnevale, L. Tornabene, C. C. Baldwin, C. Li, and E. J. Hilton. 2021. Confronting sources of systematic error to resolve historically contentious relationships: a case study using gadiform fishes (Teleostei, Paracanthopterygii, Gadiformes). Systematic Biology 70(4):739–755. Google Scholar

    1385.

    Roa-Varón, A., and G. Ortí. 2009. Phylogenetic relationships among families of Gadiformes (Teleostei, Paracanthopterygii) based on nuclear and mitochondrial data. Molecular Phylogenetics and Evolution 52(3):688–704. Google Scholar

    1386.

    Roberts, C. D. 1993. Comparative morphology of spined scales and their phylogenetic significance in the Teleostei. Bulletin of Marine Science 52:60–113. Google Scholar

    1387.

    Roberts, T. R. 1973. Interrelationships of ostariophysans. In: P. H. Greenwood, R. S. Miles, and C. Patterson, eds. Interrelationships of Fishes. London: Academic Press. pp. 373–395. Google Scholar

    1388.

    Roberts, T. R. 2012. Systematics, biology,and distribution of the species of the oceanic oarfish genus Regalecus: Teleostei, Lampridiformes, Regalecidae. Mémoires du Muséum Nationa' d'Histoire Naturelle 202:1–268. Google Scholar

    1389.

    Robins, C. H., and C. R. Robins. 1989. Family Synaphobranchidae. In: E. B. Böhlke, ed.. Orders Anguilliofrmes and Saccopharyngiformes, Volume 1. New Haven, CT: Sears Foundation for Marine Research, Yale University. pp. 207–253. (Memoir 1, Fishes of the Western North Atlantic 9.) Google Scholar

    1390.

    Robins, C. R. 1989. The phylogenetic relationships of anguilliform fishes. In: E. B. Böhlke, ed. Orders Anguilliofrmes and Saccopharyngiformes, Volume 1. New Haven, CT: Sears Foundation for Marine Research, Yale University. pp. 9–23. (Memoir 1, Fishes of the Western North Atlantic 9.) Google Scholar

    1391.

    Rodiles-Hernández, R., D. A. Hendrickson, J. G. Lundberg, and J. M. Humphries. 2005. Lacantunia enigmatica (Teleostei: Siluriformes): a new and phylogenetically puzzling freshwater fish from Mesoamerica. Zootaxa 1000(1):1–24.  https://doi.org/10.11646/zootaxa.1000.1.1 Google Scholar

    1392.

    Roe, L. J. 1991. Phylogenetic and ecological significance of Channidae (Osteichthyes, Teleostei) from the early Eocene Kuldana Formation of Kohat, Pakistan. Contributions from the Museum of Paleontology University of Michigan 28(5):93–100. Google Scholar

    1393.

    Rosas Puchuri, U. F. 2021. What is Protacanthopterygii? A phylogenomic perspective of early-branching lineages of euteleost fishes [master's thesis]. George Washington University. 101 pp. Google Scholar

    1394.

    Rosen, D. E. 1962. Comments on the relationships of the North American cave fishes of the family Amblyopsidae. American Museum Novitates 2109:1–35. Google Scholar

    1395.

    Rosen, D. E. 1964. The relationships and taxonomic position of the halfbeaks, killifishes, silversides and their relatives. Bulletin of the American Museum of Natural History 127:217–268. Google Scholar

    1396.

    Rosen, D. E. 1973. Interrelationships of higher euteleostean fishes. In: P. H. Greenwood, R. S. Miles, and C. Patterson, eds. Interrelationships of Fishes. London: Academic Press. pp. 397–513. Google Scholar

    1397.

    Rosen, D. E. 1974. Phylogeny and zoogeography of salmoniform fishes and relationships of Lepidogalaxias salamandroides. Bulletin of the American Museum of Natural History 153:269–325. Google Scholar

    1398.

    Rosen, D. E. 1982. Teleostean interrelationships, morphological function and evolutionary inference. American Zoologist 22(2):261–273. Google Scholar

    1399.

    Rosen, D. E. 1984. Zeiforms as primitive plectognath fishes. American Museum Novitates 2782:1–45. Google Scholar

    1400.

    Rosen, D. E. 1985. An essay on euteleostean classification. American Museum Novitates 2827:1–57. Google Scholar

    1401.

    Rosen, D. E., P. L. Forey, B. G. Gardiner, and C. Patterson. 1981. Lungfishes, tetrapods, paleontology, and plesiomorphy. Bulletin of the American Museum of Natural History 167:163–275. Google Scholar

    1402.

    Rosen, D. E., and P. H. Greenwood. 1970. Origin of the Weberian apparatus and the relationships of the ostariophysan and gonorynchiform fishes. American Museum Novitates 2428:1–25. Google Scholar

    1403.

    Rosen, D. E.. 1976. A fourth Neotropical species of synbranchid eel and the phylogeny and systematics of synbranchiform fishes. Bulletin of the American Museum of Natural History 157:1–69. Google Scholar

    1404.

    Rosen, D. E., and L. R. Parenti. 1981. Relationships of Oryzias and the groups of atherinomorph fishes. American Museum Novitates 2719:1–25. Google Scholar

    1405.

    Rosen, D. E., and C. Patterson. 1969. The structure and relationships of the paracanthopterygian fishes. Bulletin of the American Museum of Natural History 141:357–474. Google Scholar

    1406.

    Rosen, D. E.. 1990. On Müller's and Cuvier's concepts of pharyngognath and labyrinth fishes and the classification of percomorph fishes, with an atlas of percomorph dorsal gill arches. American Museum Novitates 2983:1–57. Google Scholar

    1407.

    Roth, O., M. H. Solbakken, O. K. Tørresen, T. Bayer, M. Matschiner, H. T. Baalsrud, S. N. K. Hoff, M. S. O. Brieuc, D. Haase, R. Hanel, ET AL. 2020. Evolution of male pregnancy associated with remodeling of canonical vertebrate immunity in seahorses and pipefishes. Proceedings of the National Academy of Sciences 117(17):9431–9439. Google Scholar

    1408.

    Rüber, L., R. Britz, S. O. Kullander, and R. Zardoya. 2004. Evolutionary and biogeographic patterns of the Badidae (Teleostei: Perciformes) inferred from mitochondrial and nuclear DNA sequence data. Molecular Phylogenetics and Evolution 32(3):1010–1022. Google Scholar

    1409.

    Rüber, L., R. Britz, and R. Zardoya. 2006. Molecular phylogenetics and evolutionary diversification of labyrinth fishes (Perciformes: Anabantoidei). Systematic Biology 55(3):374–397. Google Scholar

    1410.

    Rüber, L., M. Kottelat, H. H. Tan, P. K. L. Ng, and R. Britz. 2007. Evolution of miniaturization and the phylogenetic position of Paedocypris, comprising the wo'ld's smallest vertebrate. BMC Evolutionary Biology 7:38.  https://doi.org/10.1186/1471-2148-7-38 Google Scholar

    1411.

    Rüber, L., H. H. Tan, and R. Britz. 2020. Snakehead (Teleostei: Channidae) diversity and the Eastern Himalaya biodiversity hotspot. Journal of Zoological Systematics and Evolutionary Research 58(1):356–386. Google Scholar

    1412.

    Rüber, L., J. L. Van Tassell, and R. Zardoya. 2003. Rapid speciation and ecological divergence in the American seven-spined gobies (Gobiidae, Gobiosomatini) inferred from a molecular phylogeny. Evolution 57(7):1584–1598. Google Scholar

    1413.

    Russell, A. 1794. Natural History of Aleppo. 2nd ed. London: Millar. 430 pp. Google Scholar

    1414.

    Russell, B. C. 1990. Nemipterid Fishes of the World (Threadfin Breams, Whiptail Breams, Monocle Breams, Dwarf Monocle Breams and Coral Breams). Vol. 12 of FAO Species Catalogue. Rome: Food and Agriculture Organization of the United Nations. 303 pp. (FAO Fisheries Synopsis 125.) Google Scholar

    1415.

    Rutenko, O. A., S. V. Turanov, and Y. PH. Kartavtsev. 2019. Complete mitochondrial genome of ocellated blenny, Opisthocentrus ocellatus (Tilesius, 1811) (Zoarcales: Opisthocentidae). Mitochondrial DNA Part B 4(1):1553–1555. Google Scholar

    1416.

    Sabaj, M. H., M. Arce H, D. Donahue, A. Cramer, and L. M. Sousa. 2022. Synbranchus of the Middle to Lower Xingu Basin, Brazil, with the description of a new rheophilic species, S. royal (Synbranchiformes: Synbranchidae). Proceedings of the Academy of Natural Sciences of Philadelphia 166(1):1–24. Google Scholar

    1417.

    Saeed, B., W. Ivantsoff, and L. E. L. M. Crowley. 1994. Systematic relationships of atheriniform families within Division I of the Series Atherinomorpha (Acanthopterygii) with relevant historical perspectives. Journal of Ichthyology 34:27–72. Google Scholar

    1418.

    Sagemehl, M. 1885. Beiträge zur vergleichenden anatomie der fische. III. Das Cranium der Characiniden nebst allgemeinen Bemerkungen über die mit einem'Weber'schen Apparat versehenen Physostomenfamilien. Morphologisches Jahrbuch 10:1–119. Google Scholar

    1419.

    Saitoh, K., M. Miya, J. G. Inoue, N. B. Ishiguro, and M. Nishida. 2003. Mitochondrial genomics of ostrariophysan fishes: perspectives on phylogeny and biogeography. Journal of Molecular Evolution 56(4):464–472. Google Scholar

    1420.

    Saitoh, K., T. Sado, M. H. Doosey, H. L. Bart, J. G. Inoue, M. Nishida, R. L. Mayden, and M. Miya. 2011. Evidence from mitochondrial genomics supports the lower Mesozoic of South Asia as the time and place of basal divergence of cypriniform fishes (Actinopterygii: Ostariophysi). Zoological Journal of the Linnean Society 161(3):633–662. Google Scholar

    1421.

    Saitoh, K., T. Sado, R. L. Mayden, N. Hanzawa, K. Nakamura, M. Nishida, and M. Miya. 2006. Mitogenomic evolution and interrelationships of the Cypriniformes (Actinopterygii: Ostariophysi): the first evidence toward resolution of higher-level relationships of the'world's largest freshwater fish clade based on 59 whole mitogenome sequences. Journal of Molecular Evolution 63:826–841. Google Scholar

    1422.

    Sakamoto, K., T. Uyeno, and N. Micklich. 2004. Oligopleuronectes germanicus gen. et sp. nov., an Oligocene pleuronectid flatfish from Frauenweiler, S-Germany. Bulletin of the National Science Museum, Series C, Geology and Paleontology 30:89–94. Google Scholar

    1423.

    Sanciangco, M. D., K. E. Carpenter, and R. Betancur-R. 2016. Phylogenetic placement of enigmatic percomorph families (Teleostei: Percomorphaceae). Molecular Phylogenetics and Evolution 94(Part B):565–576. Google Scholar

    1424.

    Sanford, C. P. J. 1990. The phylogenetic relationships of salmonoid fishes. Bulletin of the British Museum (Natural History), Zoology 56:145–153. Google Scholar

    1425.

    Santaquiteria, A., A. C. Siqueira, E. Duarte-Ribeiro, G. Carnevale, W. T. White, J. J. Pogonoski, C. C. Baldwin, G. Ortí, D. Arcila, and R. Betancur-R. 2021. Phylogenomics and historical biogeography of seahorses, dragonets, goatfishes, and allies (Teleostei: Syngnatharia): assessing factors driving uncertainty in biogeographic inferences. Systematic Biology 70(6):1145–1162. Google Scholar

    1426.

    Santini, F., and G. Carnevale. 2015. First multilocus and densely sampled timetree of trevallies, pompanos and allies (Carangoidei, Percomorpha) suggests a Cretaceous origin and Eocene radiation of a major clade of piscivores. Molecular Phylogenetics and Evolution 83:33–39. Google Scholar

    1427.

    Santini, F., L. J. Harmon, G. Carnevale, and M. E. Alfaro. 2009. Did genome duplication drive the origin of teleosts? A comparative study of diversification in ray-finned fishes. BMC Evolutionary Biology 9:194.  https://doi.org/10.1186/1471-2148-9-194 Google Scholar

    1428.

    Santini, F., X. H. Kong, L. Sorenson, G. Carnevale, R. S. Mehta, and M. E. Alfaro. 2013. A multi-locus molecular timescale for the origin and diversification of eels (order: Anguilliformes). Molecular Phylogenetics and Evolution 69(3):884–894. Google Scholar

    1429.

    Santini, F., M. T. T. Nguyen, L. Sorenson, T. B. Waltzek, J. W. L. Alfaro, J. M. Eastman, and M. E. Alfaro. 2013. Do habitat shifts drive diversification in teleost fishes? An example from the pufferfishes (Tetraodontidae). Journal of Evolutionary Biology 26(5):1003–1018. Google Scholar

    1430.

    Santini, F., L. Sorenson, and M. E. Alfaro. 2013. A new phylogeny of tetraodontiform fishes (Tetraodontiformes, Acanthomorpha) based on 22 loci. Molecular Phylogenetics and Evolution 69(1):177–187. Google Scholar

    1431.

    Santini, F., and J. C. Tyler. 2003. A phylogeny of the families of fossil and extant tetraodontiform fishes (Acanthomorpha, Tetraodontiformes), Upper Cretaceous to Recent. Zoological Journal of the Linnean Society 139(4):565–617. Google Scholar

    1432.

    Santini, F.. 2004. The importance of even highly incomplete fossil taxa in reconstructing the phylogenetic relationships of the Tetraodontiformes (Acanthomorpha: Pisces). Integrative and Comparative Biology 44(5):349–357. Google Scholar

    1433.

    Santini, F., J. C. Tyler, A. F. Bannikov, and D. S. Baciu. 2006. A phylogeny of extant and fossil buckler dory fishes, family Zeidae (Zeiformes, Acanthomorpha). Cybium 30(2):99–107. Google Scholar

    1434.

    Sato, M. 2022. Morphological diversity of the lateral line system in Teleostei. In: Y. Kai, H. Motomura, and K. Matsuura, eds. Fish Diversity of Japan: Evolution, Zoogeography, and Conservation. Singapore: Springer Nature Singapore. pp. 283–310. Google Scholar

    1435.

    Sato, T., and T. Nakabo. 2002. Paraulopidae and Paraulopus, a new family and genus of aulopiform fishes with revised relationships within the order. Ichthyological Research 49:25–46. Google Scholar

    1436.

    Satoh, T. P. 2018. Complete mitochondrial genome sequence of Glaucosoma buergeri (Pempheriformes: Glaucosomatidae) with implications based on the phylogenetic position. Mitochondrial DNA Part B 3(1):107–109. Google Scholar

    1437.

    Satoh, T. P., and E. Katayama. 2022. Complete mitochondrial genomes of two sand diver species (Perciformes, Trichonotidae): novel gene orders and phylogenetic position within Gobiiformes. Mitochondrial DNA Part B 7(1):12–14. Google Scholar

    1438.

    Sawada, Y. 1982. Phylogeny and zoogeography of the superfamily Cobitoidea (Cyprinoidei, Cypriniformes). Memoirs of the Faculty of Fisheries, Hokkaido University 28(2):65–223. Google Scholar

    1439.

    Schaefer, S. A. 1990. Anatomy and relationships of the scoloplacid catfishes. Proceedings of the Academy of Natural Sciences of Philadelphia 142:167–210. Google Scholar

    1440.

    Schaefer, S. A. 2006. Morphological Evolution, Aptations, Homoplasies, Constraints and Evolutionary Trends: Catfishes as a Case Study on General Phylogeny and Macroevolution. By Rui Diogo [book review]. Quarterly Review of Biology 81(3):276–278. Google Scholar

    1441.

    Schaefer, S. A., and G. V. Lauder. 1986. Historical transformation of functional design: evolutionary morphology of feeding mechanisms in loricarioid catfishes. Systematic Zoology 35(4):489–508. Google Scholar

    1442.

    Schaeffer, B., and C. Patterson. 1984. Jurassic fishes from the western United States, with comments on Jurassic fish distribution. American Museum Novitates 2796:1–86. Google Scholar

    1443.

    Schedel, F. D. B., A. Chakona, B. L. Sidlauskas, M. O. Popoola, N. Usimesa Wingi, D. Neumann, E. J. W. M. N. Vreven, and U. K. Schliewen. 2022. New phylogenetic insights into the African catfish families Mochokidae and Austroglanididae. Journal of Fish Biology 100(5):1171–1186. Google Scholar

    1444.

    Schlesinger, G. 1909. Zur phylogenie und ethologie der scombresociden. Verhandlungen der Kaiserlich-Königlichen Zoologisch-Botanischen Gesellschaft in Wien 59:302–339. Google Scholar

    1445.

    Schomburgk, R. H. 1848. The History of Barbados: Comprising a Geographical and Statistical Description of the Island; a Sketch of the Historical Events Since the Settlement; and an Account of Its Geology and Natural Productions. London: Longman, Brown, Green and Longmans. 722 pp. Google Scholar

    1446.

    Schrøder, A. E., and G. Carnevale. 2023. The argentiniform Surlykus longigracilis gen. et sp. nov., the most abundant fish from the Eocene Fur Formation of Denmark. Bulletin of the Geological Society of Denmark 72:1–18. Google Scholar

    1447.

    Schrøder, A. E., J. A. Rasmussen, P. R. Møller, and G. Carnevale. 2022. A new beardfish (Teleostei, Polymixiiformes) from the Eocene Fur Formation, Denmark. Journal of Vertebrate Paleontology 42(2):e2142914.  https://doi.org/10.1080/02724634.2022.2142914 Google Scholar

    1448.

    Schultz, L. P. 1948. A revision of six subfamilies of Atherine fishes, with descriptions of new genera and species. Proceedings of the United States National Museum 98(3220):1–48. Google Scholar

    1449.

    Schultze, H.-P., and G. Arratia. 1988. Reevaluation of the caudal skeleton of some actinopterygian fishes. II. Hiodon, Elops, and Albula. Journal of Morphology 195(3):257–303. Google Scholar

    1450.

    Schultze, H.-P., and E. O. Wiley. 1984. The neopterygian Amia as a living fossil. In: N. Eldredge and S. M. Stanley, eds. Living Fossils. New York: Springer. pp. 153–159. Google Scholar

    1451.

    Schwarzhans, W. 2003. Fish Otoliths from the Paleocene of Denmark. Copenhagen: Danmarks og Grønlands geologiske undersøgelse. 94 pp. (Geologic Survey of Denmark and Greenland Bulletin 2.) Google Scholar

    1452.

    Schwarzhans, W. 2007. The otoliths from the middle Eocene of Osteroden near Bramsche, north-western Germany. Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 244:299–369. Google Scholar

    1453.

    Schwarzhans, W. 2010. Otolithen aus den Gerhartsreiter Schichten (Oberkreide: Maastricht) des Gerhartsreiter Grabens (Oberbayern). Munich: Verlag Dr. Friedrich Pfeil. 100 pp. (Palaeo Ichthyologica 4.) Google Scholar

    1454.

    Schwarzhans, W. 2019. Reconstruction of the fossil marine bony fish fauna (Teleostei) from the Eocene to Pleistocene of New Zealand by means of otoliths. With studies of Recent congroid, morid and trachinoid otoliths. Memorie della Società Italianan de Scienze Naturali e del Museo Civico di Storia Naturale di Milano 46:1–326. Google Scholar

    1455.

    Schwarzhans, W., and A. Bratishko. 2011. The otoliths from the middle Paleocene of Luzanivka (Cherkasy district, Ukraine). Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 261:83–110. Google Scholar

    1456.

    Schwarzhans, W., and G. Carnevale. 2021. The rise to dominance of lanternfishes (Teleostei: Myctophidae) in the oceanic ecosystems: a paleontological perspective. Paleobiology 47(3):446–463. Google Scholar

    1457.

    Schwarzhans, W., and J. W. M. Jagt. 2021. Silicified otoliths from the Maastrichtian type area (Netherlands, Belgium) document early gadiform and perciform fishes during the Late Cretaceous, prior to the K/Pg boundary extinction event. Cretaceous Research 127:104921. https://doi.org/10.1016/j.cretres.2021.104921 Google Scholar

    1458.

    Schwarzhans, W., and G. Stringer. 2020. Fish otoliths from the Late Maastrichtian Kemp Clay (Texas, USA) and the Early Dannian Clayton Formation (Arkansas, USA) and an assessment of extinction and survival of teleost lineages across the K-Pg boundary based on otoliths. Rivista Italiana di Paleontologia e Stratigrafia 126(2):395–446. Google Scholar

    1459.

    Semple, G. P. 1985. Reproductive behaviour and development of the Glassperchlet, Ambassis agrammus Gunther (Pisces: Ambassidae), from the Alligator rivers system, northern territory. Australian Journal of Marine and Freshwater Research 36(6):797–805. Google Scholar

    1460.

    Setiamarga, D. H. E., M. Miya, Y. Yamanoue, K. Mabuchi, T. P. Satoh, J. G. Inoue, and M. Nishida. 2008. Interrelationships of Atherinomorpha (medakas, flyingfishes, killifishes, silversides, and their relatives): the first evidence based on whole mitogenome sequences. Molecular Phylogenetics and Evolution 49(2):598–605. Google Scholar

    1461.

    Sferco, E., A. López-Arbarello, and A. M. Báez. 2015. Phylogenetic relationships of †Luisiella feruglioi (Bordas) and the recognition of a new clade of freshwater teleosts from the Jurassic of Gondwana. BMC Evolutionary Biology 15:1–15.  https://doi.org/10.1186/s12862-015-0551-6 Google Scholar

    1462.

    Shaw, G. 1791. Description of the Stylephorus chordatus, a new fish. Transactions of the Linnean Society of London 1:90–92. Google Scholar

    1463.

    Shaw, G., and F. P. Nodder. 1799. The Naturalist's Miscellany, or Coloured Figures of Natural Objects; Drawn and Described from Nature, Volume 10. London: J. Cooper. 297 pp. Google Scholar

    1464.

    Shedko, S. V. 2022. Molecular dating of phylogeny of sturgeons (Acipenseridae) based on total evidence analysis. Russian Journal of Genetics 58(6):718–729. Google Scholar

    1465.

    Shedlock, A. M., T. W. Pietsch, M. G. Haygood, P. Bentzen, and M. Hasegawa. 2004. Molecular systematics and life history evolution of anglerfishes (Teleostei: Lophiiformes): evidence from mitochondrial DNA. Steenstrupia 28(2):129–144. Google Scholar

    1466.

    Shen, C., and G. Arratia. 2021. Re-description of the sexually dimorphic peltopleuriform fish Wushaichthys exquisitus (Middle Triassic, China): taxonomic implications and phylogenetic relationships. Journal of Systematic Palaeontology 19(19):1317–1342. Google Scholar

    1467.

    Shen, Y., N. Yang, Z. Liu, Q. Chen, and Y. Li. 2020. Phylogenetic perspective on the relationships and evolutionary history of the Acipenseriformes. Genomics 112(5):3511–3517. Google Scholar

    1468.

    Shi, W., S. Chen, X. Kong, L. Si, L. Gong, Y. Zhang, and H. Yu. 2018. Flatfish monophyly refereed by the relationship of Psettodes in Carangimorphariae. BMC Genomics 19:400.  https://doi.org/10.1186/s12864-018-4788-5 Google Scholar

    1469.

    Shinohara, G. 1994. Comparative morphology and phylogeny of the suborder Hexagrammoidei and related taxa (Pisces: Scorpaeniformes). Memoirs of the Faculty of Fisheries, Hokkaido University 41:1–97. Google Scholar

    1470.

    Siebert, D. J. 1987. Interrelationships among families of the order Cypriniformes (Teleostei) [dissertation]. City University of New York. 380 pp. Google Scholar

    1471.

    Siebert, D. J. 1990. Papers on the Systematics of Gadiform Fishes. D.M. Cohen(ed.)[bookreview].Copeia1990(3):889–893. Google Scholar

    1472.

    Silva, G. S. C., B. F. Melo, F. F. Roxo, L. E. Ochoa, O. A. Shibatta, M. H. Sabaj, and C. Oliveira. 2021. Phylogenomics of the bumblebee catfishes (Siluriformes: Pseudopimelodidae) using ultraconserved elements. Journal of Zoological Systematics and Evolutionary Research 59(8):1662–1672. Google Scholar

    1473.

    Silva, H. M. A., and V. Gallo. 2011. Taxonomic review and phylogenetic analysis of Enchodontoidei. Anais da Academia Brasileira de Ciências 83:483–511. Google Scholar

    1474.

    Silva Santos, R. 1985. Laeliichthys ancestralis, novo gênero e espécie de Osteoglossiformes do Aptiano da Formação Areado, Estado de Minas Gerais, Brasil. Série Geologia. Seção Paleontologia e Estratigrafia 27:151–167. Google Scholar

    1475.

    Simion, P., F. Delsuc, and H. Philippe. 2020. To what extent current limits of phylogenomics can be overcome? In: C. Scornavacca, F. Delsuc, and N. Galtier, eds. Phylogenetics in the Genomic Era. HAL Open Science [online archive].. 35 pp.  https://hal.science/hal-02535366  Google Scholar

    1476.

    Simons, A. M., and N. J. Gidmark. 2010. Systematics and phylogenetic relationships of Cypriniformes. In: T. Grande, F. J. Poyato-Ariza, and R. Diogo, eds. Gonorynchiformes and Ostariophysan Relationships: A Comprehensive Review. Enfield, NH: Science Publishers. pp. 409–440. Google Scholar

    1477.

    Siqueira, A. C., D. R. Bellwood, and P. F. Cowman. 2019. Historical biogeography of herbivorous coral reef fishes: the formation of an Atlantic fauna. Journal of Biogeography 46(7):1611–1624. Google Scholar

    1478.

    Skelton, P. H., L. Risch, and L. D. G. De Vos. 1984. On the generic identity of the Gephyroglanis catfishes from southern Africa (Pisces, Siluroidei, Bagridae). Revue de Zoologie Africaine 98(2):337–372. Google Scholar

    1479.

    Šlechtová, V., J. Bohlen, and H.-H. Tan. 2007. Families of Cobitoidea (Teleostei; Cypriniformes) as revealed from nuclear genetic data and the position of the mysterious genera Barbucca, Psilorhynchus, Serpenticobitis and Vaillantella. Molecular Phylogenetics and Evolution 44(3):1358–1365. Google Scholar

    1480.

    Smith, C. L. 1975. The evolution of hermaphroditism in fishes. In: R. Reinboth, ed. Intersexuality in the Animal Kingdom. Berlin: Springer. pp. 295–310. Google Scholar

    1481.

    Smith, D. G. 1984. A redescription of the rare eel Myroconger compressus (Pisces: Myrocongridae), with notes on its osteology, relationships and distribution. Copeia 1984(3):585–594. Google Scholar

    1482.

    Smith, D. G., and G. D. Johnson. 2007. A new species of Pteropsaron (Teleostei: Trichonotidae: Hemerocoetinae) from the western Pacific, with notes on related species. Copeia 2007(2):364–377. Google Scholar

    1483.

    Smith, J. A. 1865 March 23. Report in [Edinburgh] Daily Review, NO. 1242, 2. Google Scholar

    1484.

    Smith, J. L. B. 1935. The “Galjoen” fishes of South Africa. Transactions of the Royal Society of South Africa 23(3):265–276. Google Scholar

    1485.

    Smith, J. L. B. 1955a. The fishes of Aldabra. Part II. Annals and Magazine of Natural History, Series 12, 8(93):689–697. Google Scholar

    1486.

    Smith, J. L. B. 1955b. The fishes of the family Pomacanthidae in the western Indian Ocean. Annals and Magazine of Natural History, Series 12, 8(89):377–384. Google Scholar

    1487.

    Smith, M. L., and M. W. Hahn. 2022. The frequency and topology of pseudoorthologs. Systematic Biology 71(3):649–659. Google Scholar

    1488.

    Smith, W. L. 2005. The limits and relationships of mail-cheeked fishes (Teleostei: Percomorpha) and the evolution of venom in fishes [dissertation]. Columbia University. 414 pp. Google Scholar

    1489.

    Smith, W. L., and M. S. Busby. 2014. Phylogeny and taxonomy of sculpins, sandfishes, and snailfishes (Perciformes: Cottoidei) with comments on the phylogenetic significance of their early-life-history specializations. Molecular Phylogenetics and Evolution 79:332–352. Google Scholar

    1490.

    Smith, W. L., and M. T. Craig. 2007. Casting the percomorph net widely: the importance of broad taxonomic sampling in the search for the placement of serranid and percid fishes. Copeia 2007(1):35–55. Google Scholar

    1491.

    Smith, W. L., E. Everman, and C. Richardson. 2018. Phylogeny and taxonomy of flatheads, scorpionfishes, sea robins, and stonefishes (Percomorpha: Scorpaeniformes) and the evolution of the lachrymal saber. Copeia 106(1):94–119. Google Scholar

    1492.

    Smith, W. L., M. J. Ghedotti, O. Domínguez-Domínguez, C. D. Mcmahan, E. Espinoza, R. P. Martin, M. G. Girard, and M. P. Davis. 2022. Investigations into the ancestry of the Grape-eye Seabass (Hemilutjanus macrophthalmos) reveal novel limits and relationships for the Acropomatiformes (Teleostei: Percomorpha). Neotropical Ichthyology 20(3):e210160.  https://doi.org/10.1590/1982-0224-2021-0160 Google Scholar

    1493.

    Smith, W. L., J. H. Stern, M. G. Girard, and M. P. Davis. 2016. Evolution of venomous cartilaginous and ray-finned fishes. Integrative and Comparative Biology 56(5):950–961. Google Scholar

    1494.

    Smith, W. L., and W. C. Wheeler. 2004. Polyphyly of the mail-cheeked fishes (Teleostei: Scorpaeniformes): evidence from mitochondrial and nuclear sequence data. Molecular Phylogenetics and Evolution 32(2):627–646. Google Scholar

    1495.

    Smith, W. L.. 2006. Venom evolution widespread in fishes: a phylogenetic road map for the bioprospecting of piscine venoms. Journal of Heredity 97(3):206–217. Google Scholar

    1496.

    Smith-Vaniz, W. F. 1984. Carangidae: relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 522–530. (Special Publication 1.) Google Scholar

    1497.

    Song, H. Y., K. Mabuchi, T. P. Satoh, J. A. Moore, Y. Yamanoue, M. Miya, and M. Nishida. 2014. Mitogenomic circumscription of a novel percomorph fish clade mainly comprising “Syngnathoidei“(Teleostei). Gene 542(2):146–155. Google Scholar

    1498.

    Song, S., J. Zhao, and C. Li. 2017. Species delimitation and phylogenetic reconstruction of the sinipercids (Perciformes: Sinipercidae) based on target enrichment of thousands of nuclear coding sequences. Molecular Phylogenetics and Evolution 111:44–55. Google Scholar

    1499.

    Sorbini, L. 1970. Un nuovo genere fossile nell'ittiofauna di M. Bolca: Eolates nov. gen. Memorie del Museo Civico di Storia Naturale di Verona, prima serie 18:11–29. Google Scholar

    1500.

    Sorbini, L. 1981. The Cretaceous fishes of Nardò. I. Order Gasterosteiformes (Pisces). Bollettino del Museo Civico di Storia Naturale di Verona 8:1–27. Google Scholar

    1501.

    Sparks, J. S., P. V. Dunlap, and W. L. Smith. 2005. Evolution and diversification of a sexually dimorphic luminescent system in ponyfishes (Teleostei: Leiognathidae), including diagnoses for two new genera. Cladistics 21(4):305–327. Google Scholar

    1502.

    Sparks, J. S., R. C. Schelly, W. L. Smith, M. P. Davis, D. Tchernov, V. A. Pieribone, and D. F. Gruber. 2014. The covert world of fish biofluorescence: a phylogenetically widespread and phenotypically variable phenomenon. PLOS ONE 9(1):e83259.  https://doi.org/10.1371/journal.pone.0083259 Google Scholar

    1503.

    Sparks, J. S., and W. L. Smith. 2004a. Phylogeny and biogeography of cichlid fishes (Teleostei: Perciformes: Cichlidae). Cladistics 20(6):501–517. Google Scholar

    1504.

    Sparks, J. S.. 2004b. Phylogeny and biogeography of the Malagasy and Australasian rainbowfishes (Teleostei: Melanotaenioidei): Gondwanan vicariance and evolution in freshwater. Molecular Phylogenetics and Evolution 33(3):719–734. Google Scholar

    1505.

    Speijer, R. P., H. Pälike, C. J. Hollis, J. J. Hooker, and J. G. Ogg. 2020. The Paleogene Period. In: F. M. Gradstein, J. G. Ogg, M. D. Schmitz, and G. M. Ogg, eds. Geologic Time Scale 2020, Volume 2. Amsterdam: Elsevier. pp. 1087–1140. Google Scholar

    1506.

    Springer, V. G. 1983. Tyson belos, New Genus and Species of Western Pacific Fish (Gobiidae, Xenisthminae), with Discussions of Gobioid Osteology and Classification. Washington, DC: Smithsonian Institution Press. 40 pp. (Smithsonian Contributions to Zoology 390.) Google Scholar

    1507.

    Springer, V. G. 1993. Definition of the suborder Blennioidei and its included families (Pisces: Perciformes). Bulletin of Marine Science 52:472–495. Google Scholar

    1508.

    Springer, V. G., and W. C. Freihofer. 1976. Study of the Monotypic Fish Family Pholidichthyidae (Perciformes). Washington, DC: Smithsonian Institution Press. 43 pp. (Smithsonian Contributions to Zoology 216.) Google Scholar

    1509.

    Springer, V. G., and G. D. Johnson. 2004. Study of the Dorsal Gill-arch Musculature of Teleostome Fishes, with Special Reference to the Actinopterygii. Washington, DC: Biological Society of Washington, Smithsonian Institution. 260 pp. (Bulletin of the Biological Society of Washington 11.) Google Scholar

    1510.

    Springer, V. G.. 2015. The gill-arch musculature of Protanguilla, the morphologically most primitive eel (Teleostei: Anguilliformes), compared with that of other putatively primitive extant eels and other elopomorphs. Copeia 103(3):595–620. Google Scholar

    1511.

    Springer, V. G., and T. M. Orrell. 2004. Appendix: phylogenetic analysis of 147 families of acanthomorph fishes based primarily on dorsal gill-arch muscles and skeleton. In: V. G. Springer, G. D. Johnson, T. M. Orrell, and K. Darrow, eds. Study of the Dorsal Gill-arch Musculature of Teleostome Fishes, with Special Reference to the Actinopterygii. Washington, DC: Biological Society of Washington, Smithsonian Institution. pp. 236–260. (Bulletin of the Biological Society of Washington 11.) Google Scholar

    1512.

    Stanhope, M. J., V. G. Waddell, O. Madsen, W. De Jong, S. B. Hedges, G. C. Cleven, D. Kao, and M. S. Springer. 1998. Molecular evidence for multiple origins of Insectivora and for a new order of endemic African insectivore mammals. Proceedings of the National Academy of Sciences of the United States of America 95(17):9967–9972. Google Scholar

    1513.

    Starks, E. C. 1904. The osteology of some berycoid fishes. Proceedings of the United States National Museum 27(1366):601–619. Google Scholar

    1514.

    Starks, E. C. 1908. The characters of Atelaxia, a new suborder of fishes. Bulletin of the Museum of Comparative Zoology 52(2):1–22. Google Scholar

    1515.

    Steindachner, F. 1867. Notizen. Anzeiger der Kaiserlichen Akademie der Wissenschaften, Mathematisch-Naturwissenschaftliche Classe 4:119–120. Google Scholar

    1516.

    Steindachner, F. 1875. Ichthyologische Beiträge (II). I. Die Fische von Juan Fernandez in den Sammlungen des Wiener Museums. II. Über einige neue Fischarten von der Ost- und Westküste Süd-Amerikas. Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften, Mathematisch-Naturwissenschaftliche Classe 71(1. Abth.):443–480. Google Scholar

    1517.

    Steindachner, F. 1878. Ichthyologische Beiträge (VII). Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften, Mathematisch-Naturwissenschaftliche Classe 78(1. Abth.):377–400. Google Scholar

    1518.

    Steindachner, F. 1880. Beitäge zur Kenntniss der Flussfische Südamerikas (II) und Ichthyologische Beiträge (IX). Anzeiger der Akademie der Wissenschaften in Wien, Mathematisch-Naturwissenschaftliche Classe 17(19):157–159. Google Scholar

    1519.

    Steindachner, F. 1908. Über eine während der brasilianischen Expedition entdeckte Brachyplatystoma-Art aus dem Rio Parnahyba und über eine dicht gefleckte und gestrichelte Varietät von Giton fasciatus aus den Gewässern von Santos (Staat Sao Paulo). Anzeiger der Kaiserlichen Akademie der Wissenschaften, Wien, Mathematisch-Naturwissenschaftliche Classe 45(9):126–130. Google Scholar

    1520.

    Stewart, D. J. 2013a. A new species of Arapaima (Osteoglossomorpha: Osteoglossidae) from the Solimões River, Amazonas State, Brazil. Copeia 2013(3):470–476. Google Scholar

    1521.

    Stewart, D. J. 2013b. Re-description of Arapaima agassizii (Valenciennes), a rare fish from Brazil (Osteoglossomorpha: Osteoglossidae). Copeia 2013(1):38–51. Google Scholar

    1522.

    Stewart, K. M. 2001. The freshwater fish of Neogene Africa (Miocene–Pleistocene): systematics and biogeography. Fish and Fisheries 2(3):177–230. Google Scholar

    1523.

    Stewart, K. M. 2003. Fossil fish remains from the Mio-Pliocene deposits at Lothagam, Kenya. In: M. G. Leakey and J. M. Harris, eds. Lothagam: The Dawn of Humanity in Eastern Africa. New York: Columbia University Press. pp. 75–111. Google Scholar

    1524.

    Steyskal, G. C. 1980. The grammar of family-group names as exemplified by those of fishes. Proceedings of the Biological Society of Washington 93(1):168–177. Google Scholar

    1525.

    Stiassny, M. L. J. 1986. The limits and relationships of acanthomorph teleosts. Journal of Zoology 1(2):411–460. Google Scholar

    1526.

    Stiassny, M. L. J. 1990. Notes on the anatomy and relationships of the bedotiid fishes of Madagascar, with a taxonomic revision of the genus Rheocles (Atherinomorpha: Bedotiidae). American Museum Novitates 2979:1–33. Google Scholar

    1527.

    Stiassny, M. L. J. 1993. What are grey mullets? Bulletin of Marine Science 52:197–219. Google Scholar

    1528.

    Stiassny, M. L. J. 1996. Basal ctenosquamate relationships and the interrelationships of the myctophiform (Scopelomorph) fishes. In: M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson, eds. Interrelationships of Fishes. San Diego: Academic Press. pp. 174–191. Google Scholar

    1529.

    Stiassny, M. L. J., and J. S. Jensen. 1987. Labroid intrarelationships revisited: morphological complexity, key innovations, and the study of comparative diversity. Bulletin of the Museum of Comparative Zoology 151(5):269–319. Google Scholar

    1530.

    Stiassny, M. L. J., and J. A. Moore. 1992. A review of the pelvic girdle of acanthomorph fishes, with comments on hypotheses of acanthomorph intrarelationships. Zoological Journal of the Linnean Society 104(3):209–242. Google Scholar

    1531.

    Stiassny, M. L. J., R. C. Schelly, and V. Mamonekene. 2009. A new Alestes (Characiformes, Alestidae) from the Mpozo River in the Democratic Republic of Congo. Copeia 2009(1):110–116. Google Scholar

    1532.

    Stiassny, M. L. J., E. O. Wiley, G. D. Johnson, and M. R. De Carvalho. 2004. Gnathostome fishes. In: J. Cracraft and M. J. Donoghue, eds. Assembling the Tree of Life. New York: Oxford University Press. pp. 410–429. Google Scholar

    1533.

    Stiller, J., G. Short, H. Hamilton, N. Saarman, S. Longo, P. Wainwright, G. W. Rouse, and W. B. Simison. 2022. Phylogenomic analysis of Syngnathidae reveals novel relationships, origins of endemic diversity and variable diversification rates. BMC Biology 20:75.  https://doi.org/10.1186/s12915-022-01271-w Google Scholar

    1534.

    Stout, C. C., M. Tan, A. R. Lemmon, E. M. Lemmon, and J. W. Armbruster. 2016. Resolving Cypriniformes relationships using an anchored enrichment approach. BMC Evolutionary Biology 16:244.  https://doi.org/10.1186/s12862-016-0819-5 Google Scholar

    1535.

    Straube, N., C. Li, M. Mertzen, H. Yuan, and T. Moritz. 2018. A phylogenomic approach to reconstruct interrelationships of main clupeocephalan lineages with a critical discussion of morphological apomorphies. BMC Evolutionary Biology 18:158.  https://doi.org/10.1186/s12862-018-1267-1 Google Scholar

    1536.

    Streelman, J. T., and S. A. Karl. 1997. Reconstructing labroid evolution with single-copy nuclear DNA. Proceedings of the Royal Society of London B, Biological Sciences 264(1384):1011–1020. Google Scholar

    1537.

    Stringer, G., and W. Schwarzhans. 2021. Upper Cretaceous teleostean otoliths from the Severn Formation (Maastrichtian) of Maryland, USA, with an unusual occurrence of Siluriformes and Beryciformes and the oldest Atlantic coast Gadiformes. Cretaceous Research 125:104867.  https://doi.org/10.1016/j.cretres.2021.104867 Google Scholar

    1538.

    Stundl, J., A. Pospisilova, D. Jandzik, P. Fabian, B. Dobiasova, M. Minarik, B. D. Metscher, V. Soukup, and R. Cerny. 2019. Bichir external gills arise via heterochronic shift that accelerates hyoid arch development. eLife 8:e43531.  https://doi.org/10.7554/eLife.43531 Google Scholar

    1539.

    Su, Y., H.-C. Lin, and H.-C. Ho. 2022a. Hoplostethus roseus, a new roughy fish from the western Pacific based on morphology and DNA barcoding (family Trachichthyidae). Journal of Fish Biology 101(3):441–452. Google Scholar

    1540.

    Su, Y.. 2022b. A new cryptic species of the pineapple fish genus Monocentris (family Monocentridae) from the western Pacific Ocean, with redescription of M. japonica (Houttuyn, 1782). Zootaxa 5189(1):180–203.  https://doi.org/10.11646/zootaxa.5189.1.18 Google Scholar

    1541.

    Sullivan, J. P., S. Lavoué, and C. D. Hopkins. 2000. Molecular systematics of the African electric fishes (Mormyroidea: Teleostei) and a model for the evolution of their electric organs. Journal of Experimental Biology 203:665–683. Google Scholar

    1542.

    Sullivan, J. P.. 2016. Cryptomyrus: a new genus of Mormyridae (Teleostei, Osteoglossomorpha) with two new species from Gabon, West-Central Africa. ZooKeys 561:117–150. Google Scholar

    1543.

    Sullivan, J. P., J. G. Lundberg, and M. Hardman. 2006. A phylogenetic analysis of the major groups of catfishes (Teleostei: Siluriformes) using rag1 and rag2 nuclear gene sequences. Molecular Phylogenetics and Evolution 41(3):636–662. Google Scholar

    1544.

    Sullivan, J. P., J. Muriel-Cunha, and J. G. Lundberg. 2013. Phylogenetic relationships and molecular dating of the major groups of catfishes of the Neotropical superfamily Pimelodoidea (Teleostei, Siluriformes). Proceedings of the Academy of Natural Sciences of Philadelphia 162(1):89–110. Google Scholar

    1545.

    Sullivan, J. P., Z. Peng, J. G. Lundberg, J. Peng, and S. He. 2008. Molecular evidence for diphyly of the Asian catfish family Amblycipitidae (Teleostei: Siluriformes) and exclusion of the South American Aspredinidae from Sisoroidea. Proceedings of the Academy of Natural Sciences of Philadelphia 157(1):51–65. Google Scholar

    1546.

    Sun, Z. Y., A. Tintori, X. Yaozhong, C. Lombardo, N. Peigang, and J. Dayong. 2017. A new non-parasemionotiform order of the Halecomorphi (Neopterygii, Actinopterygii) from the Middle Triassic of Tethys. Journal of Systematic Palaeontology 15(3):223–240. Google Scholar

    1547.

    Suttkus, R. D. 1963. Order Lepisostei. In: B. Bigelow, and W. C. Schroeder, eds. Soft-rayed bony fishes. Class Osteichthyes. Sturgeons, Gars, Tarpon, Ladyfish, Bonefish, Salmon, Charrs, Anchovies, Herring, Shads, Smelt, Capelin, et al. New Haven, CT: Sears Foundation of Marine Research, Yale University. pp. 61–88. (Memoir 1, Fishes of the Western North Atlantic 3.) Google Scholar

    1548.

    Suzuki, D., M. C. Brandley, and M. Tokita. 2010. The mitochondrial phylogeny of an ancient lineage of ray-finned fishes (Polypteridae) with implications for the evolution of body elongation, pelvic fin loss, and craniofacial morphology in Osteichthyes. BMC Evolutionary Biology 10:21.  https://doi.org/10.1186/1471-2148-10-21 Google Scholar

    1549.

    Swainson, W. 1838. On the Natural History and Classification of Fishes, Amphibians, and Reptiles, or Monocardian Animals, Volume 1. London: Longman, Orme, Brown, Green, and Longmans. 368 pp. Google Scholar

    1550.

    Swann, J. B., S. J. Holland, M. Petersen, T. W. Pietsch, and T. Boehm. 2020. The immunogenetics of sexual parasitism. Science 369(6511):1608–1615. Google Scholar

    1551.

    Sytchevskaya, E. K. 1999. Freshwater fish fauna from the Triassic of northern Asia. In: G. Arratia and H.-P. Schultze, eds. Mesozoic Fishes 2: Systematics and Fossil Record. Munich: Verlag Dr. Friedrich Pfeil. pp. 445–468. Google Scholar

    1552.

    Sytchevskaya, E. K., and A. M. Prokofiev. 2002. First findings of Xiphioidea (Perciformes) in the late Paleocene of Turkmenistan. Journal of Ichthyology 42(3):227–237. Google Scholar

    1553.

    Tagliacollo, V. A., M. J. Bernt, J. M. Craig, C. Oliveira, and J. S. Albert. 2016. Model-based total evidence phylogeny of Neotropical electric knifefishes (Teleostei, Gymnotiformes). Molecular Phylogenetics and Evolution 95:20–33. Google Scholar

    1554.

    Takezaki, N. 2021. Resolving the early divergence pattern of teleost fish using genome-scale data. Genome Biology and Evolution 13(5):evab052.  https://doi.org/10.1093/gbe/evab118 Google Scholar

    1555.

    Tan, M., and J. W. Armbruster. 2018. Phylogenetic classification of extant genera of fishes of the order Cypriniformes (Teleostei: Ostariophysi). Zootaxa 4476(1):6–39.  https://doi.org/10.11646/zootaxa.4476.1.4 Google Scholar

    1556.

    Tang, K. L., M. K. Agnew, M. V. Hirt, T. Sado, L. M. Schneider, J. Freyhof, Z. Sulaiman, E. Swartz, C. Vidthayanon, M. Miya, K. Saitoh, A. M. Simons, R. M. Wood, and R. L. Mayden. 2010. Systematics of the subfamily Danioninae (Teleostei: Cypriniformes: Cyprinidae). Molecular Phylogenetics and Evolution 57(1):189–214. Google Scholar

    1557.

    Tang, K. L., P. B. Berendzen, E. O. Wiley, J. F. Morrissey, R. Winterbottom, and G. D. Johnson. 1999. The phylogenetic relationships of the suborder Acanthuroidei (Teleostei: Perciformes) based on molecular and morphological evidence. Molecular Phylogenetics and Evolution 11(3):415–425. Google Scholar

    1558.

    Tang, K. L., and C. Fielitz. 2013. Phylogeny of moray eels (Anguilliformes: Muraenidae), with a revised classification of true eels (Teleostei: Elopomorpha: Anguilliformes). Mitochondrial DNA 24(1):55–66. Google Scholar

    1559.

    Tang, K. L., D. N. Lumbantobing, and R. L. Mayden. 2013. The phylogenetic placement of Oxygaster van Hasselt, 1823 (Teleostei: Cypriniformes: Cyprinidae) and the taxonomic status of the family-group Name Oxygastrinae Bleeker, 1860. Copeia 2013(1):13–22. Google Scholar

    1560.

    Tang, Q., H. Liu, R. Mayden, and B. Xiong. 2006. Comparison of evolutionary rates in the mitochondrial DNA cytochrome b gene and control region and their implications for phylogeny of the Cobitoidea (Teleostei: Cypriniformes). Molecular Phylogenetics and Evolution 39(2):347–357. Google Scholar

    1561.

    Tang, T., Y. Huang, C. Peng, Y. Liao, Y. Lv, Q. Shi, and B. Gao. 2023. A chromosome-level genome assembly of the Reef Stonefish (Synanceia verrucosa) provides novel insights into stonustoxin (sntx) genes. Molecular Biology and Evolution 40(10):msad215. Google Scholar

    1562.

    Tao, W., R. L. Mayden, and S. He. 2013. Remarkable phylogenetic resolution of the most complex clade of Cyprinidae (Teleostei: Cypriniformes): a proof of concept of homology assessment and partitioning sequence data integrated with mixed model Bayesian analyses. Molecular Phylogenetics and Evolution 66(3):603–616. Google Scholar

    1563.

    Tao, W., L. Yang, R. L. Mayden, and S. He. 2019. Phylogenetic relationships of Cypriniformes and plasticity of pharyngeal teeth in the adaptive radiation of cyprinids. Science China Life Sciences 62:553–565. Google Scholar

    1564.

    Tate, M., R. E. Mcgoran, C. R. White, and S. J. Portugal. 2017. Life in a bubble: the role of the labyrinth organ in determining territory, mating and aggressive behaviours in anabantoids. Journal of Fish Biology 91(3):723–749. Google Scholar

    1565.

    Taverne, L. 1979. Ostéologie, phylogénèse, et systématique des téléostéens fossils et actuels du super-orde de ostéoglossomorphes, troisième partie. Évolution des structures ostéologiques et conclusions générales relatives à la phylogénèse et à la systématique du super-ordre. Addendum. Académie Royale de Belgique Mémoires de la Classe des Sciences 43:1–168. Google Scholar

    1566.

    Taverne, L. 1981. Les actinopterygiens de l'Aptien Inférieur (Tock) d'Helgoland. Mitteilungen aus dem Geologisch-Paläontologischen Institut der Universität Hamburg 51:43–82. Google Scholar

    1567.

    Taverne, L. 1982. Sur Pattersonella formosa (Traquair R.H. 1911) et Nybelinoides brevis (Traquair R.H. 1911), téléosténs salmoniformes argentinoides du Wealdien inférieur de Bernissart, Belgique, précédemment attribués au genre Leptolepis Agassiz L. 1832. Bulletin de l'Institut Royal des Sciences Naturelles de Belgique, Sciences de la Terre 54(3):1–27. Google Scholar

    1568.

    Taverne, L. 1992. Révision de Kermichthys daguini (Arambourg, 1954) nov. gen., téléostéen salmoniforme du Crétacé de la Mésogée eurafricaine. Biologisch Jaarboek Dodonaea 60:76–95. Google Scholar

    1569.

    Taverne, L. 1995. Description de l'appareil de weber du téléostéen crétacé marin Clupavus maroccanus et ses implications phylogénétiques. Belgian Journal of Zoology 125(2):267–282. Google Scholar

    1570.

    Taverne, L. 1997a. Les Clupéomorphes (Pisces, Teleostei) du Cénomanien (Crétace) de Kipala (Kwango, Zaire): ostéologie et phylogénie. Belgian Journal of Zoology 127(1):75–97. Google Scholar

    1571.

    Taverne, L. 1997b. Ostéologie et position systématique d'Audenaerdia casieri, téléostéen clupéomorphe (Pisces) du Santonien (Crétacé) de Vonso, Bas-Zaire. Annales du Musee Royal de l'Afrique Centrale, Serie 8, Sciences Zoologiques (Tervuren, Belgium) 1997:203–213. Google Scholar

    1572.

    Taverne, L. 1998. Les ostéoglossomorphes marins de l'Éocène du Monte Bolca (Italie): Monopterus Volta 1796, Thrissopterus Heckel, 1856 et Foreyichthys Taverne, 1979. Considérations sur la phylogénie de téléostéens ostéoglossomorphes. Studi e Ricerche sui Giacimenti Terziari di Bolca 7:67–158. Google Scholar

    1573.

    Taverne, L. 2002. Les poissons crétacés de Nardò. 18°. Nardoclupea granderi gen. et sp. nov. et (Teleostei, Clupeiformes, Dussumieriinae). Bollettino del Museo Civico di Storia Naturale di Verona, Geologia Paleontologia Preistoria 26:3–23. Google Scholar

    1574.

    Taverne, L. 2004. Les poissons crétacés de Nardò. 18°. Pugliaclupea nolardi gen. et sp. nov. et (Teleostei, Clupeiformes, Clupeidae). Bollettino del Museo Civico di Storia Naturale di Verona, Geologia Paleontologia Preistoria 28:17–28. Google Scholar

    1575.

    Taverne, L. 2007a. Les poissons crétacés de Nardò. 25°. Italoclupea nolfi gen. et sp. nov. (Teleostei, Clupeidae). Bollettino del Museo Civico di Storia Naturale di Verona, Geologia Paleontologia Preistoria 31:21–35. Google Scholar

    1576.

    Taverne, L. 2007b. Les poissons crétacés de Nardò. 26°. Un second Dussumieriinae: Portoselvaggioclupea whiteheadi gen. et sp. nov. et complément a l'étude de Pugliaclupea nolardi (Teleostei, Clupeidae). Bollettino del Museo Civico di Storia Naturale di Verona, Geologia Paleontologia Preistoria 31:37–42. Google Scholar

    1577.

    Taverne, L. 2011a. Les poissons crétacés de Nardò. 33°. Lecceclupea ehiravaensis gen. et sp. nov. (Teleostei, Clupeidae). Bollettino del Museo Civico di Storia Naturale di Verona, Geologia Paleontologia Preistoria 35:3–17. Google Scholar

    1578.

    Taverne, L. 2011b. Ostéologie et relations de Catervariolus (Teleostei, “Pholidophoriformes”) du Jurassique moyen de Kisangani (Formation de Stanleyville) en République Démocratique du Congo. Bulletin de l'Institut Royal des Sciences Naturelles de Belgique, Sciences de la Terre 81:175–212. Google Scholar

    1579.

    Taverne, L. 2013. Osteology and relationships of Songaichthys luctacki gen. and sp. nov. (Teleostei, Ankylophoriformes ord. nov.) from the Middle Jurassic (Songa Limestones) of Kisangani (Democratic Republic of Congo). Geo-Eco-Trop 37:33–52. Google Scholar

    1580.

    Taverne, L., and A. Filleul. 2003. Osteology and relationships of the genus Spaniodon (Teleostei, Salmoniformes) from the Santonian (Upper Cretaceous) of Lebanon. Palaeontology 46(5):927–944. Google Scholar

    1581.

    Tchernavin, V. V. 1947. Further notes on the structure of the bony fishes of the order Lyomeri (Eurypharynx). Zoological Journal of the Linnean Society 41(280):377–393. Google Scholar

    1582.

    Tedesco, P. A., E. Paradis, C. Lévêque, and B. Hugueny. 2017. Explaining global-scale diversification patterns in actinopterygian fishes. Journal of Biogeography 44(4):773–783. Google Scholar

    1583.

    Teletchea, F., V. Laudet, and C. Hänni. 2006. Phylogeny of the Gadidae (sensu Svetovidov, 1948) based on their morphology and two mitochondrial genes. Molecular Phylogenetics and Evolution 38(1):189–199. Google Scholar

    1584.

    Temminck, C. J., and H. Schlegel. 1846. Pisces. In: P. F. Von Siebold, ed. Fauna Japonica, sive, Descriptio animalium, quae in itinere per Japoniam, jussu et auspiciis, superiorum, qui summum in India Batava imperium tenent, suscepto, annis 1823–1830 [Parts 10–14]. Lugduni Batavorum [Leiden]: Apud Auctorem. pp. 173–269. Google Scholar

    1585.

    Temminck, C. J.1850. Pisces. In: P. F. Von Siebold, ed. Fauna Japonica, sive, Descriptio animalium, quae in itinere per Japoniam, jussu et auspiciis, superiorum, qui summum in India Batava imperium tenent, suscepto, annis 1823–1830 [Part 15]. Lugduni Batavorum [Leiden]: Apud Auctorem. pp. 270–324. Google Scholar

    1586.

    Teugels, G. 2003. State of the art of recent siluriform systematics. In: G. Arratia, B. G. Kapoor, M. Chardon, and R. Diogo, eds. Catfishes, Volume 1. Enfield, NH: Science Publishers. pp. 317–352. Google Scholar

    1587.

    Thacker, C. E. 2003. Molecular phylogeny of the gobioid fishes (Teleostei: Perciformes: Gobioidei). Molecular Phylogenetics and Evolution 26(3):354–368. Google Scholar

    1588.

    Thacker, C. E. 2009. Phylogeny of Gobioidei and placement within Acanthomorpha, with a new classification and investigation of diversification and character evolution. Copeia 2009(1):93–104. Google Scholar

    1589.

    Thacker, C. E. 2013. Phylogenetic placement of the European sand gobies in Gobionellidae and characterization of gobionellid lineages (Gobiiformes: Gobioidei). Zootaxa 3619(1):369–382.  https://doi.org/10.11646/zootaxa.3619.3.6 Google Scholar

    1590.

    Thacker, C. E. 2014. Species and shape diversification are inversely correlated among gobies and cardinalfishes (Teleostei: Gobiiformes). Organisms Diversity and Evolution 14(4):419–436. Google Scholar

    1591.

    Thacker, C. E. 2015. Biogeography of goby lineages (Gobiiformes: Gobioidei): origin, invasions and extinction throughout the Cenozoic. Journal of Biogeography 42(9):1615–1625. Google Scholar

    1592.

    Thacker, C. E. 2017. Patterns of divergence in fish species separated by the Isthmus of Panama. BMC Evolutionary Biology 17:111.  https://doi.org/10.1186/s12862-017-0957-4 Google Scholar

    1593.

    Thacker, C. E., and C. Gkenas. 2019. Morphometric convergence among European sand gobies in freshwater (Gobiiformes: Gobionellidae). Ecology and Evolution 9(14):8087–8103. Google Scholar

    1594.

    Thacker, C.E., C. Gkenas, A. Triantafyllidis, S. Malavasi, and I. Leonardos. 2019. Phylogeny, systematics and biogeography of the European sand gobies (Gobiiformes: Gobionellidae). Zoological Journal of the Linnean Society 185(1):212–225. Google Scholar

    1595.

    Thacker, C. E., and M. A. Hardman. 2005. Molecular phylogeny of basal gobioid fishes: Rhyacichthyidae, Odontobutidae, Xenisthmidae, Eleotridae (Teleostei: Perciformes: Gobioidei). Molecular Phylogenetics and Evolution 37(3):858–871. Google Scholar

    1596.

    Thacker, C. E., W. T. Mccraney, R. C. Harrington, T. J. Near, J. J. Shelley, M. Adams, M. P. Hammer, and P. J. Unmack. 2023. Diversification of the sleepers (Gobiiformes: Gobioidei: Eleotridae) and evolution of the root gobioid families. Molecular Phylogenetics and Evolution 186:107841.  https://doi.org/10.1016/j.ympev.2023.107841 Google Scholar

    1597.

    Thacker, C. E., and D. M. Roje. 2009. Phylogeny of cardinalfishes (Teleostei: Gobiiformes: Apogonidae) and the evolution of visceral bioluminescence. Molecular Phylogenetics and Evolution 52(3):735–745. Google Scholar

    1598.

    Thacker, C. E.. 2011. Phylogeny of Gobiidae and identification of gobiid lineages. Systematics and Biodiversity 9(4):329–347. Google Scholar

    1599.

    Thacker, C. E., T. P. Satoh, E. Katayama, R. C. Harrington, R. I. Eytan, and T. J. Near. 2015. Molecular phylogeny of Percomorpha resolves Trichonotus as the sister lineage to Gobioidei (Teleostei: Gobiiformes) and confirms the polyphyly of Trachinoidei. Molecular Phylogenetics and Evolution 93:172–179. Google Scholar

    1600.

    Thacker, C. E., J. J. Shelley, W. T. Mccraney, M. Adams, M. P. Hammer, and P. J. Unmack. 2022. Phylogeny, diversification, and biogeography of a hemiclonal hybrid system of native Australian freshwater fishes (Gobiiformes: Gobioidei: Eleotridae: Hypseleotris). BMC Ecology and Evolution 22:22.  https://doi.org/10.1186/s12862-022-01981-3 Google Scholar

    1601.

    Thacker, C. E., J. J. Shelley, W. T. Mccraney, P. J. Unmack, and M. D. Mcgee. 2022. Delayed adaptive radiation among new zealand stream fishes: joint estimation of divergence time and trait evolution in a newly delineated island species flock. Systematic Biology 71(1):13–23. Google Scholar

    1602.

    Thieme, P., N. K. Schnell, K. Parkinson, and T. Moritz. 2022. Morphological characters in light of new molecular phylogenies: the caudal-fin skeleton of Ovalentaria. Royal Society Open Science 9(1):211605.  https://doi.org/10.1098/rsos.211605 Google Scholar

    1603.

    Thompson, A. W., M. B. Hawkins, E. Parey, D. J. Wcisel, T. Ota, K. Kawasaki, E. Funk, M. Losilla, O. E. Fitch, Q. Pan, ET AL. 2021. The bowfin genome illuminates the developmental evolution of ray-finned fishes. Nature Genetics 53:1373–1384. Google Scholar

    1604.

    Thompson, D. W. 1947. A Glossary of Greek Fishes. London: Oxford University Press. 302 pp. Google Scholar

    1605.

    Tibbetts, I. R. 1991. The trophic ecology, functional morphology and phylogeny of the Hemiramphidae (Beloniformes) [dissertation]. University of Queensland. 388 pp. Google Scholar

    1606.

    Tominaga, Y. 1968. Internal morphology, mutual relationships and systematic position of the fishes belonging to the family Pempheridae. Japanese Journal of Ichthyology 15(2):43–95. Google Scholar

    1607.

    Tornabene, L., Y. J. Chen, and F. Pezold. 2013. Gobies are deeply divided: phylogenetic evidence from nuclear DNA (Teleostei: Gobioidei: Gobiidae). Systematics and Biodiversity 11(3):345–361. Google Scholar

    1608.

    Tornabene, L., R. Manning, D. R. Robertson, J. L. Van Tassell, and C. C. Baldwin. 2023. A new lineage of deep-reef gobies from the Caribbean, including two new species and one new genus (Teleostei: Gobiidae: Gobiosomatini). Zoological Journal of the Linnean Society 197(2):322–343. Google Scholar

    1609.

    Tornabene, L., J. L. Van Tassell, D. R. Robertson, and C. C. Baldwin. 2016. Repeated invasions into the twilight zone: evolutionary origins of a novel assemblage of fishes from deep Caribbean reefs. Molecular Ecology 25(15):3662–3682. Google Scholar

    1610.

    Travers, R. A. 1981. The interarcual cartilage; a review of its development, distribution and value as an indicator of phyletic relationships in euteleostean fishes. Journal of Natural History 15(5):853–871. Google Scholar

    1611.

    Travers, R. A. 1984. A review of the Mastacembeloidei, a suborder of synbranchiform teleost fishes. Part 2: Phylogenetic analysis. Bulletin of the British Museum (Natural History), Zoology 47(2):83–150. Google Scholar

    1612.

    Trewavas, E. 1932. A contribution to the classification of the fishes of the order Apodes, based on the osteology of some rare eels. Proceedings of the Zoological Society of London 102(3):639–659. Google Scholar

    1613.

    Triques, M. L. 1993. Filogenia dos gêneros de Gymnotiformes (Actinopterygii, Ostariophysi), com base em caracteres esqueléticos. Comunicações do Museu de Ciências de PUCRS, série Zoologia 6:85–130. Google Scholar

    1614.

    Troyer, E. M., R. Betancur-R, L. C. Hughes, M. Westneat, G. Carnevale, W. T. White, J. J. Pogonoski, J. C. Tyler, C. C. Baldwin, G. Ortí, ET AL. 2022. The impact of paleoclimatic changes on body size evolution in marine fishes. Proceedings of the National Academy of Sciences 119(29):e2122486119.  https://doi.org/10.1073/pnas.2122486119  Google Scholar

    1615.

    Tsuneki, K. 1992. A systematic survey of the occurrence of the hypothalamic saccus vasculosus in teleost fish. Acta Zoologica 73(1):67–77. Google Scholar

    1616.

    Turanov, S., Y. Kartavtsev, and V. Zemnukhov. 2012. Molecular phylogenetic study of several eelpout fishes (Perciformes, Zoarcoidei) from Far Eastern seas on the basis of the nucleotide sequences of the mitochondrial cytochrome oxidase 1 gene (Co-1). Russian Journal of Genetics 48(2):208–223. Google Scholar

    1617.

    Turanov, S. V., Y. P. Kartavtsev, Y. H. Lee, and D. Jeong. 2017. Molecular phylogenetic reconstruction and taxonomic investigation of eelpouts (Cottoidei: Zoarcales) based on Co-1 and Cyt-b mitochondrial genes. Mitochondrial DNA Part A 28:547–557. Google Scholar

    1618.

    Tyler, J. C. 2005a. A new genus for the surgeon fish Acanthurus gaudryi De Zigno 1887 from the Eocene of Monte Bolca, Italy, a morphologically primitive basal taxon of Acanthuridae (Acanthuroidea, Perciformes). Studi e Ricerche sui Giacimenti Terziari di Bolca 11:149–163. Google Scholar

    1619.

    Tyler, J. C. 2005b. Redescription and basal phylogenetic placement of the acanthurid surgeon fish Gazolaichthys vestenanovae from the Eocene of Monte Bolca, Italy (Perciformes; Acanthuroidea). Studi e Ricerche sui Giacimenti Terziari di Bolca 11:97–117. Google Scholar

    1620.

    Tyler, J. C., and A. F. Bannikov. 1992. A Remarkable New Genus of Tetraodontiform Fish with Features of both Balistids and Ostraciids from the Eocene of Turkmenistan. Washington, DC: Smithsonian Institution Press. 14 pp. (Smithsonian Contributions to Paleobiology 72.) Google Scholar

    1621.

    Tyler, J. C.. 1997. Relationships of the Fossil and Recent Genera of Rabbitfishes (Acanthuroidei: Siganidae). Washington, DC: Smithsonian Institution Press. 35 pp. (Smithsonian Contributions to Paleobiology 84.) Google Scholar

    1622.

    Tyler, J. C.. 2005. Massalongius, gen. and fam. nov., a new clade of acanthuroid fishes (Perciformes, Acanthuridae) from the Eocene of Monte Bolca, Italy, related to Zanclidae. Studi e Ricerche sui Giaciamenti Terziari di Bolca 11:75–95. Google Scholar

    1623.

    Tyler, J. C., P. Bronzi, and A. Ghiandoni. 2000. The Cretaceous fishes of Nardò. 11°. A new genus and species of Zeiformes, Cretazeus rinaldii, the earliest record for the order. Bollettino del Museo Civico di Storia Naturale di Verona, Geologia Paleontologia Preistoria 24:11–28. Google Scholar

    1624.

    Tyler, J. C., G. D. Johnson, I. Nakamura, and B. B. Collette. 1989. Morphology of Luvarus imperialis (Luvaridae), with a Phylogenetic Analysis of the Acanthuroidei (Pisces). Washington, DC: Smithsonian Institution Press. 78 pp. (Smithsonian Contributions to Zoology 485.) Google Scholar

    1625.

    Tyler, J. C., and M. Križnar. 2013. A new genus and species, Slovenitriacanthus saksidai, from southwestern Slovenia, of the Upper Cretaceous basal tetraodontiform fish family Cretatriacanthidae (Plectocretacicoidea). Bollettino del Museo Civico di Storia Naturale di Verona, Geologia Paleontologia Preistoria 37:45–56. Google Scholar

    1626.

    Tyler, J. C., M. Mirzaie, and A. Nazemi. 2006. New genus and species of basal tetraodontoid puffer fish from the Oligocene of Iran, related to the Zignoichthydae (Tetraodontiformes). Bollettino del Museo Civico di Storia Naturale di Verona, Geologia Paleontologia Preistoria 30:49–58. Google Scholar

    1627.

    Tyler, J. C., B. O'toole, and R. Winterbottom. 2003. Phylogeny of the Genera and Families of Zeiform Fishes, with Comments on Their Relationships with Retraodontiforms and Caproids. Washington, DC: Smithsonian Institution Press. 110 pp. (Smithsonian Contributions to Zoology 618.) Google Scholar

    1628.

    Tyler, J. C., and F. Santini. 2002. Review and reconstructions of the tetraodontiform fishes from the Eocene of Monte Bolca, Italy, with comments on related Tertiary taxa. Studi e Ricerche sui Giancimenti Terziari di Bolca 9:47–119. Google Scholar

    1629.

    Tyler, J. C.. 2005. A phylogeny of the fossil and extant zeiform-like fishes, Upper Cretaceous to Recent, with comments on the putative zeomorph clade (Acanthomorpha). Zoologica Scripta 34(2):157–175. Google Scholar

    1630.

    Tyler, J. C., and C. Sorbini. 1999. Phylogeny of the fossil and recent genera of fishes of the family Scatophagidae (Squamipinnes). Bollettino del Museo Civico di Storia Naturale di Verona 23:353–393. Google Scholar

    1631.

    Tyler, J. C., and L. Sorbini. 1996. New Superfamily and Three New Families of Tetraodontiform Fishes from the Upper Cretaceous: The Earliest and Most Morphologically Primitive Plectognaths. Washington, DC: Smithsonian Institution Press. 59 pp. (Smithsonian Contributions to Paleobiology 82.) Google Scholar

    1632.

    Underkoffler, K. E., M. A. Luers, J. R. Hyde, and M. T. Craig. 2018. A taxonomic review of Lampris guttatus (Brünnich 1788) (Lampridiformes; Lampridae) with descriptions of three new species. Zootaxa 4413(3):551–565.  https://doi.org/10.11646/zootaxa.4413.3.9 Google Scholar

    1633.

    Uribe, M. C., H. J. Grier, and V. Mejía-Roa. 2014. Comparative testicular structure and spermatogenesis in bony fishes. Spermatogenesis 4(3):e983400.  https://doi.org/10.4161/21565562.2014.983400 Google Scholar

    1634.

    Vaillant, L. 1902. Résultats zoologiques de l'expédition scientifique Néerlandaise au Bornéo central. Poissons. Notes from the Leyden Museum 24(1–3):1–166. Google Scholar

    1635.

    Van Der Laan, R., W. N. Eschmeyer, and R. Fricke. 2014. Family-group names of Recent fishes. Zootaxa 3882(1):1–230.  https://doi.org/10.11646/zootaxa.3882.1.1 Google Scholar

    1636.

    Van Der Laan, R., R. Fricke, and W. N. Eschmeyer, EDS. 2023. Eschmeyer's Catalog of Fishes: Classification [online database]. Accessed 3 November 2023.  http://www.calacademy.org/scientists/catalog-of-fishes-classification/  Google Scholar

    1637.

    Vari, R. P. 1979. Anatomy, relationships, and classification of the families Citharinidae and Distichodontidae (Pisces, Characoidea). Bulletin of the British Museum (Natural History), Zoology 36(5):261–344. Google Scholar

    1638.

    Vari, R. P. 1983. Phylogenetic Relationships of the Families Curimatidae, Prochilodontidae, Anostomidae, and Chilodontidae (Pisces: Characiformes). Washington, DC: Smithsonian Institution Press. 60 pp. (Smithsonian Contributions to Zoology 378.) Google Scholar

    1639.

    Vari, R. P. 1995. The Neotropical Fish Family Ctenoluciidae (Teleostei: Ostariophysi: Characiformes): Supra and Intrafamilial Phylogenetic Relationships, with a Revisionary Study. Washington, DC: Smithsonian Institution Press. 97 pp. (Smithsonian Contributions to Zoology 564.) Google Scholar

    1640.

    Vaz, D. F. B. 2020. Morphology and systematics of Batrachoidiformes (Percomorphacea: Teleostei) [dissertation]. College of William and Mary. pp. 456. Google Scholar

    1641.

    Vaz, D. F. B., and E. J. Hilton. 2020. The caudal skeleton of Batrachoidiformes (Teleostei: Percomorphacea): a study of morphological diversity, intraspecific variation, and phylogenetic inferences. Zoological Journal of the Linnean Society 189(1):228–286. Google Scholar

    1642.

    Vaz, D. F. B.. 2023. Skeletal ontogeny of the Plainfin Midshipman, Porichthys notatus (Percomorphacea: Batrachoidiformes). Journal of Anatomy 242(3):447–494. Google Scholar

    1643.

    Vecchioni, L., F. Marrone, E. Belaiba, F. Tiralongo, L. Bahri-Sfar, and M. Arculeo. 2019. The DNA barcoding of Mediterranean combtooth blennies suggests the paraphyly of some taxa (Perciformes, Blenniidae). Journal of Fish Biology 94(2):339–344. Google Scholar

    1644.

    Velez, J. A., W. Watson, E. M. Sandknop, W. Arntz, and M. Wolff. 2003. Larval and osteological development of the mote sculpin (Normanichthys crockeri) (Pisces: Normanichthyidae) from the Independencia Bight, Pisco, Peru. Journal of Plankton Research 25(3):279–290. Google Scholar

    1645.

    Vernygora, O. 2020. Systematics of Clupeomorpha (Osteichthyes: Teleostei) with methodological considerations for morphological phylogenetics [dissertation]. University of Alberta. 388 pp. Google Scholar

    1646.

    Vernygora, O., A. M. Murray, and M. V. H. Wilson. 2016. A primitive clupeomorph from the Albian Loon River Formation (Northwest Territories, Canada). Canadian Journal of Earth Sciences 53(4):331–342. Google Scholar

    1647.

    Veysey, A. J., P. M. Brito, and D. M. Martill. 2020. A new crossognathiform fish (Actinopterygii, Teleostei) from the Upper Cretaceous (Turonian) of southern Morocco with hypertrophied fins. Cretaceous Research 114:104207.  https://doi.org/10.1016/j.cretres.2019.104207 Google Scholar

    1648.

    Vialle, R. A., J. E. S. De Souza, K. De Paiva Lopes, D. G. Teixeira, P. De Azevedo Alves Sobrinho, A. M. Ribeiro-Dos-Santos, C. Furtado, T. Sakamoto, F. A. Oliveira Silva, E. Herculano Corrêa De Oliveira, ET AL. 2018. Whole genome sequencing of the Pirarucu (Arapaima gigas) supports independent emergence of major teleost clades. Genome Biology and Evolution 10(9):2366–2379. Google Scholar

    1649.

    Vincent, M., and J. T. Thomas. 2011. Kryptoglanis shajii, an enigmatic subterranean-spring catfish (Siluriformes, incertae sedis) from Kerala, India. Ichthyological Research 58:161–165. Google Scholar

    1650.

    Vinnikov, K. A., R. C. Thomson, and T. A. Munroe. 2018. Revised classification of the righteye flounders (Teleostei: Pleuronectidae) based on multilocus phylogeny with complete taxon sampling. Molecular Phylogenetics and Evolution 125:147–162. Google Scholar

    1651.

    Voigt, E. 1926. Über ein bemerkenswertes Vorkommen neuer Fischotolithen in einem Senongeschiebe von Cöthen in Anhalt. Zeitschrift für Geschiebeforschung 2:172–187. Google Scholar

    1652.

    Volta, G. S. 1796. Ittiolitologia Veronese del Museo Bozziano: Ora Annesso a Quello del Conte Giovambattista Gazola e di altri Gabinetti di Fossili Veronesi. Verona: Dalla Stamperia Giuliari. 323 pp. Google Scholar

    1653.

    Von Der Heyden, S., and C. A. Matthee. 2008. Towards resolving familial relationships within the Gadiformes, and the resurrection of the Lyconidae. Molecular Phylogenetics and Evolution 48(2):764–769. Google Scholar

    1654.

    Voskoboinikova, O. S. 2004. Ontogenetic bases of the origin and relationships of the fishes from the suborder Notothenioidei (Perciformes). Journal of Ichthyology 44(6):418–432. Google Scholar

    1655.

    Wainwright, P. C., W. L. Smith, S. A. Price, K. L. Tang, J. S. Sparks, L. A. Ferry, K. L. Kuhn, R. I. Eytan, and T. J. Near. 2012. The evolution of pharyngognathy: a phylogenetic and functional appraisal of the pharyngeal jaw key innovation in labroid fishes and beyond. Systematic Biology 61(6):1001–1027. Google Scholar

    1656.

    Walbaum, J. J. 1792. Petri Artedi Sueci Genera Piscium: In Quibus Systema Totum Ichthyologiae Proponitur Cum Classibus, Ordinibus, Generum Characteribus, Specierum Differentiis, Observationibus Plurimis. Redactis Speciebus 242 Ad Genera 52. Ichthyologiae Pars III. Grypeswaldiae [Greifswald]: Ant. Ferdin, Rose. 723 pp. Google Scholar

    1657.

    Walters, V. 1957. Protolophotus, a new genus of allotriognath fish from the Oligocene of Iran. Copeia 1957(1):60–61. Google Scholar

    1658.

    Walters, V., and J. E. Fitch. 1960. The families and genera of the lampridiform (Allotriognath) suborder Trachipteroidei. California Fish and Game 46(4):441–451. Google Scholar

    1659.

    Wang, C.-H., C.-H. Kuo, H.-K. Mok, and S.-C. Lee. 2003. Molecular phylogeny of elopomorph fishes inferred from mitochondrial 12S ribosomal RNA sequences. Zoologica Scripta 32(3):231–241. Google Scholar

    1660.

    Wang, J., B. Lu, R. Zan, J. Chai, W. Ma, W. Jin, R. Duan, J. Luo, R. W. Murphy, ET AL. 2016. Phylogenetic relationships of five Asian schilbid genera including Clupisoma (Siluriformes: Schilbeidae). PLOS ONE 11(1):e0145675.  https://doi.org/10.1371/journal.pone.0145675 Google Scholar

    1661.

    Wang, J.-F., H.-Y. Yu, S.-B. Ma, Q. Lin, D.-Z. Wang, and X. Wang. 2023. Phylogenetic and evolutionary comparison of mitogenomes reveal adaptive radiation of lampriform fishes. International Journal of Molecular Sciences 24(10):8756.  https://doi.org/10.3390/ijms24108756 Google Scholar

    1662.

    Wang, Q., L. Purrafee Dizaj, J. Huang, K. Kumar Sarker, C. Kevrekidis, B. Reichenbacher, H. Reza Esmaeili, N. Straube, T. Moritz, and C. Li. 2022. Molecular phylogenetics of the Clupeiformes based on exon-capture data and a new classification of the order. Molecular Phylogenetics and Evolution 175:107590.  https://doi.org/10.1016/j.ympev.2022.107590 Google Scholar

    1663.

    Wang, Q., C. Xu, C. Xu, and R. J. Wang. 2015. Complete mitochondrial genome of the southern catfish (Silurus meridionalis Chen) and Chinese catfish (S. asotus Linnaeus): structure, phylogeny, and intraspecific variation. Genetics and Molecular Research 14(4):18198–18209. Google Scholar

    1664.

    Wang, X. Z., X. N. Gan, J. B. Li, R. L. Mayden, and S. He. 2012. Cyprinid phylogeny based on Bayesian and maximum likelihood analyses of partitioned data: implications for Cyprinidae systematics. Science China Life Sciences 55:761–773. Google Scholar

    1665.

    Wang, X., J. Li, and S. He. 2007. Molecular evidence for the monophyly of East Asian groups of Cyprinidae (Teleostei: Cypriniformes) derived from the nuclear recombination activating gene 2 sequences. Molecular Phylogenetics and Evolution 42(1):157–170. Google Scholar

    1666.

    Washington, B. B., W. M. Eschmeyer, and K. M. Howe. 1984. Sorpaeniformes: relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 438–447. (Special Publication 1.) Google Scholar

    1667.

    Waters, J. M., T. Saruwatari, T. Kobayashi, I. Oohara, R. M. Mcdowall, and G. P. Wallis. 2002. Phylogenetic placement of retropinnid fishes: data set incongruence can be reduced by using asymmetric character state transformation costs. Systematic Biology 51(3):432–449. Google Scholar

    1668.

    Watson, W., A. C. Matarese, and E. G. Stevens. 1984. Trachinoidei: development and relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 554–561. (Special Publication 1.) Google Scholar

    1669.

    Wcisel, D. J., J. T. Howard, J. A. Yoder, and A. Dornburg. 2020. Transcriptome Ortholog Alignment Sequence Tools (TOAST) for phylogenomic dataset assembly. BMC Evolutionary Biology 20:41.  https://doi.org/10.1186/s12862-020-01603-w Google Scholar

    1670.

    Webb, S. A., J. A. Graves, C. Macias-Garcia, A. E. Magurran, D. O. Foighil, and M. G. Ritchie. 2004. Molecular phylogeny of the livebearing Goodeidae (Cyprinodontiformes). Molecular Phylogenetics and Evolution 30(3):527–544. Google Scholar

    1671.

    Weber, M. 1895. Fische von Ambon, Java, Thursday Island, dem Burnett-Fluss und von der Süd-Küste von Neu-Guinea. In: R. W. Semon, ed. Zoologische Forschungsreisen in Australien und dem Malayischen Archipel. Mit Unterstützung des Herrn Dr. Paul von Ritter ausgeführt in den Jahren 1891–1893, von Richard Semon, Volume 8. Jena: G. Fischer. pp. 257–276. Google Scholar

    1672.

    Weitzman, S. H. 1967. The origin of the stomiatoid fishes, with comments on the classification of salmoniform fishes. Copeia 1967(3):507–540. Google Scholar

    1673.

    Weitzman, S. H. 1974. Osteology and evolutionary relationships of the Sternoptychidae with a new classification of stomiatoid families. Bulletin of the American Museum of Natural History 153:327–478. Google Scholar

    1674.

    Wenz, S. 1999. Pliodetes nigeriensis, gen. nov. et. sp. nov., a new semionotid fish from the Lower Cretaceous of Gadoufaoua (Niger Republic): phylogenetic comments. In: G. Arratia and H.-P. Schultze, eds. Mesozoic Fishes 2: Systematics and Fossil Record –. Munich: Verlag Dr. Friedrich Pfeil. pp. 107–120. Google Scholar

    1675.

    Westneat, M. W., and M. E. Alfaro. 2005. Phylogenetic relationships and evolutionary history of the reef fish family Labridae. Molecular Phylogenetics and Evolution 36(2):370–390. Google Scholar

    1676.

    Wettstein, A. 1886. Über die Fischfauna des tertiären Glarnerschiefers. Mémoires de la Société Paléontologique Suisse, Bâle 13:5–103. Google Scholar

    1677.

    White, B. N. 1985. Evolutionary Relationships of the Atherinopsinae (Pisces: Atherinidae). Los Angeles: Natural History Museum of Los Angeles County. 20 pp. (Contributions in Science 368.) Google Scholar

    1678.

    White, B. N., R. J. Lavenberg, and G. E. Mcgowen. 1984. Atheriniformes: development and relationships. In: H. G. Moser, W. J. Richards, D. M. Cohen, M. P. Fahay, J. A.W. Kendall, and S. L. Richardson, eds. Ontogeny and Systematics of Fishes. Lawrence, KS: American Society of Ichthyologists and Herpetologists. pp. 355–362. (Special Publication 1.) Google Scholar

    1679.

    Whitehead, P. J. P. 1962. A contribution to the classification of Clupeoid fishes. Annals and Magazine of Natural History, Series 13, 5(60):737–750. Google Scholar

    1680.

    Whitehead, P. J. P. 1985. Clupeiod Fishes of the World (Suborder Clupeioidei). Part 1. Chirocentridae, Clupeidae and Pristigasteridae. Vol. 7 of FAO Species Catalogue. Rome: Food and Agriculture Organization of the United Nations. 303 pp. (FAO Fisheries Synopsis 125.) Google Scholar

    1681.

    Whitley, G. P. 1945. New sharks and fishes from Western Australia. Part 2. Australian Zoologist 11:1–42. Google Scholar

    1682.

    Wiley, E. O. 1976. The phylogeny and biogeography of fossil and recent gars (Actinopterygii: Lepisosteidae). Lawrence, KS: University of Kansas Museum of Natural History. 111 pp. (Miscellaneous Publications 64.) Google Scholar

    1683.

    Wiley, E. O. 1979. Ventral gill arch muscles and the interrelationships of gnathostomes, with a new classification of the Vertebrata. Zoological Journal of the Linnean Society 67(2):149–179. Google Scholar

    1684.

    Wiley, E. O., and G. D. Johnson. 2010. A teleost classification based on monophyletic groups. In: J. S. Nelson, H.-P. Schultze, and M. V. H. Wilson, eds. Origin and Phylogenetic Interrelationships of Teleosts. Munich: Verlag Dr. Friedrich Pfeil. pp. 123–182. Google Scholar

    1685.

    Wiley, E. O., G. D. Johnson, and W. W. Dimmick. 1998. The phylogenetic relationships of lampridiform fishes (Teleostei: Acanthomorpha), based on a total-evidence analysis of morphological and molecular data. Molecular Phylogenetics and Evolution 10(3):417–425. Google Scholar

    1686.

    Wiley, E. O.. 2000. The interrelationships of acanthomorph fishes: a total evidence approach using molecular and morphological data. Biochemical Systematics and Ecology 28(4):319–350. Google Scholar

    1687.

    Wiley, E. O., and H.-P. Schultze. 1984. Family Lepisosteidae (gars) as living fossils. In: N. Eldredge and S. M. Stanley, eds. Living Fossils. New York: Springer. pp. 160–165. Google Scholar

    1688.

    Williams, J. T. 1990. Phylogenetic Relationships and Revision of the Blenniid Fish Genus Scartichthys. Washington, DC: Smithsonian Institution Press. 30 pp. (Smithsonian Contributions to Zoology 492.) Google Scholar

    1689.

    Williams, R. R. G. 1987. The phylogenetic relationships of the salmoniform fishes based on the suspensorium and its muscles [dissertation]. University of Alberta. 752 pp. Google Scholar

    1690.

    Williford, D., N. S. Beeken, J. Anderson, P. Hajovsky, and R. Weixelman. 2022. Phylogenetic origins and age-based proportions of Malacho (Elops smithi) relative to Ladyfish (Elops saurus): species on the move in the western Gulf of Mexico. Gulf and Caribbean Research 33(1):46–59. Google Scholar

    1691.

    Wilson, A. B., and J. W. Orr. 2011. The evolutionary origins of Syngnathidae: pipefishes and seahorses. Journal of Fish Biology 78(6):1603–1623. Google Scholar

    1692.

    Wilson, C. D., C. F. Mansky, and J. S. Anderson. 2021. A platysomid occurrence from the Tournaisian of Nova Scotia. Scientific Reports 11:8375.  https://doi.org/10.1038/s41598-021-87027-y Google Scholar

    1693.

    Wilson, M. V. H. 1977. Middle Eocene Freshwater Fishes from British Columbia. Toronto: Royal Ontario Museum. 61 pp. (Life Sciences Contributions 113.) Google Scholar

    1694.

    Wilson, M. V. H. 1980. Oldest known Esox (Pisces: Esocidae) part of a new Paleocene teleost fauna from western Canada. Canadian Journal of Earth Sciences 17(3):307–312. Google Scholar

    1695.

    Wilson, M. V. H., D. B. Brinkman, and A. G. Neuman. 1992. Cretaceous Esocoidei (Teleostei): early radiation of the pikes in North American fresh waters. Journal of Paleontology 66(5):839–846. Google Scholar

    1696.

    Wilson, M. V. H., and A. M. Murray. 1996. Early Cenomanian acanthomorph teleost in the Cretaceous Fish Scale Zone, Albian/Cenomanian boundary, Alberta, Canada. In: G. Arratia and G. Viohl, eds. Mesozoic Fishes: Systematics and Paleoecology. Munich: Verlag Dr. Friedrich Pfeil. pp. 369–382. Google Scholar

    1697.

    Wilson, M. V. H.2008. Osteoglossomorpha: phylogeny, biogeography, and fossil record and the significance of key African and Chinese fossil taxa. In: L. Calvin, A. Longbottom, and M. Richter, eds. Fishes and the Break-Up of Pangaea. London: Geological Society of London. pp. 185–219. (Special Publications 295.) Google Scholar

    1698.

    Wilson, M. V. H., and P. Veilleux. 1982. Comparative osteology and relationships of the Umbridae (Pisces: Salmoniformes). Zoological Journal of the Linnean Society 76(4):321–352. Google Scholar

    1699.

    Wilson, M. V. H., and R. R. G. Williams. 1991. New Paleocene genus and species of smelt (Teleostei: Osmeridae) from freshwater deposits of the Paskapoo Formation, Alberta, Canada, and comments on osmerid phylogeny. Journal of Vertebrate Paleontology 11(4):434–451. Google Scholar

    1700.

    Wilson, M. V. H.. 2010. Salmoniform fishes: key fossils, supertree, and possible morphological synapomorphies. In: J. S. Nelson, H.-P. Schultze, and M. V. H. Wilson, eds. Origin and Phylogenetic Interrelationships of Teleosts. Munich: Verlag Dr. Friedrich Pfeil. pp. 379–409. Google Scholar

    1701.

    Winterbottom, R. 1974. The Familial Phylogeny of the Tetraodontiformes (Acanthopterygii: Pisces) as Evidenced by Their Comparative Myology. Washington, DC: Smithsonian Institution Press. 201 pp. (Smithsonian Contributions to Zoology 155.) Google Scholar

    1702.

    Winterbottom, R. 1993a. Myological evidence for the phylogeny of recent genera of surgeonfishes (Percomorpha, Acanthuridae), with comments on the Acanthuroidei. Copeia 1993(1):21–39. Google Scholar

    1703.

    Winterbottom, R. 1993b. Search for the gobioid sister group (Actinopterygii: Percomorpha). Bulletin of Marine Science 52:395–414. Google Scholar

    1704.

    Winterbottom, R., and D. A. Mclennan. 1993. Cladogram versatility: evolution and biogeography of acanthuroid fishes. Evolution 47(5):1557–1571. Google Scholar

    1705.

    Wirtz, P. 1993. Does the larva of Pholidichthys leucotaenia give a clue to the systematic position of the monotypic fish family Pholidichthyidae? In: J. H. Schröder, J. Bauer, and M. Schartl, eds. Trends in Ichthyology: An International Perspective. Oxford: Blackwell Scientific. pp. 237–238. Google Scholar

    1706.

    Woese, C. R., and G. E. Fox. 1977. Phylogenetic structure of prokaryotic domain: the primary kingdoms. Proceedings of the National Academy of Sciences of the United States of America 74:5088–5090. Google Scholar

    1707.

    Woodburne, M. O. 2004. Global events and the North American mammalian biochronology. In: M. O. Woodburne, ed. Late Cretaceous and Cenozoic Mammals of North America. New York: Columbia University Press. pp. 315–343. Google Scholar

    1708.

    Woods, L. P., and P. M. Sonoda. 1973. Order Berycomorphi (Beryciformes). In: D. M. Cohen, J. W. Atz, R. H. Gibbs, F. H. Berry, E. A. Lachner, J. E. Böhlke, G. W. Mead, and D. Merriman, eds. Orders Heteromi (Notacanthiformes), Berycomorphi (Beryciformes), Xenoberyces (Stephanoberyciformes), and Anacanthini (Gadiformes): Halosuriforms, Killfishes, Squirrelfishes and other Beryciforms, Stephanoberyciforms, Grenadiers. New Haven, CT: Sears Foundation for Marine Research, Yale University. pp. 263–396. (Memoir 1, Fishes of the Western North Atlantic 6.) Google Scholar

    1709.

    Woodward, A. S. 1891. Catalogue of the Fossil Fishes in the British Museum (Natural History), Volume 2. London: British Museum (Natural History). 567 pp. Google Scholar

    1710.

    Woodward, A. S. 1901. Catalogue of the Fossil Fishes in the British Museum (Natural History), Volume 4. London: British Museum (Natural History). 639 pp. Google Scholar

    1711.

    Wu, F., D. He, G. Fang, and T. Deng. 2019. Into Africa via docked India: a fossil climbing perch from the Oligocene of Tibet helps solve the anabantid biogeographical puzzle. Science Bulletin 64(7):455–463. Google Scholar

    1712.

    Wu, F., D. Miao, M.-M. Chang, G. Shi, and N. Wang. 2017. Fossil climbing perch and associated plant mega-fossils indicate a warm and wet central Tibet during the late Oligocene. Scientific Reports 7:878.  https://doi.org/10.1038/s41598-017-00928-9 Google Scholar

    1713.

    Wu, X.-W., Y.-Y. Chen, X.-L. Chen, and J.-X. Chen. 1981. A taxonomical system and phylogenetic relationship of the families of the suborder Cyprinoidei (Pisces). Scientia Sinica 24(4):563–572. Google Scholar

    1714.

    Xu, G.-H. 2021. A new stem-neopterygian fish from the Middle Triassic (Anisian) of Yunnan, China, with a reassessment of the relationships of early neopterygian clades. Zoological Journal of the Linnean Society 191(2):375–394. Google Scholar

    1715.

    Xu, G.-H., and M.-M. Chang. 2009. Redescription of †Paralycoptera wui Chang and Chou, 1977 (Teleostei: Osteoglossoidei) from the Early Cretaceous of eastern China. Zoological Journal of the Linnean Society 157(1):83–106. Google Scholar

    1716.

    Xu, G.-H., and K.-Q. Gao. 2011. A new scanilepiform from the Lower Triassic of northern Gansu Province, China, and phylogenetic relationships of non-teleostean Actinopterygii. Zoological Journal of the Linnean Society 161(3):595–612. Google Scholar

    1717.

    Xu, G.-H., K.-Q. Gao, and J. A. Finarelli. 2014. A revision of the middle Triassic scanilepiform fish Fukangichthys longidorsalis from Xinjiang, China, with comments on the phylogeny of the Actinopteri. Journal of Vertebrate Paleontology 34(4):747–759. Google Scholar

    1718.

    Xu, G.-H., and X.-Y. MA. 2016. A Middle Triassic stem-neopterygian fish from China sheds new light on the peltopleuriform phylogeny and internal fertilization. Science Bulletin 61(22):1766–1774. Google Scholar

    1719.

    Xu, G.-H., X.-Y. Ma, and Y. Ren. 2018. Fuyuanichthys wangi gen. et sp. nov. from the Middle Triassic (Ladinian) of China highlights the early diversification of ginglymodian fishes. PeerJ 6:e6054.  https://doi.org/10.7717/peerj.6054 Google Scholar

    1720.

    Xu, G.-H., X.-Y. Ma, F.-X. Wu, and Y. Ren. 2019. A Middle Triassic kyphosichthyiform from Yunnan, China, and phylogenetic reassessment of early ginglymodians. Vertebrata PalAsiatica 57(3):181–204. Google Scholar

    1721.

    Xu, G.-H., and C.-C. Shen. 2015. Panxianichthys imparilis gen. et sp. nov., a new ionoscopiform (Halecomorphi) from the Middle Triassic of Guizhou, China. Vertebrata Palasiatica 53(1):1–15. Google Scholar

    1722.

    Xu, G.-H., and F.-X. Wu. 2012. A deep-bodied ginglymodian fish from the Middle Triassic of eastern Yunnan Province, China, and the phylogeny of lower neopterygians. Chinese Science Bulletin 57:111–118. Google Scholar

    1723.

    Xu, G.-H., and L.-J. Zhao. 2016. A Middle Triassic stem-neopterygian fish from China shows remarkable secondary sexual characteristics. Science Bulletin 61(4):338–344. Google Scholar

    1724.

    Xu, G.-H., L.-J. Zhao, and M. I. Coates. 2014. The oldest ionoscopiform from China sheds new light on the early evolution of halecomorph fishes. Biology Letters 10:20140204.  https://doi.org/10.1098/rsbl.2014.0204 Google Scholar

    1725.

    Xu, G.-H., L.-J. Zhao, K.-Q. Gao, and F.-X. Wu. 2012. A new stem-neopterygian fish from the Middle Triassic of China shows the earliest over-water gliding strategy of the vertebrates. Proceedings of the Royal Society B,Biological Sciences 280(1750):20122261. Google Scholar

    1726.

    Yabe, M. 1985. Comparative osteology and myology of the superfamily Cottoidea (Pisces: Scorpaeniformes), and its phylogenetic classification. Memoirs of the Faculty of Fisheries, Hokkaido University 32:1–130. Google Scholar

    1727.

    Yabe, M., and T. Uyeno. 1996. Anatomical description of Normanichthys crockeri (Scorpaeniformes, incertae sedis: family Normanichthyidae). Bulletin of Marine Science 58:494–510. Google Scholar

    1728.

    Yagishita, N., T. Kobayashi, and T. Nakabo. 2002. Review of monophyly of the Kyphosidae (sensu Nelson, 1994), inferred from the mitochondrial ND2 gene. Ichthyological Research 49:103–108. Google Scholar

    1729.

    Yagishita, N., M. Miya, Y. Yamanoue, S. M. Shirai, K. Nakayama, N. Suzuki, T. P. Satoh, K. Mabuchi, M. Nishida, and T. Nakabo. 2009. Mitogenomic evaluation of the unique facial nerve pattern as a phylogenetic marker within the perciform fishes (Teleostei: Percomorpha). Molecular Phylogenetics and Evolution 53(1):258–266. Google Scholar

    1730.

    Yamada, T., T. Sugiyama, N. Tamaki, A. Kawakita, and M. Kato. 2009. Adaptive radiation of gobies in the interstitial habitats of gravel beaches accompanied by body elongation and excessive vertebral segmentation. BMC Evolutionary Biology 9:145.  https://doi.org/10.1186/1471-2148-9-145 Google Scholar

    1731.

    Yamaguchi, M. 2000. Phylogenetic analyses of myctophid fishes using morphological characters: progress, problems, and future perspectives. Japanese Journal of Ichthyology 47(2):87–107. Google Scholar

    1732.

    Yamanoue, Y., M. Miya, K. Matsuura, N. Yagishita, K. Mabuchi, H. Sakai, M. Katoh, and M. Nishida. 2007. Phylogenetic position of tetraodontiform fishes within the higher teleosts: Bayesian inferences based on 44 whole mitochondrial genome sequences. Molecular Phylogenetics and Evolution 45(1):89–101. Google Scholar

    1733.

    Yang, X., Y. Song, R. Zhang, M. Yu, X. Guo, H. Guo, X. Du, S. Sun, C. Li, X. Mao, G. Fan, and X. Liu. 2023. Unraveling the genomic features, phylogeny and genetic basis of tooth ontogenesis in Characiformes through analysis of four genomes. DNA Research 30(5):dsad022.  https://doi.org/10.1093/dnares/dsad022 Google Scholar

    1734.

    Yedii̇er, S., and D. Bostanci. 2022. Molecular and otolith shape analyses of Scorpaena spp. in the Turkish seas. Turkish Journal of Zoology 46(1):78–92. Google Scholar

    1735.

    Yokoyama, R., and A. Goto. 2005. Evolutionary history of freshwater sculpins, genus Cottus (Teleostei; Cottidae) and related taxa, as inferred from mitochondrial DNA phylogeny. Molecular Phylogenetics and Evolution 36(3):654–668. Google Scholar

    1736.

    Yuan, Z., G.-H. Xu, X. Dai, F. Wang, X. Liu, E. Jia, L. Miao, and H. Song. 2022. A new perleidid neopterygian fish from the Early Triassic (Dienerian, Induan) of South China, with a reassessment of the relationships of Perleidiformes. PeerJ 10:e13448.  https://doi.org/10.7717/peerj.13448 Google Scholar

    1737.

    Zanata, A. M., and R. P. Vari. 2005. The family Alestidae (Ostariophysi, Characiformes): a phylogenetic analysis of a trans-Atlantic clade. Zoological Journal of the Linnean Society 145(1):1–144. Google Scholar

    1738.

    Zaragüeta Bagils, R. 2004. Basal clupeomorphs and ellimmichthyiform phylogeny. In: G. Arratia and A. Tintori, eds. Mesozoic Fishes 3: Systematics, Paleoenvironments and Biodiversity. Munich: Verlag Dr. Friedrich Pfeil. pp. 391–404. Google Scholar

    1739.

    Zehren, S. J. 1979. The Comparative Osteology and Phylogeny of the Beryciformes. Chicago: University of Chicago. 389 pp. (Evolutionary Monographs 1.) Google Scholar

    1740.

    Zehren, S. J. 1987. Osteology and evolutionary relationships of the boarfish genus Antigonia (Teleostei, Caproidae). Copeia 1987(3):564–592. Google Scholar

    1741.

    Zemnukhov, V. V. 2012. Genus Leptostichaeus and its position in the taxonomy of fishes (Perciformes: Zoarcoidei, Stichaeidae). Journal of Ichthyology 52:363–368. Google Scholar

    1742.

    Zhang, H., I. Jarić, D. L. Roberts, Y. He, H. Du, J. Wu, C. Wang, and Q. Wei. 2020. Extinction of one of the world's largest freshwater fishes: lessons for conserving the endangered Yangtze fauna. Science of the Total Environment 710:136242.  https://doi.org/10.1016/j.scitotenv.2019.136242 Google Scholar

    1743.

    Zhang, J.-Y. 1998. Morphology and phylogenetic relationships of †Kuntulunia (Teleostei: Osteoglossomorpha). Journal of Vertebrate Paleontology 18(2):280–300. Google Scholar

    1744.

    Zhang, J.-Y. 2006. Phylogeny of Osteoglossomorpha. Vertebrata PalAsiatica 44(1):43–59. Google Scholar

    1745.

    Zhang, K., Y. Liu, J. Chen, H. Zhang, L. Gong, L. Jiang, L. Liu, Z. Lü, and B. Liu. 2021. Characterization of the complete mitochondrial genome of Macrotocinclus affinis (Siluriformes; Loricariidae) and phylogenetic studies of Siluriformes. Molecular Biology Reports 48:677–689. Google Scholar

    1746.

    Zhang, K., K. Zhu, Y. Liu, H. Zhang, L. Gong, L. Jiang, L. Liu, Z. Lü, and B. Liu. 2021. Novel gene rearrangement in the mitochondrial genome of Muraenesox cinereus and the phylogenetic relationship of Anguilliformes. Scientific Reports 11:2411.  https://doi.org/10.1038/s41598-021-81622-9 Google Scholar

    1747.

    Zhou, J.-J. 1990. The Cyprinidae fossil from middle Miocene of Shanwang Basin. Vertebrata PalAsiatica 28(2):95–127. Google Scholar

    1748.

    Zhuang, X., M. Qu, X. Zhang, and S. Ding. 2013. A comprehensive description and evolutionary analysis of 22 grouper (Perciformes, Epinephelidae) mitochondrial genomes with emphasis on two novel genome organizations. PLOS ONE 8(8):e73561.  https://doi.org/10.1371/journal.pone.0073561 Google Scholar

    1749.

    Zugmayer, E. 1914. Diagnoses de quelques poissons nouveaux provenant des campagnes du yacht Hirondelle II (1911–1913). Bulletin de l'Institut Océanographique de Monaco 288:1–4. Google Scholar

    Appendices

    Appendix 1

    Fossil Taxa Included in the Phylogenetic Trees
    The age intervals of the stages follow the Geologic Time Scale 2020 (Gradstein and Ogg 2020). Fossil species are indicated with a dagger (†) and are listed in an approximate order according to their phylogenetic relationships.

    img-A2yr_03.gif

    Continued

    img-z185-1_03.gif

    Continued

    img-z186-1_03.gif

    Continued

    img-z187-1_03.gif

    Continued

    img-z188-1_03.gif

    Continued

    img-z189-1_03.gif

    Continued

    img-z190-1_03.gif

    Continued

    img-z191-1_03.gif

    Continued

    img-z192-1_03.gif

    Continued

    img-z193-1_03.gif

    Continued

    img-z194-1_03.gif

    Continued

    img-z195-1_03.gif

    Continued

    img-z196-1_03.gif

    Continued

    img-z197-1_03.gif

    Continued

    img-z198-1_03.gif

    Continued

    img-z199-1_03.gif

    Continued

    img-z200-1_03.gif

    Continued

    img-z201-1_03.gif

    Continued

    img-z202-1_03.gif

    Continued

    img-z203-1_03.gif

    Continued

    img-z204-1_03.gif

    Appendix 2
    Classification of Living Lineages of Actinopterygii

    An asterisk (*) identifies family-group names that are monotypic or monogeneric.

    A double dagger (‡) identifies taxa currently not classified in a taxonomic family.

    Names in bold are formal names defined in the clade accounts.

    Continued

    img-z205-6_03.jpg

    Continued

    img-z206-1_03.jpg

    Continued

    img-z207-1_03.jpg

    Continued

    img-z208-1_03.jpg

    Continued

    img-z209-1_03.jpg

    Continued

    img-z210-1_03.jpg

    Continued

    img-z211-1_03.jpg

    Continued

    img-z212-1_03.jpg

    Continued

    img-z213-1_03.jpg

    Continued

    img-z214-1_03.jpg

    Continued

    img-z215-1_03.jpg

    Continued

    img-z216-1_03.jpg

    Continued

    img-z217-1_03.jpg

    Appendix 3
    Taxonomic Index

    Fossil taxa are indicated with a dagger (†).

    Numbers in bold italic are page ranges for clade accounts.

    • †Abisaadia 28

    • †Abisaadia hakelensis 192

    • Acanthistiinae 156, 215

    • Acanthistius 152, 153, 155

    • Acanthistius cinctus 153

    • Acanthomorpha 3, 4, 5, 11, 86, 89, 90, 91, 92–93, 94, 98, 102, 107, 161, 185, 197–206, 211

    • Acanthopterygii 11, 85, 92, 93, 97, 100, 104–105, 106, 107, 112, 138, 185, 199–206, 212

    • Acanthuridae 177, 178, 179, 218

    • Acanthuriformes 9, 112, 153, 173, 174, 175–177, 178, 184, 204–206, 218

    • Acanthuroidei 174, 175, 176, 177, 177–179, 205, 218

    • Acanthurus lineatus 173, 177

    • Acestrorhynchidae 74, 75, 208

    • Acestrorhynchus 65

    • Acheilognathidae 61, 62, 209

    • Achiridae 145, 146, 215

    • Achiropsettidae 145, 146, 215

    • Acipenser sturio 10, 15, 17

    • Acipenseridae 17, 19, 207

    • Acipenseriformes 10, 16, 17, 17, 18, 19, 19, 188, 207

    • Acipenseroidei 17

    • Acropoma 172

    • Acropoma japonicum 152, 171, 173

    • †Acropoma sp. 172

    • Acropomatidae 172, 173, 217

    • Acropomatiformes 9, 109, 112, 153, 159, 168, 171–173, 175, 177, 217

    • Actinopteri 10, 11, 12, 13, 15, 15–17, 20, 186–187, 188, 207

    • Actinopterygii 3, 4, 6,7, 8, 9, 10–11, 12, 13, 16, 17, 21, 25, 78, 107, 186, 207

    • Adrianichthyidae 130, 134, 135, 137, 214

    • Adrianichthys oophorus 133

    • Aenigmachanna 151

    • Aenigmachannidae 151

    • †Aethalionopsis 53, 54

    • †Aethalionopsis robustus 194

    • Aethotaxis 158, 216

    • Aethotaxis mitopteryx 157

    • Afronandus sheljuzhkoi 151

    • Afrotheria 4

    • †Agnevichthys 167

    • †Agnevichthys gretchinae 165, 166, 204

    • Agonidae 162, 163, 216

    • Ailiidae 70, 71, 72, 209

    • †Aipichthyoides 93

    • †Aipichthyoides galeatus 197

    • †Aipichthys 93

    • †Aipichthys minor 197

    • Akysidae 70, 71, 72, 209

    • Albula 27, 30, 31

    • Albula vulpes 27, 29, 30

    • Albulidae 25, 26, 27, 28, 29, 30–31, 31, 185, 191–192, 207

    • Albuliformes 30

    • Albuloidae 30

    • Albuloidei 30

    • Alepisauridae 88, 89, 211

    • Alepisaurus ferox 85, 87

    • Alepocephali 50

    • Alepocephalidae 50, 51, 208

    • Alepocephaliformes 44, 45, 46, 47, 49–51, 75, 76, 77, 78, 79, 85, 208

    • Alepocephaloidea 50

    • Alepocephaloidei 50

    • Alepocephalus bairdii 49

    • Alepocephalus rostratus 49

    • Alestes inferus 72

    • Alestidae 74, 75, 208

    • Allenbatrachus 117

    • Allotriognathi 95

    • Alosidae 48, 49, 208

    • Amarsipidae 125, 213

    • Amarsipus carlsbergi 123, 125

    • Ambassidae 138, 139, 214

    • Ambassis urotaenia 127

    • Amblycipitidae 70, 71, 72, 209

    • Amblyopsidae 98, 99, 135, 211

    • Amia 19, 21, 22

    • Amia calva 10, 15, 19, 20, 21, 22

    • Amiidae 23, 207

    • †Amiopsis 22, 23

    • †Amiopsis lepidota 189

    • Ammodytidae 139, 163, 170, 171, 217

    • †Ampheristus 115

    • †Ampheristus americanus 114, 115, 200

    • Amphichthys 117

    • Amphiliidae 71, 72, 209

    • †Amphiplaga 99

    • †Amphiplaga brachyptera 99, 198

    • †Amphistium 144, 145, 146

    • †Amphistium paradoxum 144, 145, 203

    • Anabantaria 149

    • Anabantidae 151, 152, 215

    • Anabantiformes 149, 152

    • Anabantoidei 142, 148, 149, 150–152, 185, 202–203, 215

    • Anabantomorpha 152

    • Anabantomorphariae 149

    • Anabas testudineus 148, 150

    • Anablepidae 136, 137, 214

    • †Anaethalion 28

    • †Anaethalion zapporum 27, 28, 191

    • Anarhichadidae 166, 167, 216

    • Anchariidae 70, 71, 72, 209

    • †Anchichanna 151, 152

    • †Anchichanna kuldanensis 152, 202

    • †Andinichthyidae 54, 55, 56, 195

    • †Andinichthys bolivianensis 195

    • †Anenchelum 123, 125

    • †Anenchelum eocaenicum 125, 201

    • Anguilla anguilla 32, 35

    • Anguilla rostrata 24, 27

    • †Anguillavus 28

    • †Anguillavus mazeni 192

    • Anguillidae 33, 35, 36, 207

    • Anguilliformes 25, 27, 28, 30, 31, 32–34, 34, 35, 36, 37, 185, 207

    • Anguilloidei 26, 32, 33, 34, 35–36, 185, 207

    • †Ankylophoriformes 20, 21, 187

    • †Ankylophorus similis 187

    • Anomalopidae 108, 111, 212

    • Anoplogaster 108, 111

    • Anoplogastridae 108, 212

    • Anoplopomatidae 159, 160, 161, 162, 216

    • Anostomidae 72, 74, 75, 208

    • Anotophysa 54

    • Anotophysi 54

    • Antennariidae 179, 181, 182, 218

    • Anthiadidae 152, 155, 156, 215

    • Antigonia 100

    • †Apateodus 88, 89

    • †Apateodus corneti 196

    • Apeltes quadracus 163, 164, 165

    • Aphaniidae 136, 137, 214

    • Aphos 117

    • Aphredoderidae 99, 211

    • Aphredoderus sayanus 98, 99

    • Aphyonidae 115, 116

    • Aphyonus gelatinosus 114, 115

    • Aplocheilidae 135, 136, 137, 214

    • Aplocheilus lineatus 135

    • Aplodactylidae 169, 217

    • Aplodactylus 167, 169

    • Aplodactylus 167, 169

    • Apodes 32, 34

    • Apogon imberbis 121

    • Apogonidae 118, 119, 120, 121, 122, 212

    • †Apogonidarum 119, 122

    • †Apogoniscus 119, 122

    • Apogonoidei 9, 106, 118, 119, 121–122, 212

    • Apteronotidae 63, 64, 209

    • Apteronotus albifrons 62

    • Aracanidae 183, 184, 185, 219

    • Arapaima 40, 41, 42

    • Arapaiminae 40, 41

    • †Araripelepidotes 22, 23

    • †Araripelepidotes temnurus 189

    • †Archaemacruroides bratishkoi 104

    • †Archaemacruroides vanknippenbergi 104

    • †Archaeozeus 97, 98

    • †Archaeozeus skamolensis 198

    • †Archaeus 147, 148

    • †Archaeus oblongus 147, 203

    • Argentina sphyraena 77

    • †Argentina voigti 78

    • Argentinidae 77, 78, 210

    • Argentiniformes 43, 46, 50, 75, 76, 77, 77–78, 78, 79, 85, 185, 210

    • Argentinoidea 78

    • Argentinoidei 78

    • †Argestichthys 123, 125

    • †Argestichthys vysotzkyi 125, 201

    • †Argillichthys 88, 89

    • †Argillichthys toombsi 196

    • Ariidae 70, 71, 72, 210

    • Ariomma 123, 125

    • Ariommatidae 125, 213

    • Arrhamphus 134, 135, 214

    • Arrhamphus sclerolepis 134

    • Arripidae 125, 213

    • Arripis 122, 123, 125

    • Arripis trutta 122

    • Artedidraconidae 157

    • †Ascalabos 20, 21

    • Ascalabos voithi 188

    • †Aspidorhynchidae 20, 21, 187

    • †Aspidorhynchus crassus 187

    • Aspredinidae 70, 71, 72, 210

    • Astroblepidae 70, 209

    • Astroblepus 69

    • Astroscopus y–graecum 170

    • †Atacamichthys 20, 21

    • †Atacamichthys greeni 187

    • Atelaxia 95

    • Ateleopodidae 85, 87, 211

    • Ateleopus 90

    • Ateleopus japonicus 85

    • Atherina hepsetus 130, 132

    • Atherinella panamensis 132

    • Atherinidae 132, 133, 213

    • †Atherinidarum 133

    • Atheriniformes 4, 9, 112, 128, 129, 130, 130–131, 133, 135, 138, 185, 202, 213

    • Atherinoidei 9, 85, 129, 130, 131, 132–133, 135, 185, 213

    • Atherinomorpha 112, 131

    • Atherinomorphae 131

    • Atherinopsidae 132, 133, 213

    • Atherinopsis californiensis 132

    • Atherion 132, 133

    • Atherion elymus 132

    • Atherionidae 133, 213

    • †Atolvorator longipectoralis 87, 89

    • Auchenipteridae 70, 71, 72, 210

    • Auchenoglanididae 71, 72, 210

    • †Audenaerdia casieri 49

    • Aulichthys 163, 164

    • Aulichthys japonicus 163, 164, 165

    • Aulopa 88

    • Aulopidae 88, 89, 211

    • Aulopiformes 85, 86, 87, 87–89, 89, 91, 196, 211

    • Aulopus filamentosus 87

    • Aulorhynchidae 163, 164

    • Aulorhynchus 164

    • Aulorhynchus flavidus 163, 164, 165

    • Aulostomidae 127, 213

    • Aulostomus 126, 127

    • Aulotrachichthys prosthemius 107, 109

    • †Australosomus kochi 186

    • †Australosomus 15, 16, 17

    • Austroglanididae 72, 210

    • Austroglanis 71

    • Austrolebias nigripinnis 135

    • †Avitoluvarus 177, 178, 179

    • †Avitoluvarus dianae 205

    • †Avitosmerus 42, 44

    • †Avitosmerus canadensis 193

    • Azygopterus corallinus 165

    • †Babelichthys olneyi 94

    • †Bacchiaichthys 111, 112, 113

    • †Bacchiaichthys zucchiae 200

    • Bachmannia 65, 67, 69

    • †Bachmannia chubutensis 195

    • Badis 151

    • Badis badis 148, 150

    • Bagridae 70, 71, 72, 210

    • †Bajaichthys 97, 98

    • †Bajaichthys elegans 199

    • Balistes vetula 182

    • Balistidae 184, 185, 219

    • Balitoridae 56, 59, 60, 209

    • †Balkaria 183, 184, 185

    • †Balkaria histiopterygia 206

    • Banjos 172

    • Banjosidae 173, 217

    • †Bannikovichthys 123, 125

    • †Bannikovichthys paelignus 125, 200

    • Barbourisia 110

    • Barbourisia rufa 110, 111

    • Barbourisiidae 111, 212

    • Barbucca 59, 60

    • Barbuccidae 60

    • †Barcarenichthys 75, 76, 77

    • †Barcarenichthys joneti 195

    • Barchatus 117

    • Bathyclupeidae 172, 173, 217

    • Bathydraconidae 157, 158, 216

    • Bathygadidae 103, 104, 212

    • Bathylaco nigricans 49

    • Bathylaconoidei 50

    • Bathylagidae 77, 78, 210

    • Bathymaster signatus 165

    • Bathymasteridae 165, 166, 167, 216

    • Bathymyrinae 37

    • Bathysauridae 89, 211

    • Bathysauroides gigas 88, 89

    • Bathysauroididae 89, 211

    • Bathysauropsidae 89, 211

    • Bathysauropsis 88, 89

    • Bathysaurus 88, 89

    • Bathysomi 94

    • Batrachoidaria 117

    • Batrachoides 117

    • Batrachoides pacifici 116

    • Batrachoidida 117

    • Batrachoididae 96, 102, 107, 112, 113, 115, 116–118, 118, 155, 185, 212

    • †Batrachoididarum trapezoidalis 117

    • Batrachoidiformes 117

    • Batrachoidimorpharia 117

    • Batrachoidinae 117, 118

    • Batrachomoeus 117

    • Batrichthys 117

    • †Baugeichthys 28

    • †Baugeichthys caeruleus 191

    • †Beckerophotus 91, 92

    • †Beckerophotus gracilis 197

    • Bedotiidae 132, 133, 213

    • †Bellwoodilabrus 170, 171

    • †Bellwoodilabrus landinii 171, 204

    • Belone 134

    • Belone belone 133

    • Belonidae 133, 134, 135, 214

    • Beloniformes 134

    • Belonoidei 9, 129, 130, 131, 133–135, 135, 185, 202, 213

    • Bembridae 159, 160, 161, 216

    • Bembropidae 155, 156, 159, 216

    • Bembrops 155

    • †Berybolcensis 109, 110

    • Berybolcensis leptacanthus 199

    • Berycidae 105, 108, 110, 111, 212

    • Beryciformes 100, 104, 105, 106, 107, 108, 109–110, 111, 112, 113, 199, 212

    • Berycoidei 106, 109, 110, 110–111, 212

    • †Berycomorus 109, 110

    • Berycomorus firdoussii 199

    • †Berycopsia 98

    • †Berycopsis 98

    • Beryx decadactylus 92, 104, 107, 109, 110, 111

    • †Beurlenichthys 42, 44

    • †Beurlenichthys ouricuriensis 193

    • †Bidenichthys crepidatus 114, 115, 200

    • Bifax 117, 118

    • Bifax lacinia 117

    • Blenniicae 140

    • Blenniidae 137, 139, 140, 141, 214

    • Blenniiformes 128, 129, 130, 137–139, 140, 151, 185, 202, 214

    • Blennioidei 128, 129, 138, 139, 139–141, 214

    • Blennius ocellaris 137, 139

    • †Blochius 147, 148

    • †Blochius longirostris 147, 203

    • †Bobasatraniidae 15, 16, 17, 186

    • Bodianus rufus 170

    • †Bolcabalistes 183, 184, 185

    • †Bolcabalistes varii 206

    • †Bolcapogon 119, 122

    • †Boltyshia 79, 80, 81

    • †Boltyshia brevicauda 80, 196

    • †Boreosomus piveteaui 186

    • †Boreosomus 16, 17

    • Bothidae 145, 146, 215

    • Botia almorhae 59

    • Botiidae 59, 60, 209

    • Bovichtidae 157, 158, 216

    • Bovichtus diacanthus 157

    • Brachiopterygii 15

    • Brama japonica 122

    • Bramidae 123, 125, 213

    • †Brannerion 28

    • †Brannerion latum 191

    • Bregmaceros 103

    • Bregmaceros cantori 102

    • Bregmacerotidae 104, 212

    • †Brembodus ridens 186

    • Brienomyrus 39

    • Brotula barbata 114

    • Bryconidae 74, 75, 208

    • †Bullichthys 28

    • †Bullichthys santanensis 191

    • Butidae 120, 121, 212

    • Bythites fuscus 115

    • Bythitidae 115, 116, 212

    • †Bythitidarum rasmussenae 116, 200

    • Bythitoidei 106, 114, 115, 115–116, 185, 212

    • Caesionidae 175

    • Caesioscorpididae 169, 217

    • Caesioscorpis 169

    • Caesioscorpis theagenes 155, 167

    • Cairnsichthys 132

    • †Calamostoma 126, 127

    • †Calamostoma breviculum 127, 201

    • Callanthiidae 175, 176, 177, 218

    • Callichthyidae 69, 70, 209

    • Calliclinus 140, 141, 214

    • Callionymidae 120, 126, 127, 213

    • Callionymus curvicornis 126

    • Caproidae 175, 176, 177, 218

    • Capromimus 100

    • Capros aper 100

    • †Caprosimilis 173, 176, 177

    • †Caprosimilis carpathicus 205

    • Caraibops trispinosus 172

    • Carangaria 143

    • Carangidae 125, 141, 147, 148, 215

    • Carangiformes 112, 113, 141, 142, 143, 145, 148, 175, 185, 203–204, 215

    • Carangimorphariae 143

    • Carangoidei 142, 143, 146–148, 185, 203, 215

    • †Carangopsis 123, 125

    • †Carangopsis maximus 125, 200

    • Caranx 141, 144, 145

    • Caranx hippos 141, 146

    • Caranx melampygus 141, 146

    • Carapidae 114

    • Carapus 31

    • Carapus acus 114

    • Carapus bermudensis 92, 104, 111

    • Caristiidae 109, 123, 125, 213

    • †Carlomonnius 119, 120, 121

    • †Carlomonnius quasigobius 200

    • †Carrionellus 135, 137

    • †Carrionellus diumortuus 136, 202

    • Cataetyx 115

    • Cataphracti 160

    • †Catervariolus 20, 21

    • †Catervariolus hornemanni 187

    • Catostomidae 56, 57, 58, 59, 60, 61, 209

    • Catostomus catostomus 56

    • †Caturus 22, 23

    • Caturus furcatus 188

    • Caulophrynidae 179, 181, 182, 218

    • Cebidichthyidae 166, 167, 217

    • Cebidichthys 166

    • Centrarchidae 167, 169, 173, 217

    • Centrarchiformes 85, 104, 111, 112, 153, 167, 168, 169, 173, 175, 177, 217

    • Centrarchoidei 167

    • Centrarchus macropterus 167, 173

    • Centriscidae 126, 127, 213

    • Centriscus scutatus 126

    • Centrogenyidae 171, 217

    • Centrogenys 170, 171, 173

    • Centrogenys vaigiensis 170

    • Centrolophidae 123, 125, 213

    • Centrophryne 179

    • Centrophryne spinulosa 179, 181

    • Centrophrynidae 182, 218

    • Centropomidae 143, 215

    • Centropomus 143

    • Centropomus medius 141

    • Cephalopholis cruentata 153

    • Cepolidae 176, 177, 218

    • Ceratiidae 179, 181, 182, 218

    • Cetomimidae 110, 111, 212

    • Cetopsidae 70, 71, 72, 210

    • Cetopsis coecutiens 70

    • Cetostoma regani 110

    • Chaca 71

    • Chacidae 72, 210

    • Chaenopsidae 139, 140, 141, 214

    • Chaetodontidae 175, 176, 177, 218

    • Chalceidae 75, 208

    • Chalceus 74

    • Champsodon 159, 172

    • Champsodontidae 173, 217

    • Chanidae 54, 208

    • Channa 151

    • Channa argus 148, 150

    • Channichthyidae 155, 157, 158, 216

    • Channidae 149, 151, 152, 215

    • †Chanoides 51, 52

    • †Chanoides macropoma 195

    • Chanos 53

    • Chanos chanos 52, 53, 54

    • Characidae 74, 75, 208

    • Characiformes 51, 52, 54, 55, 56, 65, 72–75, 185, 208

    • Characoidea 74

    • Characoidei 74

    • Charax gibbosus 51, 54, 72

    • Charax metae 72

    • †Charitopsis 53, 54

    • †Charitopsis spinosus 194

    • †Charitosomus 53, 54

    • †Charitosomus formosus 194

    • †Chasmoclupea aegyptica 49

    • Chatrabus 117

    • Chaudhuriidae 149, 150, 215

    • Chaunacidae 179, 181, 182, 218

    • †Chaychanus 138, 139

    • †Chaychanus gonzalezorum 138, 139, 202

    • Cheilodactylidae 169, 217

    • Cheilodactylus 167, 169

    • Cheilodipterus quinquelineatus 121

    • Cheimarrichthyidae 171, 217

    • Cheimarrichthys 170

    • Cheimarrichthys fosteri 170, 171

    • Chiasmodontidae 123, 125, 213

    • Chilodontidae 72, 74, 75, 208

    • Chirocentridae 47, 48, 49, 208

    • Chirocentrus 48

    • Chironemidae 169, 217

    • Chironemus 167, 169

    • Chitala 40

    • Chlopsidae 33, 34, 36, 37, 207

    • Chlorophthalmidae 88, 89, 211

    • †Choichix 93

    • †Choichix alvaradoi 197

    • Chologaster cornuta 98

    • †Chondrostei 19

    • †Chondrosteus acipenseroides 186

    • †Chondrosteus 16, 17

    • Chriodorus 134, 135, 214

    • Chriodorus atherinoides 134

    • Chromis chromis 137

    • Cichla temensis 128

    • Cichlidae 128, 138, 139, 170, 171, 214

    • †Cimolichthys 88, 89

    • †Cimolichthys nepaholica 196

    • Cirrhitidae 167, 169, 217

    • Citharidae 144, 145, 146, 215

    • Citharinidae 64, 65, 209

    • Cithariniformes 9, 45, 52, 55, 56, 64–65, 72, 74, 185, 209

    • Citharinoidei 65

    • Citharinus citharus 64

    • Citharinus congicus 72

    • Citharus linguatula 144

    • Cladistia 15

    • Clariidae 70, 71, 72, 210

    • Claroteidae 71, 72, 210

    • Clinidae 139, 140, 141, 214

    • †Clupavus 51, 52

    • †Clupavus maroccanus 195

    • Clupea harengus 42, 46, 48, 75

    • Clupei 47

    • Clupeidae 48, 49, 208

    • Clupeiformes 42, 44, 45, 46, 46–48, 48, 50, 51, 56, 76, 77, 193, 208

    • Clupeocephala 11, 23, 24, 25, 27, 42, 43, 44, 44, 192–193, 208

    • Clupeoidei 45, 47, 48, 48–49, 208

    • Clupeomorpha 23, 25, 47

    • Cobitidae 59, 60, 209

    • Cobitidoidea 59, 60

    • †Cobitis longipectoralis 60

    • †Cobitis nanningensis 60

    • Cobitis taenia 56, 59

    • Cobitoidea 59, 60

    • Cobitoidei 56, 57, 58, 59, 59–60, 185, 209

    • Colletteichthys 117

    • Coloconger 33, 37

    • Colocongridae 37, 207

    • Cololabis 134

    • Conger conger 36, 37

    • Conger oceanicus 32, 36

    • Congiopodidae 159, 160, 161, 216

    • Congiopodus leucopaecilus 160

    • Congridae 26, 33, 37, 207

    • Congrinae 37

    • Congriscus 37

    • Congrogadidae 128, 138, 139, 214

    • Congroidei 26, 33, 34, 36–37, 185, 207

    • Conorhynchos 72, 210

    • Conorhynchos conirostris 71, 72

    • †Corydoras revelatus 69

    • Coryphaena 125, 147, 148

    • Coryphaenidae 148, 215

    • Cottales 163

    • Cottida 160

    • Cottidae 158, 162, 163, 216

    • Cottiformes 163

    • Cottoidae 163

    • Cottoidea 154, 159, 160, 161, 161–163, 216

    • Cottoidei 160, 163

    • Cottus carolinae 153, 158, 161

    • Cottus gobio 161, 162

    • †Cottus otiakensis 163

    • Cranoglanididae 72, 210

    • Cranoglanis 70, 71

    • Creediidae 120, 172, 173, 217

    • Crenimugil crenilabis 137

    • Crenuchidae 74, 75, 208

    • Crenuchus spilurus 72

    • †Cretatriacanthus 183

    • †Cretazeus 101

    • †Cretazeus rinaldii 101, 199

    • †Cretophareodus 41, 42

    • †Cretophareodus alberticus 42, 191

    • Cromeria 54

    • Cryptacanthodes 166

    • Cryptacanthodidae 167, 217

    • Cryptopsaras couesii 179

    • Cryptotremini 140, 141, 214

    • Ctenoluciidae 72, 74, 75, 208

    • †Ctenoplectus 183, 184, 185

    • †Ctenoplectus williamsi 206

    • Ctenosquamata 11, 85, 86, 87, 89–90, 185, 196–197, 211

    • †Ctenothrissa signifer 90, 197

    • †Ctenothrissiformes 89, 90, 197

    • Cubanichthyidae 136, 137, 214

    • Culaea inconstans 164, 165

    • †Cuneatus 22, 23

    • †Cuneatus wileyi 190

    • Curimatidae 72, 74, 75, 208

    • Cyclopsettidae 145, 146, 215

    • Cyclopteridae 158, 162, 163, 216

    • Cyclosquamata 88

    • †Cyclurus 22, 23

    • †Cyclurus kehreri 189

    • Cyema 35

    • Cyema atrum 33

    • Cyematidae 33, 35, 36, 207

    • †Cynoclupea 47, 48

    • †Cynoclupea nelsoni 193

    • Cynodontidae 74, 75, 208

    • Cynoglossidae 145, 146, 215

    • Cyprinidae 58, 61, 62, 209

    • Cypriniformes 42, 44, 51, 52, 54, 55, 56, 56–59, 59, 60, 61, 195, 209

    • Cypriniphysae 58

    • Cyprinodon variegatus 130, 135

    • Cyprinodontidae 136, 137, 214

    • Cyprinodontiformes 7, 137

    • Cyprinodontoidei 9, 129, 130, 131, 135–137, 185, 202, 214

    • Cyprinoidae 58

    • Cyprinoidea 58, 62

    • Cyprinoidei 56, 57, 58, 59, 60–62, 185, 209

    • Cyprinus carpio 42, 44, 51, 54, 56, 60, 135

    • Cyttidae 101, 211

    • Cyttopsis rosea 100

    • Cyttus 100, 101

    • Cyttus australis 100

    • Dactylopteridae 126, 127, 140, 158, 159, 213

    • Dactyloscopidae 139, 140, 141, 214

    • Dactyloscopus lacteus 139

    • Daector 117

    • †Daitingichthys 28

    • †Daitingichthys tischlingeri 191

    • †Dakotaichthys hogansoni 104

    • Dallia 80, 81

    • Dallia admirabilis 80

    • Dallia pectoralis 80

    • †Dalmatichthys 98

    • †Danatinia casca 94, 95

    • Danio rerio 60

    • Danionella 61

    • Danionidae 56, 60, 61, 62, 209

    • †Dapediidae 20, 21, 187

    • †Dapedium noricum 187

    • Dario 151

    • †Dastilbe 53, 54

    • †Dastilbe crandalli 194

    • Datnioides 175

    • †Deltaichthys 30, 31

    • †Deltaichthys albuloides 191

    • Dentatherinidae 132

    • Denticeps 47

    • Denticeps clupeoides 46, 47, 48

    • Denticipitidae 48, 208

    • †Dercetidae 31

    • Derichthyidae 33, 37, 207

    • Derichthys 37

    • Derichthys serpentinus 36

    • Diceratiidae 179, 181, 182, 218

    • Dichistiidae 169, 217

    • Dichistius 167, 169

    • Dictyosoma 166

    • Dinematichthyidae 116, 212

    • Dinematichthyini 116

    • Dinematichthys iluocoeteoides 114, 115

    • Dinolestes lewini 171, 172

    • Dinolestidae 173, 217

    • Dinopercidae 175, 176, 177, 218

    • Diodontidae 183, 184, 185, 219

    • Diplomystes chilensis 65

    • Diplomystes nahuelbutaensis 70

    • Diplomystidae 65, 67, 68, 69, 209

    • †Dipteronotus ornatus 186

    • †Dipteronotus 15, 16, 17

    • Diretmidae 108, 111, 212

    • Diretmoides pauciradiatus 111

    • Diretmus argenteus 92, 104, 107, 109

    • Discoserra pectinodon 16, 186

    • Dissostichus 157, 158, 216

    • Distichodontidae 64, 65, 209

    • Distichodus 65

    • Distichodus mossambicus 64

    • Doederleinia berycoides 172

    • Doradidae 70, 71, 72, 210

    • Dorosomatidae 48, 49, 208

    • †Dorsetichthys 20, 21

    • †Dorsetichthys bechei 187

    • Draconettidae 126, 127, 213

    • Drepane 175, 176

    • Drepaneidae 177, 218

    • †Ductor 147, 148

    • †Ductor vestenae 147, 203

    • Dussumieria 48

    • Dussumieriidae 48, 49, 208

    • †Ebenaqua ritchei 186

    • Echeneidae 147, 148, 215

    • Echeneis naucrates 146

    • †Echidnocephalus 31, 32

    • †Echidnocephalus troscheli 32, 192

    • †Eekaulostomus 126, 127

    • †Eekaulostomus cuevasae 127, 201

    • Ehiravidae 48, 49, 208

    • Elassoma 169

    • Eleginops 157

    • Eleginops maclovinus 157

    • Eleginopsidae 158, 216

    • †Eleogobius 119, 120, 121

    • †Eleogobius gaudanti 200

    • Eleotridae 120, 121, 212

    • Eleotris pisonis 119

    • †Ellimmichthyiformes 44, 46, 193

    • Ellopostoma 59

    • Ellopostomatidae 60, 209

    • Elopidae 29, 207

    • Elopiformes 25, 26, 27, 28, 28–29, 56, 185, 191, 207

    • Elopocephalai 28

    • Elopoidei 28, 29

    • †Elopoides 29

    • †Elopoides tomassoni 191

    • Elopomorpha 23, 25, 26, 27, 27–28, 31, 33, 42, 44, 47, 76, 185, 191–192, 207

    • Eloposteoglossocephala 25

    • Elops 29

    • Elops saurus 23, 27, 28

    • Elops smithi 29

    • †Elopsomolos 29

    • †Elopsomolos frickhingeri 28, 29, 191

    • Embiotoca jacksoni 137

    • Embiotoca lateralis 127

    • Embiotocidae 128, 138, 139, 140, 170, 171, 214

    • Emmelichthyidae 176, 177, 218

    • Encheli 34

    • †Enchelurus 28

    • †Enchelurus anglicus 192

    • †Enchodontoidei 88, 89, 196

    • †Enchodus zimapanensis 196

    • Engraulidae 47, 48, 49, 208

    • Engraulis encrasicolus 23, 24, 42, 44, 46, 48

    • †Engraulis evolans 135

    • Enoplosidae 169, 217

    • Enoplosus armatus 167, 169

    • †Eoalosa 48, 49

    • †Eoalosa janvieri 193

    • †Eoanabas 151, 152

    • †Eoanabas thibetana 152, 203

    • †Eoapogon 119, 122

    • †Eoaulostomus 126, 127

    • †Eoaulostomus bolcensis 127, 202

    • †Eobothus 146

    • †Eobothus minimus 144, 145, 204

    • †Eobuglossus 146

    • †Eobuglossus eocenicus 144, 145, 204

    • †Eochanna chorlakkiensis 152

    • †Eocitharinus 55, 56

    • †Eocitharinus macrognathus 55, 195

    • †Eocoris bloti 171

    • †Eoengraulis fasoloi 49

    • †Eokrefftia prediaphus 92

    • †Eolates 141, 143, 203

    • †Eolates gracilis 143, 203

    • †Eolophotes 94, 95

    • †Eolophotes lenis 197

    • †Eomola 183, 184, 185

    • †Eomola bimaxillaria 206

    • †Eomyctophum 91, 92

    • †Eomyctophum broncus 92, 197

    • †Eophryne 182

    • †Eoplectus 183, 184, 185

    • †Eoplectus bloti 206

    • †Eoscatophagus 173, 176, 177

    • †Eoscatophagus frontalis 204

    • †Eosiganus 173, 176, 177

    • †Eosiganus kumaensis 205

    • †Eosphaeramia 119, 122

    • †Eospinus 183, 184, 185

    • †Eospinus daniltshenkoi 206

    • †Eosternoptyx discoidalis 83

    • †Eosynanceja brabantica 156, 160, 161

    • Ephippidae 175, 176, 177, 218

    • Epigonidae 172, 173, 217

    • Epigonidarum tyassminensis 172

    • Epinephelidae 152, 155, 156, 216

    • Epipneusta 16

    • †Erichalcis 42, 44

    • †Erichalcis arcta 193

    • †Erismatopterus 99

    • †Erismatopterus levatus 99, 198

    • Erpetoichthys calabaricus 13, 15

    • Erythrinidae 72, 74, 75, 208

    • Esocidae 4, 9, 43, 50, 75, 76, 77, 78, 79, 79–81, 135, 185, 196, 210

    • Esociformes 80

    • Esocinae 80

    • Esocoidea 80

    • Esocoidei 80

    • Esox 79, 80, 81

    • Esox lucius 78, 79

    • †Esox tiemani 80

    • Esselenia 166

    • Esselenichthys 166

    • †Estesesox 78, 79

    • †Estesesox foxi 196

    • Etrumeus 48

    • Euacanthomorphacea 107

    • Euacanthopterygii 107

    • Euclichthyidae 104, 212

    • Euclichthys 103

    • Euleptorhamphidae 134, 135, 214

    • Euleptorhamphus 134

    • Eulophiidae 165, 166, 167, 216

    • Eumicrotremus orbis 162

    • Eupercaria 9, 11, 106, 112, 113, 152–153, 177, 200, 215

    • Eurypharyngidae 35, 36, 207

    • Eurypharynx 35

    • Eurypharynx pelecanoides 32

    • Eurypterygii 85

    • Euteleostei 3, 11, 23, 42, 43, 44, 46, 75–77, 77, 79, 81, 84, 87, 185, 195–196, 210

    • Euteleosteomorpha 76

    • †Evenkia eunoptera 186

    • Eventognathi 58

    • Evermannellidae 88, 89, 211

    • †Exallias vectensis 140

    • Exocoetidae 133, 134, 135, 214

    • Exocoetoidei 135

    • †Ezkutuberezi carmeni 193

    • Fangfangia 60, 62

    • †Farinichthys 28

    • †Farinichthys gigas 191

    • Fierasfer 31

    • Fistularia 126, 127

    • Fistulariidae 127, 213

    • Fluviphylacidae 137, 214

    • Fluviphylax 136

    • †Francolebias 135, 137

    • †Francolebias aymardi 136, 202

    • Fundulidae 136, 137, 214

    • †Fuyuanichthys 22, 23

    • †Fuyuanichthys wangi 189

    • Gadariae 102

    • Gadidae 103, 104, 212

    • Gadiformes 9, 11, 94, 95, 96, 97, 98, 100, 102, 104, 211

    • Gadoidei 9, 96, 102, 102–104, 212

    • Gadus morhua 92, 95, 102, 104

    • Gaidropsaridae 103, 104, 212

    • Galaxiidae 50, 76, 77, 78, 79, 85, 211

    • Gambusia affinis 128

    • Gasteropelecidae 65, 74, 75, 208

    • †Gasterorhamphosus 113, 126, 127

    • †Gasterorhamphosus zuppichinii 127, 201

    • Gasterosteales 164

    • Gasterosteidae 126, 155, 159, 160, 162, 163–165, 216

    • Gasterosteiformes 126, 127, 149, 164

    • Gasterosteoidei 146, 164

    • Gasterosteus 164, 165

    • Gasterosteus aculeatus 163

    • †Gasterosteus cf. aculeatus 165

    • Gastromyzontidae 60

    • †Gaudryella 42, 44

    • †Gaudryella gaudryi 193

    • †Gazolaichthys 177, 178, 179

    • †Gazolaichthys vestenanovae 205

    • Gempylidae 123, 125, 213

    • †Gephyroberyx robustus 108

    • Gerreidae 173, 176, 177, 218

    • Gerres cinereus 152

    • †Ghabouria 42, 44

    • †Gharbouria libanica 193

    • Gibberichthyidae 111, 212

    • Gibberichthys 110, 111

    • Gibbonsia metzi 137, 139

    • Gigantactinidae 179, 181, 182, 218

    • Gigantura 88

    • Giganturidae 89, 211

    • Gillellus semicinctus 127

    • †Gilmourella 126, 127

    • †Gilmourella minuta 127, 201

    • Girellidae 167, 169, 217

    • Glaucosoma 172, 176

    • Glaucosoma 172, 176

    • Glaucosomatidae 173, 217

    • Glossanodon musceli 78

    • Gobiaria 119

    • Gobiesocidae 120, 128, 138, 139, 214

    • Gobiesociformes 7, 127

    • Gobiesocoidei 96

    • Gobiesox maeandricus 127, 137

    • Gobiida 119

    • Gobiidae 118, 119, 120, 121, 213

    • Gobiiformes 106, 112, 113, 118–119, 121, 200, 212

    • Gobio gobio 59

    • Gobioidei 106, 118, 119, 119–121, 212

    • Gobiomorpharia 119

    • Gobionidae 61, 62, 209

    • Gobionotothen 157, 158, 216

    • Gobius niger 118, 119

    • Gobius niger 118, 118

    • Gonorhynchoidei 54

    • Gonorynchidae 54, 208

    • Gonorynchiformes 44, 45, 47, 51, 52, 52–54, 56, 194, 208

    • Gonorynchus 53, 54

    • Gonorynchus gonorynchus 51, 52, 53

    • Gonorynchus greyi 44, 52, 53

    • Gonostoma denudatum 82

    • Gonostomatidae 82, 83, 210

    • Goodeidae 136, 137, 214

    • †Gordichthys 53, 54

    • †Gordichthys conquensis 194

    • Gramma 128, 138, 161

    • Grammatidae 138, 139, 140, 214

    • Grammicolepididae 100, 101, 211

    • Grammonus 115

    • Grasseichthys gabonensis 54

    • †Guildayichthyidae 15, 16, 17, 186

    • †Guildayichthys carnegiei 16

    • Gvozdarus 157, 158, 216

    • Gymnarchidae 40, 208

    • Gymnarchus 39

    • Gymnarchus niloticus 25, 39, 40

    • Gymnodontes 184

    • Gymnonoti 64

    • Gymnothorax formosus 32

    • Gymnotidae 63, 64, 209

    • Gymnotiformes 45, 51, 52, 54, 55, 56, 62–64, 65, 185, 209

    • Gymnotoidae 64

    • Gymnotoidea 64

    • Gymnotoidei 51, 54, 64

    • Gymnotus carpio 51, 54

    • Gymnotus pantherinus 62, 63

    • Gyrinocheilidae 59, 209

    • Gyrinocheilus 56, 57, 58, 59, 60, 61

    • Gyrinocheilus pustulosus 56

    • †Habroichthys minimus 187

    • †Habroichthys 15, 16, 17

    • Haemulidae 175, 176, 177, 218

    • †Hajulia 30, 31

    • †Hajulia multidens 192

    • †Hakeliosomus 53, 54

    • †Hakeliosomus hakelensis 194

    • Halecostomi 22

    • Halobatrachus 116, 117, 118

    • Halobatrachus didactylus 117

    • Halophryne 117

    • Halophryninae 116, 117, 118

    • Halosauri 32

    • Halosauridae 31, 32, 208

    • Halosaurus ovenii 31

    • Hapalogenys 175

    • Haplodoci 117

    • Harpadon 89

    • Harpagiferidae 157, 158, 216

    • †Hayenchelys 28

    • †Hayenchelys germanus 192

    • †Helgolandichthys 42, 44

    • †Helgolandichthys schmidi 192

    • †Helmolepis cyphognathus 186

    • Helostoma temminckii 151, 152

    • Helostomatidae 152, 215

    • Hemerocoetidae 120, 172, 173, 217

    • Hemibranchii 164

    • Hemilutjanus macrophthalmos 155, 172

    • †Hemingwaya 147, 148

    • †Hemingwaya sarissa 147, 203

    • Hemiodontidae 74, 75, 208

    • Hemiramphidae 133, 134, 135, 214

    • Hemiramphus 134

    • Hemiramphus edwardsi 135

    • Hemiramphus far 133

    • Hepsetidae 74, 75, 208

    • Hepsetus 72, 74

    • Heptapteridae 71, 72, 210

    • Heterenchelyidae 36, 207

    • Heteroconger 37

    • Heteroconger hassi 36

    • Heterocongrinae 37

    • Heterognathi 74

    • Heteromi 31, 32

    • †Heteronectes 144, 145, 146

    • †Heteronectes chaneti 144, 145, 203

    • Heteropneustes 70, 71

    • Heteropneustidae 72, 210

    • Heterostomata 146

    • Heterotis 40, 41, 42

    • Heterotis niloticus 40, 41

    • Hexagrammidae 158, 159, 161, 162, 163, 216

    • Hexagrammos decagrammus 161

    • Himantolophidae 182, 218

    • Himantolophus 179, 181

    • Hiodon 25, 38, 40

    • Hiodon alosoides 40

    • Hiodon tergisus 23, 37, 40

    • Hiodontidae 39, 208

    • Hippopotamyrus 39

    • Hispidoberycidae 111, 212

    • Hispidoberyx 110

    • Hispidoberyx ambagiosus 110, 111

    • †Histionotophorus 182

    • Histiophrynidae 181, 182

    • Holocentridae 92, 105, 108, 109, 110, 111, 212

    • Holocentrus rufus 104, 107, 109

    • Holostei 3, 16, 18, 19, 20, 21, 21–23, 188–190, 207

    • †Holosteus 88, 89

    • †Holosteus esocinus 196

    • †Homonotichthys 98

    • Hoplichthyidae 161, 216

    • Hoplichthys 120, 159, 160, 161

    • Hoplichthys langsdorfii 160

    • Hoplostethus 105

    • Horabagridae 70, 71, 72, 210

    • Howellidae 172, 173, 217

    • †Howqualepis 10

    • †Huashia 38, 39

    • †Huashia gracilis 190

    • †Hulettia 20, 21

    • †Hulettia americana 187

    • †Humboldtichthys kirschbaumi 64

    • Hypopomidae 63, 64, 209

    • Hypoptychus 163, 164

    • Hypoptychus dybowskii 163, 164, 165

    • Hyporhamphus 134, 135, 214

    • †Hypsidoris 65, 67, 69

    • †Hypsidoris farsonensis 195

    • †Ichthyemidion 29

    • †Ichthyemidion vidali 191

    • Ichthyococcus 82, 83, 210

    • †Ichthyodectiformes 20, 21, 24, 188

    • †Ichthyokentema 20, 21

    • †Ichthyokentema purbeckensis 187

    • Icosteidae 125, 213

    • Icosteus aenigmaticus 122, 123, 125

    • Ictaluridae 70, 71, 72, 210

    • Iguanodectidae 65, 74, 75, 208

    • Ijimaia 90

    • Indostomidae 150, 215

    • Indostomus 126, 148, 149, 150

    • Indostomus paradoxus 148, 149

    • Inimicus didactylus 160

    • Iniomi 87, 88, 91

    • †Ionoscopus 22, 23

    • Ionoscopus cyprinoides 188

    • Ipnopidae 88, 89, 211

    • †Iraniplectus 183, 184, 185

    • †Iraniplectus bakhtiari 206

    • †Iridopristis 109, 110

    • Iridopristis parrisi 199

    • Iso 132, 133

    • Isonidae 133, 213

    • Isospondyli 47

    • †Istieus 30, 31

    • †Istieus grandis 192

    • Istiophoridae 123, 141, 143, 147, 148, 215

    • †Italoclupea 48, 49

    • †Italoclupea nolfi 193

    • Jaydia 122

    • †Jianghanichthys 56, 57, 58, 59

    • †Jianghanichthys hubeiensis 58, 195

    • †Jinanichthys 38, 39

    • †Jinanichthys longicephalus 190

    • †Jiuquanichthys 25, 27

    • †Jiuquanichthys liui 190

    • †Joffrichthys 41, 42

    • †Joffrichthys symmetropterus 190

    • Jordaniidae 162, 163, 216

    • †Judeichthys 53, 54

    • †Judeichthys haasi 194

    • †Judeoberyx 105, 107

    • †Judeoberyx princeps 199

    • †Jurgensenichthys 126, 127

    • †Jurgensenichthys elongatus 127, 202

    • Kaperangus microlepis 172

    • Kasatkia 166

    • †Kenyaichthys 135, 137

    • †Kenyaichthys kipkechi 136, 202

    • †Kermichthys 75, 76, 77

    • †Kermichthys daguini 195

    • Kneria 54

    • Kneria paucisquamata 52

    • Kneriidae 53, 54, 55, 208

    • †Knightia eocaena 49

    • Kryptoglanidae 71, 72, 210

    • Kryptoglanis shajii 71

    • Kuhlia 167, 169

    • Kuhlia marginata 167

    • Kuhliidae 169, 217

    • †Kuntulunia 38, 39

    • †Kuntulunia longipterus 190

    • Kurtidae 122, 212

    • Kurtiformes 122

    • Kurtus 118, 119, 120, 121, 122

    • Kurtus indicus 118, 121

    • †Kushlukia 177, 178, 179

    • †Kushlukia permira 205

    • Kyphosidae 167, 169, 217

    • Labidesthes sicculus 128

    • Labridae 170, 171, 173, 217

    • Labriformes 112, 153, 168, 170–171, 173, 185, 204, 217

    • Labrisomidae 139, 140, 141, 214

    • Labrisomus 161

    • †Labrobolcus 170, 171

    • †Labrobolcus giorgioi 171, 204

    • †Labrophagus 88, 89

    • †Labrophagus esocinus 196

    • Labrus 170, 173

    • Labrus bergylta 152, 170

    • Labrus mixtus 170, 173

    • Labyrinthici 149, 152

    • Labyrinthiformes 152

    • Lacantunia enigmatica 71

    • Lacantuniidae 71, 72, 210

    • Lactariidae 143, 215

    • Lactarius lactarius 143

    • †Laeliichthys 39, 40

    • †Laeliichthys ancestralis 40, 190

    • Lampridacea 95

    • Lampridae 95, 211

    • Lampridiformes 95

    • Lampriformes 86, 92, 93, 94–95, 97, 102, 107, 197, 211

    • Lamprimorpha 95

    • Lampripterygii 95

    • Lampris 94

    • Lampris guttatus 92, 94

    • Lamprologus callipterus 137

    • †Latellagnathus 131, 133

    • Lateolabracidae 173, 218

    • Lateolabrax 172

    • †Lateopisciculus 99

    • †Lateopisciculus turrifumosus 99, 198

    • Latidae 143, 215

    • Latridae 167, 169, 217

    • Lebiasinidae 72, 74, 75, 208

    • †Lebonichthys 28

    • †Lebonichthys namourensis 191

    • †Lecceclupea 48, 49

    • †Lecceclupea ehiravaensis 193

    • Leiognathidae 175, 176, 177, 218

    • Lepidoblennius 139

    • †Lepidocottus 119, 120, 121

    • †Lepidocottus aries 200

    • Lepidocybium 125, 213

    • Lepidocybium flavobrunneum 123, 125

    • Lepidogalaxias 85

    • Lepidogalaxias salamandroides 6, 42, 75, 76

    • Lepidogalaxii 6

    • Lepidogalaxiidae 6, 77, 210

    • Lepidogalaxiiformes 6

    • Lepidonectes corallicola 139

    • Lepidosteidae 15

    • †Lepidotes 22, 23

    • †Lepidotes semiserratus 189

    • Lepisosteidae 15, 19, 20, 21, 22, 23, 207

    • Lepisosteus osseus 10, 15, 19, 21

    • Leptobarbidae 62, 209

    • Leptobarbus 61, 62

    • Leptobrama 143, 147, 148

    • Leptobrama muelleri 146

    • Leptobramidae 148, 215

    • †Leptolepis 20, 21, 24

    • Leptolepis coryphaenoides 12, 188

    • †Leptolumamia 119, 122

    • Leptoscopidae 170, 171, 217

    • Lethrinidae 175, 176, 177, 218

    • Leuciscidae 61, 62, 209

    • Leuciscus leuciscus 60

    • †Libotonius 99

    • †Libotonius blakeburnensis 99, 198

    • Lichia 147

    • Lichia amia 147

    • †Lindoeichthys 99

    • †Lindoeichthys albertensis 99, 198

    • Linophrynidae 179, 181, 182, 218

    • Liparidae 158, 162, 163, 216

    • Lipogramma 128, 138

    • †Lissoberyx 105, 107

    • †Lissoberyx dayi 199

    • Lobotes 175

    • Lobotidae 175, 176, 177, 218

    • Lophichthyidae 182, 218

    • Lophichthys 179

    • Lophichthys boschmai 179, 181

    • Lophiidae 179, 181, 182, 218

    • Lophiiformes 9, 112, 176, 182

    • Lophioidei 4, 9, 96, 102, 107, 112, 115, 116, 175, 176, 177, 179, 180, 181–182, 184, 185, 205–206, 218

    • Lophius gastrophysus 179

    • Lophius piscatorius 179

    • Lophotidae 94, 95, 211

    • Loricaria cataphracta 65, 69

    • Loricaria simillima 69, 70

    • Loricariidae 69, 70, 209

    • Loricarioidea 69

    • Loricarioidei 65, 66, 67, 68, 69, 69–70, 185, 209

    • Lotidae 103, 104, 212

    • †Louckaichthys novosadi 117

    • Luciocephaloidei 152

    • †Luenchelys 28

    • †Luenchelys minimus 192

    • Lumpenidae 166, 167, 217

    • †Lusitanichthys 51, 52

    • †Lusitanichthys africanus 194

    • Lutjanidae 175, 176, 177, 218

    • Luvaridae 179, 218

    • Luvarus 125, 177, 178

    • Luvarus imperialis 177, 178

    • Lyconidae 103, 104, 212

    • Lyconus 103

    • †Lycoptera 25, 27

    • †Lycoptera davidi 190

    • Lyomeri 32, 35

    • Lythrypnus dalli 118, 119

    • †Macabi 30, 31

    • †Macabi tojolabalensis 191

    • †Macroaulostomus 126, 127

    • †Macroaulostomus veronensis 127, 202

    • Macroramphosidae 127, 213

    • Macrorhamphosus 126

    • †Macrosemius 22, 23

    • †Macrosemius rostratus 189

    • Macrouridae 103, 104, 212

    • Macrurocyttus 100

    • Macrurocyttus acanthopodus 100

    • Macruronidae 103, 104, 212

    • Macruronus 103

    • Macruronus novaezelandiae 102

    • Malacanthidae 176, 177, 218

    • Malacopterygii 47

    • Malakichthyidae 171, 172, 173, 218

    • Malakichthys 172

    • Malapteruridae 71, 72, 210

    • Mallotus 84

    • Marcusenius 39

    • †Masillosteus 22, 23

    • †Masillosteus janeae 189

    • †Massalongius 177, 178, 179

    • †Massalongius gazolai 205

    • †Massamorichthys 99

    • †Massamorichthys wilsoni 99, 198

    • Mastacembelidae 149, 150, 215

    • Mastacembelus mastacembelus 148, 149

    • †Mcconichthys 99

    • †Mcconichthys longipinnis 99, 198

    • Mccoskerichthys 140

    • †Megalampris 95

    • †Megalampris keyesi 94, 197

    • Megalomycteridae 111

    • Megalopidae 29, 207

    • Megalops 29

    • Megalops cyprinoides 28

    • Melamphaidae 110, 111, 212

    • Melanocetidae 182, 218

    • Melanocetus 179, 181

    • Melanonidae 104, 212

    • Melanonus 103

    • Melanotaeniidae 132, 133, 213

    • Melapedalion 135, 214

    • Melapedalion breve 134

    • Mene 148

    • Mene maculata 143, 147

    • †Mene purdyi 147, 148, 203

    • Menidae 148, 215

    • Menidia 161

    • Merlucciidae 103, 104, 212

    • Merluccius merluccius 102

    • †Mesogaster 133

    • Metavelifer multiradiatus 94

    • Microcanthidae 167, 169, 217

    • Microcyprini 137

    • Micropterus salmoides 85, 89, 92, 104, 111, 152, 167

    • Microstoma microstoma 77

    • Microstomatidae 77, 78, 210

    • Milyeringa veritas 119

    • Milyeringidae 120, 121, 213

    • †Mimia 10

    • Mirapinnidae 111

    • Mochokidae 71, 72, 210

    • †Moclaybalistes 183, 184, 185

    • †Moclaybalistes danekrus 206

    • Mola mola 182

    • Molidae 183, 184, 185, 219

    • Monacanthidae 183, 184, 185, 219

    • Monocentridae 108, 212

    • Monocirrhus polyacanthus 151

    • Monodactylidae 172, 177, 218

    • Monodactylus 172, 176

    • Monognathidae 35, 36, 207

    • Monognathus 32

    • Moridae 103, 104, 212

    • Moringua microchir 35

    • Moringuidae 33, 35, 36, 207

    • Mormyridae 25, 39, 40, 208

    • Mormyroidea 39

    • Moronidae 100, 173, 176, 177, 218

    • Moythomasia 10

    • Mugil cephalus 127

    • Mugilidae 112, 128, 138, 139, 214

    • Mugiliformes 7

    • Mullidae 126, 127, 213

    • Mullus auratus 126

    • Muraena helena 36

    • Muraenesocidae 37, 207

    • Muraeni 34

    • Muraenidae 36, 207

    • Muraenoidei 26, 33, 34, 36, 185, 207

    • Muraenolepididae 103, 104, 212

    • Myctophata 91

    • Myctophidae 89, 91, 92, 211

    • Myctophiformes 77, 85, 86, 87, 89, 90, 90–92, 96, 185, 197, 211

    • Myctophoidei 91

    • Myctophum punctatum 90

    • Myroconger 36

    • Myroconger compressus 32, 36

    • Myrocongridae 36, 207

    • Nandidae 151, 152, 215

    • Nandus 151

    • †Nardoclupea 48, 49

    • †Nardoclupea grandei 193

    • †Nardonoides 51, 52

    • †Nardonoides chardoni 195

    • †Nardovelifer 93

    • †Nardovelifer altipinnis 197

    • †Neilpeartia 182

    • Nemacheilidae 59, 60, 209

    • Nematistiidae 148, 215

    • Nematistius 147, 148

    • Nematistius pectoralis 143, 147, 148

    • Nematogenyidae 70, 209

    • Nematogenys 69

    • Nematogenys inermis 69

    • Nematognathi 68

    • Nemichthyidae 33, 35, 36, 207

    • Nemipteridae 175, 176, 177, 218

    • Nemoossis 27, 30, 31

    • Nemoossis belloci 30

    • †Neocassandra 89, 90

    • †Neocassandra mica 90, 197

    • Neoceratias 179

    • Neoceratias spinifer 179, 181

    • Neoceratiidae 182, 218

    • Neoclinini 140, 141, 214

    • Neoclinus 140

    • Neocyema 35

    • Neocyema erythrosoma 33

    • Neocyematidae 35, 36, 207

    • Neopterygii 11, 12, 16, 17, 19–21, 187–190, 207

    • Neoscopelidae 89, 91, 92, 211

    • Neoscopelus macrolepidotus 90

    • Neosebastidae 161, 216

    • Neoteleostei 11, 75, 76, 77, 78, 79, 85, 86, 87, 211

    • Neozoarcidae 166, 167, 217

    • Nessorhamphus 37

    • Nettastomatidae 33, 37, 207

    • †Nhanulepisosteus 22, 23

    • †Nhanulepisosteus mexicanus 189

    • Niphon 155, 156

    • Niphon spinosus 155, 156

    • Niphonidae 157, 216

    • Nomeidae 123, 125, 213

    • Normanichthyidae 161, 216

    • Normanichthys 160, 161

    • Normanichthys crockeri 158, 160, 161

    • Notacanthidae 31, 32, 208

    • Notacanthiformes 25, 26, 27, 28, 30, 31–32, 207

    • Notacanthus chemnitzii 31

    • Nothobranchiidae 135, 136, 137, 214

    • Notocheirus 132

    • Notocheirus hubbsi 132

    • †Notogoneus 53, 54

    • †Notogoneus montanensis 194

    • Notopteridae 25, 39, 40, 208

    • Notopterus 40

    • Notopterus notopterus 38, 39

    • Notosudidae 88, 89, 211

    • Notothenia 157

    • Notothenia coriiceps 157

    • Nototheniidae 157, 158, 216

    • Nototheniiformes 158

    • Notothenioidae 158

    • Notothenioidei 139, 154, 155, 156, 157–158, 216

    • Novumbra 80, 81

    • †Novumbra oregonensis 80

    • †Nunaneichthys 30, 31

    • †Nunaneichthys mexicanus 31, 192

    • †Nybelinoides 75, 76, 77

    • †Nybelinoides brevis 196

    • †Obaichthyidae 22, 23

    • †Obaichthys decoratus 189

    • Odacidae 170

    • Odontobutidae 120, 121, 213

    • Ogcocephalidae 179, 181, 182, 218

    • Ogcocephalus radiatus 179

    • †Oldmanesox 78, 79

    • †Oldmanesox canadensis 195

    • †Oligobothus 146

    • †Oligobothus pristinus 144, 145, 204

    • †Oligolophotes 95

    • †Oligolophotes fragosus 94, 197

    • †Oligopleuronectes 146

    • †Oligopleuronectes germanicus 144, 145, 204

    • Oligoplites 147

    • †Oligoscatophagus 173, 176, 177

    • †Oligoscatophagus capellini 205

    • †Omosoma 98

    • †Omosomopsis 97, 98

    • †Omosomopsis simum 98, 198

    • Oncopteridae 145, 146, 215

    • Oncopterus darwini 145

    • Oneirodidae 179, 181, 182, 218

    • Ophicephaliformes 152

    • Ophichthidae 37, 207

    • Ophichthys zophochir 36

    • Ophidiaria 115

    • Ophidiicae 115

    • Ophidiida 115

    • Ophidiidae 114, 115, 212

    • Ophidiiformes 96, 104, 106, 107, 111, 112, 113, 114–115, 115, 116, 118, 139, 155, 185, 200, 212

    • Ophidiimorpharia 115

    • Ophidion 114

    • Ophidion barbatum 114

    • Opisthocentridae 166, 167, 217

    • †Opisthomyzon 147, 148

    • †Opisthomyzon glaronensis 147, 203

    • Opisthoproctidae 77, 78, 210

    • Opistognathidae 128, 138, 139, 140, 214

    • Oplegnathidae 169, 217

    • Oplegnathus 167, 169

    • Opsanus 117

    • Opsanus tau 116

    • Oreosomatidae 100, 101, 211

    • Orestiidae 136, 137, 214

    • †Orrichthys 182

    • Oryzias latipes 128, 130

    • Oseanacephala 9, 11, 23, 24, 24–25, 26, 27, 185, 190–192, 207

    • Osmeridae 81, 84, 85, 211

    • Osmeriformes 43, 50, 76, 77, 78, 79, 81, 83–85, 85, 196, 211

    • Osmeroidea 84

    • Osmeroidei 84

    • †Osmeroides 28

    • †Osmeroides lewesiensis 191

    • Osmeromorpha 76

    • Osmerus 84

    • Osmerus eperlanus 83, 84

    • Osmerus mordax 81, 83, 85

    • Osphronemidae 151, 152, 215

    • Ostarioclupeomorpha 46

    • Ostariophysen 51, 52, 55, 56

    • Ostariophysi 3, 4, 11, 44, 45, 46, 50, 51, 51–52, 55, 56, 76, 77, 185, 194–195, 208

    • Osteichthyes 13

    • Osteoglossi 38

    • Osteoglossidae 25, 26, 39, 40, 40–42, 185, 190–191, 208

    • Osteoglossiformes 26, 38, 39, 39–40, 41, 56, 185, 190–191, 208

    • Osteoglossinae 41

    • Osteoglossoidei 38

    • Osteoglossomorpha 23, 25, 26, 27, 37–39, 40, 42, 44, 47, 76, 185, 190–191, 208

    • Osteoglossum 39, 41, 42

    • Osteoglossum bicirrhosum 24, 38, 39, 40

    • Ostorhinchus doederleini 118

    • Ostraciidae 183, 184, 185, 219

    • Ostracion cubicus 182

    • Ostracoberycidae 173, 218

    • Ostracoberyx 109, 172

    • Otocephala 11, 23, 42, 44, 44, 45, 46, 50, 51, 76, 77, 78, 193–195, 208

    • Otomorpha 46

    • Otophysi 11, 45, 51, 52, 54–56, 63, 65, 74, 185, 195, 208

    • Ovalentaria 9, 11, 106, 112, 113, 127–128, 129, 130, 130, 213

    • Ovalentariae 128

    • Oxudercidae 120, 121, 213

    • Oxyporhamphus 134

    • †Pachycormidae 20, 21, 187

    • †Pachycormus macropterus 187

    • †Padovathurus 177, 178, 179

    • †Padovathurus gaudryi 205

    • Paedocyprididae 62, 209

    • Paedocypris 56, 58, 60, 61, 62

    • Paedocypris progenetica 56, 60

    • †Palaeocentrotidae 95, 197

    • †Palaeocentrotus boeggildi 94, 197

    • †Palaeoesox 79, 80, 81

    • †Palaeoesox fritzschei 196

    • †Palaeogadus weltoni 104

    • †Palaeogobio zhongyuanensis 62

    • †Palaeonotopterus 39, 40

    • †Palaeonotopterus greenwoodi 39, 190

    • †Palaeopholis 166, 167

    • †Palaeopholis laevis 165, 166, 204

    • †Palaeorhynchus 147, 148

    • †Palaeorhynchus senectus 147, 203

    • †Paleodenticeps 47, 48

    • †Paleodenticeps tanganikae 193

    • †Paleopsephurus 17, 19

    • †Paleopsephurus wilsoni 188

    • †Paleoserranus 153

    • †Paleoserranus lakamhae 152, 153, 200

    • Pan–Siluriformes 45, 55, 195

    • Pan–Teleostei 12, 19, 20, 23, 24, 25, 187–188

    • Pangasiidae 71, 72, 210

    • Pantanodon 136

    • Pantanodon stuhlmanni 135

    • Pantanodontidae 136, 137, 214

    • Pantodon 25, 39, 41,

    • Pantodon buchholzi 38, 39, 40, 41

    • Pantodontidae 40, 208

    • †Panxianichthys 22, 23

    • Panxianichthys imparilis 188

    • Papyrocranus 40

    • Parabrotula plagiophthalmus 115

    • Parabrotulidae 115, 116

    • Paracanthomorphacea 98

    • Paracanthopterygii 11, 85, 92, 93, 95–98, 100, 107, 112, 114, 115, 116, 176, 198–199, 211

    • Paracanthurus hepatus 177

    • Paracanthus hepatus 152

    • Parachanna 151

    • †Parachanos 53, 54

    • Parachanos aethiopicus 194

    • †Paraelops 28

    • †Paraelops cearensis 191

    • †Paraeoliscus 126, 127

    • †Paraeoliscus robinetae 127, 201

    • †Paralabrus rossiae 171

    • †Paralates 118, 119

    • †Paralates chapelcorneri 200

    • Paralepididae 88, 89, 211

    • Paralichthodes 144, 145

    • Paralichthodes algoensis 144, 145

    • Paralichthodidae 146, 215

    • Paralichthyidae 145, 146, 215

    • †Paralycoptera 38, 39

    • †Paralycoptera wui 190

    • †Paramphisile 126, 127

    • †Paramphisile weileri 127, 201

    • †Paranguilla tigrina 35

    • Paranotothenia 157

    • †Paraophiodon 162, 163

    • †Paraophiodon nessovi 162, 204

    • Parapercis hexophtalma 170

    • Parascombrops 172

    • Parascorpididae 169, 217

    • Parascorpis typus 167, 169

    • Paraulopidae 89, 211

    • Paraulopus 88

    • †Paravinciguerria 81

    • †Paravinciguerria praecursor 81, 196

    • †Parawenzichthys 42, 44

    • †Parawenzichthys minor 193

    • Parazenidae 100, 101, 211

    • Parodontidae 74, 75, 209

    • Parona 147

    • Parona signata 147

    • †Pastorius 111, 112, 113

    • †Pastorius methenyi 199

    • †Pavlovichthys 88, 89

    • †Pavlovichthys mariae 196

    • Pediculati 182

    • Pegasidae 126, 127, 213

    • Pegasus volitans 126

    • †Peipiaosteidae 19

    • †Peipiaosteus 16, 17

    • †Peipiaosteus pani 186

    • Pelagia 125

    • Pelagiaria 125

    • †Peltopleuridae 15, 16, 17, 186

    • †Peltopleurus lissocephalus 186

    • Pempheridae 119, 122, 172, 173, 218

    • Pempheriformes 172

    • †Pempheris huddlestoni 172

    • Pempheris schomburgkii 171

    • Pentacerotidae 172, 173, 175, 218

    • †Pepemkay 105, 107

    • †Pepemkay maya 199

    • Perca fluviatilis 10, 15, 19, 21, 23, 24, 42, 75, 98, 111, 152, 153, 156, 173

    • Percalates 167, 169, 217

    • Percalates colonorum 167

    • Percichthyidae 167, 169, 172, 217

    • Percidae 155, 156, 157, 216

    • Perciformes 3, 5, 7, 42, 111, 112, 152, 153, 153, 154, 155–156, 156, 163, 173, 177, 185, 204, 215

    • Percilia 169

    • Perciliidae 169

    • Percoidei 154, 155, 156, 156–157, 161, 172, 216

    • Percomorpha 3, 4, 5, 7, 11, 92, 105, 106, 107, 109, 111–113, 114, 115, 116, 118, 119, 120, 123, 126, 128, 130, 138, 141, 143, 148, 150, 153, 155, 171, 172, 175, 176, 177, 182, 185, 199–206, 212

    • Percomorphacea 113

    • Percomorpharia 153

    • Percophidae 158, 216

    • Percophis brasiliensis 155, 157, 158

    • Percopsacea 99

    • Percopsaria 99

    • Percopsidae 99, 211

    • Percopsiformes 95, 96, 97, 98, 98–99, 100, 104, 185, 198, 211

    • Percopsis 98, 99

    • Percopsis omiscomaycus 92, 95, 98, 104

    • Periophthalmus barbarus 119

    • Perulibatrachus 117

    • Phallostethidae 132, 133, 213

    • †Phareodus 41, 42

    • †Phareodus testis 190

    • Pholidae 165, 166, 167, 217

    • Pholidichthyidae 139, 214

    • Pholidichthys 128, 138, 139

    • Pholidichthys leucotaenia 127

    • †Pholidophoridae 20, 21, 187

    • †Pholidophorus 24

    • Phosichthyidae 82, 83, 210

    • Phosichthys argenteus 82

    • Phractolaemus ansorgii 53, 54

    • Phreatobiidae 72, 210

    • Phreatobius 71

    • Phycidae 103, 104, 212

    • †Phyllopharyngodon longipinnis 171

    • Phytichthys 165

    • Phytichthys chirus 165

    • Pimelodidae 71, 72, 210

    • Pimelodus maculatus 70

    • Pinguipedidae 170, 171, 217

    • †Pinichthys 123, 125

    • †Pinichthys pulcher 125, 200

    • †Pirskenius 119, 120, 121

    • †Pirskenius radoni 200

    • Platycephalidae 158, 159, 160, 161, 216

    • Platycephalus indicus 158, 160

    • †Platycephalus parapercoides 156

    • †Platysiagidae 15, 16, 17, 186

    • †Platysomus superbus 13

    • Platytroctes apus 49

    • Platytroctidae 50, 51, 208

    • Plecoglossidae 85, 211

    • Plecoglossus 84, 85

    • Plecoglossus altivelis 84

    • Plectocretacicoidei 183, 184

    • Plectocretacius 183

    • Plectognathes 184

    • Plectognathi 184

    • Plectospondyli 55

    • Plectrogeniidae 159, 160, 161, 216

    • †Plesioberyx 109, 110

    • †Plesioberyx discoides 110

    • †Plesioberyx maximus 110, 199

    • †Plesiolycoptera 38, 39

    • †Plesiolycoptera daquingensis 190

    • Plesiopidae 128, 138, 139, 214

    • Pleuragramma antarcticum 157

    • Pleuragrammatinae 158, 216

    • Pleuronectes platessa 144

    • Pleuronectidae 144, 145, 146, 215

    • Pleuronectiformes 7, 146

    • Pleuronectoidei 4, 9, 141, 142, 143, 144–146, 185, 203–204, 215

    • Pleuronectoideo 146

    • Pleuronichthys cornutus 141, 144

    • †Pliodetes 22, 23

    • †Pliodetes nigeriensis 189

    • †Plioplarchus whitei 169

    • Plotosidae 71, 72, 210

    • Poeciliidae 136, 137, 214

    • Poecilopsettidae 144, 145, 146, 215

    • Pollichthys 83, 210

    • Pollichthys mauli 82

    • Pollimyrus 39

    • Polycentridae 128, 138, 139, 151, 214

    • Polycentropsis abbreviata 151

    • Polycentrus 151

    • Polycentrus schomburgkii 127

    • Polymetme 82, 83, 210

    • Polymixia 97, 98, 100, 109

    • Polymixia lowei 92

    • Polymixiidae 98, 211

    • Polymixiiformes 98

    • Polynemidae 143, 215

    • Polynemus melanochir 141

    • Polyodon spathula 17

    • Polyodontidae 17, 19, 207

    • Polyprion 172

    • Polyprionidae 173, 218

    • Polypteridae 10, 13, 13–15, 16, 22, 185, 207

    • Polypteriformes 15

    • Polypterini 15

    • Polypterus 13, 15

    • Polypterus bichir 10, 13

    • †Polyspinatus 97, 98

    • †Polyspinatus fluere 198

    • Pomacanthidae 175, 176, 177, 219

    • Pomacentridae 128, 138, 139, 140, 170, 171, 214

    • Pomatomidae 125, 172, 213

    • Pomatomus saltatrix 123, 125

    • Porichthyinae 117, 118

    • Porichthys 117

    • †Portoselvaggioclupea 48, 49

    • †Portoselvaggioclupea whiteheadi 193

    • Priacanthidae 176, 177, 219

    • †Priscosturion 17, 19

    • Priscosturion longipinnis 188

    • Pristigasteridae 47, 48, 49, 208

    • Pristolepis 151

    • †Proaracana 183, 184, 185

    • †Proaracana dubia 206

    • Procatopodidae 136, 137, 214

    • Prochilodontidae 72, 74, 75, 209

    • †Proeleginops 158

    • †Proeleginops grandeastmanorum 157, 157, 204

    • Profundulidae 136, 137, 214

    • †Prohalecites 20, 21

    • †Prohalecites porroi 187

    • Prokaryota 4

    • †Prolebias 135, 137

    • †Prolebias stenoura 136, 202

    • †Prosolenostomus 126, 127

    • †Prosolenostomus lessinii 127, 201

    • Protacanthopterygii 76, 79

    • Protanguilla palau 32, 34, 35

    • Protanguillidae 35, 207

    • †Protobalistum 183, 184, 185

    • †Protobalistum imperialis 206

    • †Protolophotus elami 94

    • †Protopsephurus 17, 19

    • †Protopsephurus liui 19, 188

    • †Protosiganus 173, 176, 177

    • †Protosiganus glaronensis 205

    • Prototroctes 84

    • †Protozeus 97, 98

    • †Protozeus kuehnei 199

    • †Protriacanthus 183

    • †Proumbra 79, 80, 81

    • †Proumbra irtyshensis 196

    • Psephurus gladius 10, 15, 17

    • Psettodes 144, 145

    • Psettodes erumei 141, 144

    • Psettodidae 146, 215

    • Pseudamia gelatinosa 121

    • Pseudaphritidae 158, 216

    • Pseudaphritis urvillii 157

    • Pseudochromidae 128, 138, 139, 140, 214

    • Pseudochromis fridmani 127

    • Pseudomugilidae 132, 133, 213

    • †Pseudopholidoctenus germanicus 187

    • Pseudopimelodidae 71, 72, 210

    • Pseudotrichonotidae 89, 211

    • Pseudotrichonotus 88, 89

    • Psilorhynchidae 62, 209

    • Psilorhynchus 61, 62

    • Psychrolutidae 162, 163, 216

    • Pteropsaron evolans 171

    • Pterothrissus 27, 30, 31

    • Pterothrissus gissu 30

    • †Pterygocephalus 126, 127

    • †Pterygocephalus paradoxus 127, 201

    • Ptilichthyidae 167, 217

    • Ptilichthys goodei 135, 166

    • †Pugliaclupea 48, 49

    • †Pugliaclupea nolardi 193

    • Pungitius 164, 165

    • †Pycnodontiformes 15, 16, 17, 186

    • †Pycnosteroides 93

    • †Pycnosteroides levispinosus 197

    • †Pyrenichthys 75, 76, 77

    • †Pyrenichthys jauzaci 195

    • Pythonichthys microphthalmus 36

    • Rachycentridae 148, 215

    • Rachycentron 147

    • Rachycentron canadum 147, 148

    • Radiicephalidae 95, 211

    • Radiicephalus 94

    • †Ramallichthys 53, 54

    • †Ramallichthys orientalis 194

    • Raniceps raninus 103

    • Ranicipitidae 104, 212

    • †Redfieldiidae 15, 16, 17, 186

    • †Redfieldius gracilis 186

    • Regalecidae 94, 95, 211

    • Regalecus russelii 94

    • Retropinna 83, 84

    • Retropinna semoni 83

    • Retropinnidae 84, 85, 211

    • Rhabdamia 122

    • †Rhamphexocoetus 131, 133, 135

    • †Rhamphexocoetus volans 134, 202

    • Rhamphichthyidae 63, 64, 209

    • Rhamphocottidae 162, 163, 216

    • †Rhamphognathus 133

    • †Rhamphosus 126, 127

    • †Rhamphosus rastrum 127, 201

    • †Rhinocephalus cretaceus 104

    • Rhombosoleidae 144, 145, 146, 215

    • Rhyacichthyidae 120, 121, 213

    • Rhyacichthys 120

    • Rhyacichthys aspro 119

    • Rhycheridae 181, 182

    • Rhynchorhamphus 134

    • Riekertia 117, 118

    • Riekertia ellisi 117

    • Rita 71

    • Ritidae 72, 210

    • Rivulidae 135, 136, 137, 214

    • Rondeletia 110, 111

    • Rondeletiidae 111, 212

    • †Rubiesichthys 53, 54

    • †Rubiesichthys gregalis 54, 194

    • †Ruffoichthys 173, 176, 177

    • †Ruffoichthys spinosus 205

    • Saccopharyngidae 35, 36, 207

    • Saccopharyngiformes 33, 35

    • Saccopharyngoidei 32, 33, 35

    • Saccopharynx 32

    • †Sakhalinia 162, 163

    • †Sakhalinia multispinata 162, 204

    • Salangidae 84, 85, 211

    • Salarias 161

    • Salmo salar 75, 78, 79

    • Salmonidae 47, 50, 76, 77, 78, 79, 210

    • Salmoniformes 9, 11, 43, 50, 75, 76, 77, 78–79, 185, 195–196, 210

    • Salmopercae 99

    • Samaridae 144, 145, 146, 215

    • Sanopus 117

    • †Santanaclupea 44, 46

    • †Santanaclupea silvasantosi 193

    • †Santanasalmo 42, 44

    • †Santanasalmo elegans 192

    • †Santanichthys 51, 52

    • †Santanichthys diasii 52, 195

    • Sarcopterygii 13

    • †Sardinoides 89, 90

    • †Sardinoides monasteri 90, 196

    • Sargocentron 105

    • Saurida 89

    • †Scanilepiformes 10, 13, 186

    • Scaphirhynchus 17

    • Scaridae 170

    • Scartella cristata 137

    • Scatophagidae 175, 176, 177, 219

    • Schilbeidae 70, 71, 72, 210

    • Schuettea 171, 172, 173, 218

    • Sciaenidae 176, 177, 219

    • Scleropages 39, 41, 42

    • Scleroparei 160

    • Scoloplacidae 70, 209

    • Scoloplax 69

    • Scomber scombrus 122, 123

    • Scomberesocidae 133, 134

    • Scomberesox 134

    • Scomberoides 147

    • Scomberoidinae 147

    • Scombridae 87, 123, 125, 213

    • Scombriformes 109, 112, 113, 122–125, 200–201, 213

    • †Scombroclupeoides 42, 44

    • †Scombroclupeoides scutata 192

    • Scombroidei 123, 125, 141

    • Scombrolabracidae 125, 213

    • Scombrolabrax heterolepis 123, 125

    • Scombropidae 173, 218

    • Scombrops 172

    • Scopelarchidae 88, 89, 211

    • Scopelengys tristis 85, 89

    • Scopeliformes 88

    • Scopelomorpha 91

    • Scophthalmidae 144, 145, 146, 215

    • Scorpaena porcus 158, 160

    • Scorpaenicae 161

    • Scorpaenichthyidae 163, 216

    • Scorpaenichthys marmoratus 162

    • Scorpaenidae 158, 160, 161, 216

    • Scorpaeniformes 155, 156, 158, 160, 161

    • Scorpaenoidae 161

    • Scorpaenoidea 154, 159, 160, 160–161, 216

    • Scorpaenoidei 154, 155, 156, 158–160, 161, 162, 216

    • Scorpididae 167, 169, 175, 217

    • Scytalina 165

    • Scytalina cerdale 165, 166

    • Scytalinidae 165

    • Searsiidae 50

    • Sebastes norvegicus 153, 158, 160

    • †Semionotus 22, 23

    • †Semionotus elegans 189

    • Serpenticobitidae 60

    • Serpenticobitis 59, 60

    • Serranidae 152, 155, 156, 161, 172, 176, 216

    • Serraniformes 156

    • Serranus scriba 98

    • Serrasalmidae 65, 74, 75, 209

    • Serrivomer beanii 35

    • Serrivomeridae 33, 35, 36, 207

    • †Sharfia 179, 181, 182

    • †Sharfia mirabilis 205

    • Siganidae 177, 219

    • †Siganopygaeus 173, 176, 177

    • †Siganopygaeus rarus 205

    • Siganus 175, 176, 177, 178

    • Sillaginidae 175, 176, 177, 219

    • Siluri 68

    • Siluridae 68, 71, 72, 210

    • Siluriformes 4, 51, 52, 54, 55, 56, 63, 65, 65, 66, 67–69, 69, 71, 185, 195, 209

    • Siluroidea 68

    • Siluroidei 51, 54, 65, 66, 67, 68, 69, 70–72, 185, 209

    • Silurus glanis 51, 54, 65, 70

    • Simenchelys parasitica 34

    • †Sinamia 22, 23

    • †Sinamia zdanskyi 189

    • †Singida 41, 42

    • †Singida jacksonoides 191

    • Sinipercidae 167, 169, 217

    • †Sinoglossus 41, 42

    • †Sinoglossus lushanensis 190

    • Siphamia 122

    • Sisoridae 71, 72, 210

    • †Slovenitriacanthus 183

    • Solea solea 144

    • Soleidae 145, 146, 215

    • †Solenorhynchus 126, 127

    • †Solenorhynchus elegans 127, 201

    • Solenostomidae 127, 213

    • Solenostomus 126, 127

    • †Solnhofenamia 22, 23

    • †Solnhofenamia elongata 189

    • †Spaniodon 75, 81

    • †Spaniodon latus 196

    • Sparidae 72, 175, 176, 177, 219

    • †Speirsaenigma 84, 85

    • †Speirsaenigma lindoei 84, 85, 196

    • †Sphenocephalidae 97, 98, 198

    • Sphyraena 143

    • Sphyraenidae 143, 215

    • †Spinacanthus 183, 184, 185

    • †Spinacanthus cuneiformis 206

    • Spinachia 164

    • Spinachia spinachia 164, 165

    • Spratelloides gracilis 48

    • Spratelloididae 48, 49, 208

    • Squamipinnes 175

    • Stathmonotus 140

    • Steindachneria argentea 103

    • Steindachneriidae 103, 104, 212

    • Stenopterygii 83

    • Stephanoberycidae 110, 111, 212

    • Stephanoberyciformes 108

    • Stephanoberycoidei 110, 111

    • Stereolepididae 173, 218

    • Stereolepis 171, 172

    • Stereolepis gigas 171

    • Sternoptychidae 82, 83, 210

    • Sternoptyx diaphana 82

    • Sternopygidae 63, 64, 209

    • Sternopygus macrurus 62

    • Stiassnyiformes 128

    • Stichaeidae 165, 166, 167, 217

    • Stichaeus punctatus 165

    • †Stichocentrus 109, 110

    • †Stichocentrus elegans 110

    • †Stichocentrus liratus 109, 110

    • †Stichocentrus spinulosus 110

    • Stomias boa 81, 82

    • Stomiatia 83

    • Stomiatiformes 83

    • Stomiatii 9, 11, 43, 76, 77, 81, 210

    • Stomiatoidea 83

    • Stomiatoidei 83

    • Stomiidae 82, 83, 210

    • Stomiiformes 43, 47, 50, 75, 76, 77, 81, 82–83, 85, 87, 91, 210

    • †Stompooria 75, 76, 77

    • †Stompooria rogersmithi 195

    • Stromateidae 123, 125

    • Stromateidae 123, 125, 213

    • Stromateoidei 123, 125

    • Stylephoridae 102, 212

    • Stylephorus chordatus 92, 94, 97, 102

    • Sudidae 89, 211

    • Sudis 88

    • Sundadanio 60, 61, 62

    • Sundadanionidae 62, 209

    • †Surlykus 75, 76, 77

    • †Surlykus longigracilis 195

    • Symphysanodontidae 172, 173, 218

    • Synagropidae 171, 172, 173, 218

    • Synagrops 172

    • Synanceiidae 156, 161, 216

    • Synaphobranchidae 34, 35, 207

    • Synaphobranchoidei 26, 33, 34, 34–35, 185, 207

    • Synaphobranchus kaupi 32, 34, 35

    • Synbranchidae 149, 150, 215

    • Synbranchiformes 7, 112, 113, 142, 148–149, 150, 185, 202–203, 215

    • Synbranchoidei 142, 148, 149, 149–150, 215

    • Synbranchus marmoratus 148, 149

    • Synentognathi 135

    • Syngnatharia 127

    • Syngnathidae 126, 127, 213

    • Syngnathiformes 112, 113, 124, 126–127, 159, 201–202, 213

    • Syngnathus acus 126, 133

    • †Synhypuralis 126, 127

    • †Synhypuralis banister 127, 202

    • Synodontidae 88, 89, 211

    • Synodus 89

    • Taeniamia 122

    • Taeniopaedia 25

    • Taeniosomi 94

    • Takifugu rubripes 182

    • Tanichthyidae 62, 209

    • Tanichthys 61, 62

    • †Tarkus 179, 181, 182

    • †Tarkus squirei 206

    • Tarumania walkerae 72, 75

    • Tathicarpidae 181, 182

    • †Tchernovichthys 42, 44

    • †Tchernovichthys exspectatum 44, 192

    • Teleocephala 24

    • Teleostei 3, 11, 12, 16, 19, 20, 21, 22, 23–24, 28, 30, 36, 50, 68, 76, 83, 185, 188, 207

    • Telmatherinidae 132, 133, 213

    • †Tenuicentrum 109, 110

    • Tenuicentrum lanceolatum 199

    • Terapontidae 167, 169, 217

    • Terapontoidei 167

    • Tetragonuridae 123, 125, 213

    • Tetragonurus 123, 125

    • Tetraodon lineatus 182, 183

    • Tetraodontidae 183, 184, 185, 219

    • Tetraodontiformes 7, 9, 176, 184,

    • Tetraodontoidei 4, 9, 100, 175, 176, 177, 180, 182–185, 206, 219

    • †Thaiichthys 22, 23

    • †Thaiichthys buddhabutrensis 189

    • Thalasseleotrididae 120, 121, 213

    • Thalassophryne 117

    • Thalassophryninae 117, 118

    • †Tharrhias 53, 54

    • †Tharrhias araripis 194

    • †Tharsis 20, 21

    • Tharsis dubius 188

    • Thaumatichthyidae 179, 181, 182, 218

    • †Thoracopteridae 15, 16, 17, 187

    • †Thoracopterus wushaensis 187

    • †Thrissops subovatus 188

    • †Ticinolepis 22, 23

    • †Ticinolepis longaeva 189

    • Tinca 62

    • Tinca tinca 61

    • Tincidae 62, 209

    • †Tischlingerichthys 44, 46

    • †Tischlingerichthys viohli 194

    • Toxotes jaculatrix 146

    • Toxotidae 147, 148, 175, 215

    • Trachichthyidae 107, 108, 111, 212

    • Trachichthyiformes 100, 104, 105, 106, 107, 107–108, 109, 111, 112, 113, 212

    • Trachichthyoidei 108

    • Trachichthys australis 107

    • Trachinidae 155, 156, 157, 216

    • Trachiniformes 155, 172

    • Trachinocephalus 89

    • Trachinoidei 155, 158, 172

    • Trachinotidae 147, 148, 215

    • Trachinotinae 147

    • Trachinotus 147

    • †Trachinus falcatus 157

    • Trachinus radiatus 156

    • Trachipteridae 94, 95, 211

    • Trachypoma macracanthus 153

    • Trachyrincidae 103, 104, 212

    • Trematominae 157, 158, 216

    • Triacanthidae 183, 184, 185, 219

    • Triacanthodidae 183, 184, 185, 219

    • Triathalassothia 117, 118

    • Trichiuridae 123, 125, 213

    • Trichiurus lepturus 122

    • Trichodontidae 158, 162, 163, 216

    • Trichomycteridae 69, 70, 209

    • Trichomycterus guianense 69

    • Trichonotus 118, 119, 120

    • Trichonotus filamentosus 118

    • †Tricophanes 99

    • †Tricophanes foliarum 99, 198

    • Triglidae 159, 160, 161, 216

    • Triodon 183, 184

    • Triodon macropterus 183, 184

    • Triodontidae 185, 219

    • Triportheidae 65, 74, 75, 209

    • Tripterygiidae 139, 140, 141, 214

    • †Trollichthys bolcensis 49

    • †Tselfatia formosa 188

    • †Tselfatiiformes 23, 24, 188

    • †Turkmene finitimus 94, 95, 197

    • †Turkmenidae 95, 197

    • Ulvaria subbifurcata 165

    • Umbra 80, 81

    • Umbra krameri 79

    • Umbridae 76, 79, 80

    • Uranoscopidae 139, 170, 171, 217

    • †Urenchelys 28

    • †Urenchelys germanum 192

    • †Urosphen 126, 127

    • †Urosphen dubius 127, 201

    • Vaillantella 59

    • Vaillantellidae 60, 209

    • Valencia 136

    • Valenciidae 137, 214

    • †Varasichthyidae 20, 21, 24, 188

    • †Varasichthys ariasi 188

    • Veliferidae 94, 95, 211

    • †Venusichthys 15, 16, 17

    • †Venusichthys comptus 187

    • Verilus 172

    • †Veronavelifer sorbini 94

    • †Vidalamia 22, 23

    • †Vidalamia catalunica 189

    • Vinciguerria 82, 83, 210

    • Vinciguerria nimbaria 82

    • Vladichthys 117

    • †Watsonulus 21, 22, 23

    • †Watsonulus eugnathoides 21, 22, 188

    • †Wenzichthys 42, 44

    • †Wenzichthys congolensis 192

    • Woodsia 82

    • Xenentodon cancila 133

    • Xenisthmidae 120, 121, 213

    • Xenisthmus 120

    • Xenoconger fryeri 36

    • Xenocyprididae 61, 62, 209

    • Xenodermichthys copei 49

    • Xenomystus 40

    • †Xenyllion zonensis 198

    • Xiphias gladius 123, 141, 143, 146, 147

    • Xiphiidae 148, 215

    • †Xiphiorhynchus 147, 148

    • †Xiphiorhynchus parvus 147, 203

    • Xiphister 165

    • Xiphisterinae 166

    • †Yanbiania 38, 39

    • †Yanbiania wangqingica 190

    • Yarrella 82, 83, 210

    • Zanclidae 179, 218

    • Zanclus 177, 178

    • Zanclus cornutus 177, 178

    • Zaniolepididae 159, 162, 163, 216

    • Zaniolepis latipinnis 161

    • †Zaprora koreana 166

    • Zaprora silenus 166

    • Zaproridae 167, 217

    • Zeacea 101

    • Zeiariae 101

    • Zeidae 100, 101, 211

    • Zeiformes 96, 97, 98, 100–101, 102, 105, 107, 184, 199, 211

    • Zenarchopteridae 134, 135, 214

    • Zenion 100

    • Zeniontidae 100, 101, 211

    • Zenopsis 101

    • Zeoidei 101

    • Zeomorphi 101

    • Zeorhombiformes 146

    • Zeus 101

    • Zeus faber 92, 100

    • †Zignoichthys 183, 184, 185

    • †Zignoichthys oblongus 206

    • Zoarcales 166

    • Zoarceoidea 166

    • Zoarces elongatus 158

    • Zoarces viviparus 165

    • Zoarcicae 166

    • Zoarcidae 166, 167, 217

    • Zoarcoidea 9, 154, 155, 159, 160, 162, 165–167, 216

    • Zoarcoidei 96, 115, 139, 166

    • Zoqueichthys 93

    • Zoqueichthys carolinae 197

    • Zoroteleostei 75, 76

    • Zorzinilabrus furcatus 171

    Thomas J. Near and Christine E. Thacker "Phylogenetic Classification of Living and Fossil Ray-Finned Fishes (Actinopterygii)," Bulletin of the Peabody Museum of Natural History 65(1), 3-302, (18 April 2024). https://doi.org/10.3374/014.065.0101
    Received: 11 April 2023; Accepted: 10 October 2023; Published: 18 April 2024
    KEYWORDS
    Acanthomorpha
    Euteleostei
    Holostei
    Ostariophysi
    Perciformes
    Percomorpha
    PhyloCode
    Back to Top