Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Molecular taxonomy confirms that the northeastern Atlantic and Mediterranean Sea harbor a single lancelet, Branchiostoma lanceolatum (Pallas, 1774) (Cephalochordata: Leptocardii: Branchiostomatidae)

  • Filomena Caccavale ,

    Contributed equally to this work with: Filomena Caccavale, David Osca

    Roles Conceptualization, Data curation, Methodology, Writing – review & editing

    Affiliation Department of Biology and Evolution of Marine Organisms, Stazione Zoologica Anton Dohrn Napoli, Naples, Italy

  • David Osca ,

    Contributed equally to this work with: Filomena Caccavale, David Osca

    Roles Conceptualization, Data curation, Methodology, Software, Writing – review & editing

    Affiliation Department of Integrated Marine Ecology, Stazione Zoologica Anton Dohrn Napoli, Naples, Italy

  • Salvatore D’Aniello ,

    Roles Conceptualization, Data curation, Funding acquisition, Methodology, Resources, Supervision, Writing – original draft, Writing – review & editing

    salvatore.daniello@szn.it

    Affiliation Department of Biology and Evolution of Marine Organisms, Stazione Zoologica Anton Dohrn Napoli, Naples, Italy

  • Fabio Crocetta

    Roles Conceptualization, Data curation, Funding acquisition, Methodology, Resources, Supervision, Writing – original draft, Writing – review & editing

    Affiliation Department of Integrated Marine Ecology, Stazione Zoologica Anton Dohrn Napoli, Naples, Italy

Abstract

Branchiostomatidae (lancelets or amphioxus) comprises about 30 species, several of which are well-established models in evolutionary development. Our zoological and ecological knowledge of the family is nonetheless limited. Despite evident differences can be found among known populations, the taxonomy of Branchiostoma lanceolatum (type species of the genus Branchiostoma) has never been investigated with modern methods through its range in the northeastern Atlantic and Mediterranean Sea. We address this via a multilocus molecular approach and comparing specimens collected from different European populations. Results obtained here confirm the presence of a single species inhabiting the range between the topotypical localities of B. lanceolatum (Atlantic Ocean) and of its junior synonym B. lubricum (Mediterranean Sea), without evincing geographical structure between populations. This suggests that environment most likely drives the characteristics observed in different geographic areas. The long larval phase and the slow mutation rate in lancelets may have played a key role in the evolutionary history of this iconic species.

Introduction

The family Branchiostomatidae Bonaparte, 1846 (subphylum Cephalochordata) comprises about thirty species, known as lancelets or amphioxus [14]. They inhabit the soft bottoms of various sublittoral and coastal habitats (estuaries, coastal lagoons, river deltas, and open coasts) from temperate to tropical regions [57] and some species grow up to 10 cm in length. Lancelets are generally benthic, living half-buried and only exposing the rostral end to the water. They filter plankton through the gill-bars by generating a ciliary water current, entering from the mouth through the buccal cirri. Food particles are embedded in mucus, collected in the pharynx and then passed into the intestine [610].

Until recently, lancelets were divided in the genera Branchiostoma Costa, 1834 (the most diverse genus, exceeding twenty species) and Epigonichthys Peters, 1876 [1, 11]. However, subsequent studies reinstated the genus Asymmetron Andrews, 1893; these allocated there some taxa previously ascribed to Epigonichthys and investigated the phylogenetic relationships between the three clades, suggesting that Asymmetron diverged first and that Epigonichthys and Branchiostoma are sister groups [2, 3, 1214].

Notwithstanding morphological differences between the three genera in gonads organization, metapleural fold, and caudal process [1, 2, 14], they share the same adult morphology, a translucent and elongated body with well visible neural tube, a notochord, an endostyle, a segmented musculature and a postanal tail [1517]. Despite sharing these features with vertebrates, lancelets lack key vertebrate structures, such as migratory neural crest cells, a highly regionalized brain, or paired sense organs [18].

The morphological similarity of lancelets to vertebrates and their phylogenetic relatedness has attracted the scientific attention of biologists for centuries [19]. In particular, the European Branchiostoma lanceolatum [20], the East Asian Branchiostoma belcheri [21], and the Floridian–Caribbean Branchiostoma floridae Hubbs, 1922, have become established model organisms for the evolution of the developmental mechanisms (Evo-Devo) during the transition from invertebrate to vertebrate chordates [16, 2225]. Moreover, the lancelet genome resembles that of the chordate ancestor in terms of conserved organization, regulation, and function [2628].

Despite the general importance of Branchiostomatidae, little is known about much of this family. Several species were newly described, or their taxonomy has been clarified only recently [2, 29, 30], and species misidentification or cryptic diversity have been found using molecular approaches or integrative taxonomy [14, 3136]. Moreover, new records of lancelet larvae or adults improved our assessment of species-specific geographical distribution and ecological traits at a range of scales [3746].

Finally, even widely studied lancelet species still lack rigorous characterization. As an example, the taxonomy and phylogeography of B. lanceolatum, the type species of its genus (see [47]), has never been investigated with modern approaches through its range in the northeastern Atlantic–Mediterranean Sea. Yet, populations differ in size and morphology (Atlantic specimens are larger), developmental rate (Atlantic larvae grow slower), and spawning period (of longer duration in the Mediterranean Sea) [4851]. We addressed this using a multilocus molecular approach to compare B. lanceolatum specimens collected from diverse European populations by both Atlantic and Mediterranean coasts.

Material and methods

Sampling

Lancelet specimens were collected between 2012 and 2017 in five European localities (two from the Atlantic Ocean, two from the western Mediterranean Sea, and one from the central-eastern Mediterranean Sea) and including populations widely exploited for Evo-Devo studies [28, 5256]. Noteworthy, specimens were also sampled from near the type localities of Limax lanceolatus Pallas, 1774 (Cornwall: see [20]) and Branchiostoma lubricum Costa, 1834 (Naples, Italy: see [47]), the only confirmed subjective junior synonym of B. lanceolatum. Sampled localities are summarized in Table 1 and shown in Fig 1. Voucher specimens were fixed in 70–100% ethanol for molecular analyses, shortly after collection.

thumbnail
Fig 1. Map of the sampling sites (codes correspond to the localities reported in Table 1) highlighting the type localities of Limax lanceolatus Pallas, 1774 (red square) and Branchiostoma lubricum Costa, 1834 (red circle).

https://doi.org/10.1371/journal.pone.0251358.g001

thumbnail
Table 1. Sampling sites (codes as in Fig 1) with geographic coordinates (WGS 84), environmental data, sampling gear, date, and legit.

https://doi.org/10.1371/journal.pone.0251358.t001

Ethical statement

The field study did not involve endangered or protected species. All animal procedures were in compliance with the European Union guidelines.

DNA extraction, amplification, and sequencing

Genomic DNA was extracted from adult lancelets as previously described [56]. Partial sequences of three mitochondrial genes widely used for taxonomic studies, namely cytochrome c oxidase subunit I (COX1), 12S ribosomal ribonucleic acid (12S rRNA), and 16S ribosomal ribonucleic acid (16S rRNA), were amplified from three specimens randomly selected from each sampling locality, using the following species-specific primers designed for this study using OligoEvaluator: Cox1_forward 5’-GATTCATAATATGCGTGCTAGC-3’ and Cox1_reverse 5’-CGGCTCCTATAGACAAAACG-3’; 12S_forward 5’-GGGTTACTGATGATACATGC-3’ and 12S_reverse 5’-CTACTATTGACTACACCCTG-3’; 16S_forward 5’-CGCCTGTTTAACAAAAACAT-3’ and 16S_reverse 5’-CGGTCTGTACTCAGATCACGTA-3’.

PCR was conducted in a 25 μl volume reaction, containing 5 μl of Green GoTaq Reaction Buffer (5×), 1.75 μl of MgCl2 (25 mM), 0.5 μl dNTP mix (10 mM each), 1 μl of template DNA (50–80 ng/μl), 2.5 μl of each primer (5 μM), 0.12 μl of GoTaq® DNA Polymerase (5 u/μl), and distilled water. Amplifications were performed according to the following conditions: initial denaturation at 95°C (5 min), followed by 30 cycles of denaturation at 95°C (30 sec), annealing at 55°C (30 sec), extension at 72°C (1 min), with a final extension at 72°C (5 min).

PCR products were examined on ethidium bromide-stained 1% agarose-TAE gels, and bands of the appropriate molecular weight were extracted from the gel and purified using the GFX PCR DNA and Gel Band Purification Kit (GE Healthcare Life Sciences). For each specimen, the three amplified gene fragments were cloned in pGEM®-T Easy Vector (Promega) and Sanger sequenced from both directions using the M13 forward and reverse primers (Promega). Sequencing was carried out with an Automated Capillary Electrophoresis Sequencer 3730 DNA Analyzer (Applied Biosystems) using the BigDye® Terminator v3.1 Cycle Sequencing Kit (Life Technologies).

Sequences and sequence alignment

Sequences obtained for the three gene fragments were compared with reference sequences from the NCBI nucleotide database using BLASTn [57], and assembled into single contigs for each specimen using Sequencher v.5.4 (Gene Codes Co.; Ann Arbor, MI, USA). Complete mitochondrial genomes from the genus Branchiostoma were acquired from GenBank [58], together with those of Epigonichthys maldivensis (NC_006465) as outgroup based on its sister relationship with Branchiostoma taxa [3, 14, 36].

To construct the data set, the amino acid sequences of the partial COX1 gene were aligned using Translator X [59] to better align based on peptide sequence, whereas orthologous nucleotide sequences of the ribosomal RNA mitochondrial genes were aligned separately using MAFFT [60]. The nucleotide alignments from the three gene fragments (COX1, 12S, and 16S) were then concatenated. Alignment format conversions were performed using the ALTER webserver [61].

Phylogenetic analyses

We conducted phylogenetic analyses on the complete aligned and concatenated multilocus data sets, using two optimality criteria: Maximum likelihood estimation (ML) [62] and Bayesian inference (BI) [63]. ML analyses were conducted with RAxML v.8.1.16 [64] using the rapid hill-climbing algorithm.

BI analyses were conducted using MrBayes v.3.2.7a [65] on XSEDE through the on-line CIPRES Science Gateway v.3.3 [66]. We ran four simultaneous Markov chains for two million generations, sampling every thousand generations, and discarding the first quarter of generations, to prevent sampling before reaching stationarity (assessed by plots of log likelihood values and standard deviation of split frequencies). Two independent Bayesian inference runs were performed to assure adequate mixing of the Markov chains and detect convergence.

The best partition scheme and best-fit models of substitution for the data set were identified with Partition Finder 2 [67], applying the Akaike information criterion [68]. The partitions tested were all genes combined; all genes separated; and genes grouped by subunits—i.e. cox, ribosomal. Support for internal branches was evaluated by non-parametric bootstrapping [69] with a thousand replicates (ML) and by Bayesian posterior probabilities (BI). Maximal, high, moderate, and poor support for nodes was defined for ML as 100%, >70%, 50–70%, and below 50%, and for BI as 1, >0.95, 0.90–0.95%, 0.90, respectively.

Genetic distances

We evaluated genetic distances to assess species boundaries. Genetic distances (shown as the percentage difference in base substitutions per sites) were computed using the Kimura 2-Parameters (K2P) model [70] with a thousand bootstrap resampling using MEGA X [71], treating the concatenated dataset (COX1, 12S, and 16S) as a single locus.

Species delimitation analyses

Three different approaches were used for species delimitation: the Bayesian implementation of the Poisson Tree Processes model (bPTP) [72], the Automatic Barcode Gap Discovery (ABGD) method [73], and a statistical parsimony network analysis (TCS) [74].

The bPTP is a phylogeny-based species delimitation method that delimits species based on single-locus molecular data [72]. Species delimitation was thus analyzed using the entire mitochondrial dataset (COX1+12S+16S) as a single-locus data [75], and the bPTP was run to compare the number of species delimited by each model. The PTP model uses non-ultrametric trees to enumerate species in terms of the number of substitutions, which indicates branch length. The phylogenetic trees obtained using BI and ML provided inputs for comparison.

The bPTP analysis was performed in the Exelixis Lab species delimitation web server (http://species.h-its.org), as follows: the number of MCMC generations was 105, as recommended for small trees, thinning was set to 100 and we discarded the first quarter of samples. Convergence of the parameters was checked after the run.

The ABGD method detects the so-called ‘barcode gap’ in the distribution of pairwise distances [73, 76]. A distance value corresponding to the most reliable gap was used to group the sequences in MOTUs (Molecular Operational Taxonomic Units) [77]. The concatenated dataset alignment was processed using the ABGD program (https://bioinfo.mnhn.fr/abi/public/abgd/abgdweb.html), excluding the outgroup E. maldivensis. We applied default parameters with the Kimura two-parameter (K2P) model [70] and the following settings: a prior constraint on the maximum value of intraspecific divergence between 0.001 and 0.1, 10 recursive steps within the primary partitions defined by the first estimated gap, and a gap width of 1.5.

The statistical parsimony network analysis calculates the maximum number of mutational steps constituting a parsimonious connection between two haplotypes [74, 78]. The haplotypes are then joined into networks as per Templeton and colleagues [79], and those separated by more mutational steps (i.e. probability of secondary mutations exceeding 5%) remain disconnected. A statistical parsimony network analysis implemented in the TCS program [80] was applied to the complete Branchiostoma dataset to differentiate species in a mixed sample [81, 82]. Moreover, two complementary analyses (one with the concatenated fragments and one with the COX1 fragment only) were also carried out on the B. lanceolatum dataset to identify any geographic structure within the haplotypes found.

Results

Partial sequences of the three mitochondrial genes [COX1 (621 bp), 12S (488 bp), and 16S (501 bp)] were obtained from three specimens from each of the five different localities (Fig 1, Table 1, S1 Table). These were deposited in GenBank and accession numbers are reported in S2 Table. Our GenBank data extraction yielded complete mitochondrial genomes of 48 Branchiostoma specimens, belonging to four different taxa, as well as that of E. maldivensis (S1 Table).

We acquired fragments of all the three genes from all complete mitochondrial genomes derived from GenBank, and generated a concatenated matrix including all specimens. After alignment, the sequence data used for analyses consisted of 1653 characters (S1 File). The best partition scheme for the data set was to treat the three concatenated loci separately. The best-fit model of substitution for all was the GTR+G model. ML (-ln L = 7718.59) and BI (-ln L = 8331.61 and -ln L = 8341.92 for the respective runs) analyses reached a similar tree topology, with four well-defined terminal clades gaining maximal support (Fig 2).

thumbnail
Fig 2. ML and BI consensus tree based on the analyses of the concatenated mitochondrial dataset (COX1+12S+16S).

Values at nodes represent ML bootstrap support and Bayesian posterior probabilities, respectively. Colored bars indicate the results of the species delimitation analyses: bPTP with ML tree in red, bPTP with Bayesian tree in blue, ABGD in green, and TCS in grey. Novel sequence samples are highlighted in bold (see S1 Table for codes).

https://doi.org/10.1371/journal.pone.0251358.g002

Relationships between Branchiostoma clades were unclear, showing maximal support for the BI analysis, but moderate support for the ML analysis (Fig 2). Moreover, our results placed B. floridae as the sister species of a group composed by the remaining species, with B. lanceolatum diverging first, and B. belcheri and B. japonicum as sister species. Such data are consistent with a previous work by [83] who used protein-coding genes of the mitochondrial DNA, although they differ from other studies using phylogenomic data [84], complete mitochondrial genomes [33], and 12S rRNA [83]. Phylogenetic relationships based on complete mitochondrial genomes are presumably more reliable than those obtained here; however, this is out of the scope of the present work.

At a species level, two of the four terminal clades were monophyletic (those representing B. floridae and B. lanceolatum) and the clade representing B. lanceolatum included all our experimentally-derived specimens studied here. The two remaining clades comprised (i) a single specimen of B. belcheri (AY932825: [83]) and (ii) several specimens described as B. belcheri clustering with a single specimen of B. japonicum (NC_008069, derived from DQ407722: [83]), respectively (Fig 2; S1 Fig). This first clade (i) corresponds to B. belcheri, whereas the latter (ii) corresponds to B. japonicum, which was considered a junior synonym or a subspecies of B. belcheri until recently [13, 83].

The species delimitation approaches arrived at similar results and confirmed the topology of the phylogenetic analyses, yielding four well-defined MOTUs (Fig 2; S1 Fig). The TCS analysis defined a total of 61 haplotypes in the concatenated dataset. Two B. lanceolatum (AB194383 and AB478572) and two B. floridae (AB478581 and AB478582) specimens, respectively, had identical sequences in all three fragments.

All the interspecific genetic distances were over 20% (Table 2). The lower genetic distance was found between B. lanceolatum and B. belcheri (mean 20.6%, SEM ± 1.2%), while the highest was between B. floridae and B. japonicum (mean 25.9%, SEM ± 1.4%).

thumbnail
Table 2. Average of the uncorrected pairwise genetic distances between Branchiostoma species based on the concatenated dataset (COX1, 12S, and 16S).

https://doi.org/10.1371/journal.pone.0251358.t002

The TCS analysis on the concatenated B. lanceolatum dataset yielded a total of 25 different haplotypes found in the 26 specimens analyzed (AB194383 and AB478572 were identical, see above), but found no clear geographic structure (Fig 3). A similar result was also obtained when analyzing the COX1 fragment only, with 17 haplotypes and the following samples sharing haplotypes: (i) AB194383, AB478567, AB478572, AB478573, and BFA3; (ii) BNA2, BFA1, and BRO3; (iii) AB478568, AB478570, and BSI3; (iv) AB478571 and BAR1 (S2 Fig). In both cases, the Mediterranean coast of France had the most haplotypes (13 in the concatenated dataset and 9 in the COX1 matrix), possibly explained by this region contributing the most samples.

thumbnail
Fig 3. TCS for the concatenated matrix (COX1+12S+16S) of B. lanceolatum, showing relationships between the recorded haplotypes.

See S1 Table for codes. Abbreviations used: ATL–Atlantic Ocean; MED–Mediterranean Sea; AR–Argelès sur Mer; NA–Napoli; SI–Siracusa. Circles representing haplotypes are scaled to their frequencies. Black dots represent missing intermediate haplotypes. Branch length connecting the different haplotypes is proportional to the number of mutations, with small transversal lines along the connecting branches representing mutational steps.

https://doi.org/10.1371/journal.pone.0251358.g003

Discussion

The Mediterranean marine biogeography is mostly a subset of that of the Atlantic, having originated with the re-establishment of the Atlantic–Mediterranean connection (5.33 million years ago) [85, 86], a phenomenon that would suggest conspecificity between the biotas of the two seas. Indeed, several species possess an Atlantic–Mediterranean distribution [8790]. However, phylogeographical barriers between the Atlantic Ocean and the Mediterranean Sea and the geographical complexity of the Mediterranean have given rise to endemism, and cryptic and vicariant species, even among conspicuous species and model systems [82, 9194].

Speciation remains one of the most controversial and least understood topics in evolution due to the complexity and diversity of living organisms; it typically results from combining various intrinsic and extrinsic factors including geographic barriers to larval dispersal, duration of the larval phase, and reproductive isolation due to prezygotic and postzygotic mechanisms [9598].

Recent advances in the use of DNA barcoding and integrative taxonomy are now clarifying relationships between species and within ambiguous species-groups, including the biota of the northeastern Atlantic and the Mediterranean Sea [9294, 99, 100]. In fact, while the systematics of the Atlantic–Mediterranean biota dates back centuries [85, 86, 101, 102], we need to test historical and morphological taxonomy with modern research approaches. This is particularly important when differences among species can be subtle, as exemplified by Branchiostomatidae. Lancelets are typical examples of morphological and evolutionary stasis, exhibiting conserved morphology, genetic machinery, and development regulation, even in species that diverged million of years ago [103, 104].

Despite we tested here B. lanceolatum specimens from throughout its range in the northeastern Atlantic and the Mediterranean, none of the analyses done was able to discern them at the species level, which implies that they constitute a single species. This suggests that environment most likely drives the peculiar characteristics observed in certain populations. Indeed in diverse taxonomic groups, the Atlantic specimens outgrow their Mediterranean counterparts [105], whereas developmental rates and spawning periods may differ due to the discrepant mean temperatures between the two seas. Finally, the TCS analysis also failed to identify a clear geographic structure; while this method warrants further testing with more samples, it is indeed in agreement with studies of other lancelet populations worldwide [106, 107].

The absence of speciation at the Atlantic–Mediterranean boundary may be explained by (among other factors) the long-duration of the larval phase before settlement and the sluggish mutation rate of this clade. In fact, the planktonic larvae of B. lanceolatum dwell in the plankton until metamorphosis [48, 51], and the duration of the larval phase of lancelets is long, varying from one to three months, depending on species [9, 108110]. This species may therefore disperse widely, enhancing genetic connectivity among distant populations, and thus diminishing population structure. Moreover, cephalochordates generally exhibit a slow mutation rate, with diverse Branchiostoma species barely differing even within their complete mitochondrial sequences [111113].

Notwithstanding present results, this work refines our taxonomic and phylogeographic understanding of an iconic and important model species and, more generally, of the Atlantic–Mediterranean biota. Given that B. lanceolatum is the most widely-distributed species in the genus, with historical records from other biogeographic areas than the Atlantic–Mediterranean, including the Red Sea and the Indo–Pacific region (review in [1]), further work ought to explore the taxonomy of this species in A Phylogenomic Framework and Divergence History of Cephalochordata Amphioxus Framework global context.

Supporting information

S1 Table. GenBank identification numbers for Branchiostoma and Epigonichthys maldivensis sequences used in the present analyses and associated specimen data (localities obtained from GenBank and/or relevant paper/s).

*Misidentifications for Branchiostoma japonicum (see [13, 83]). Codes as in Table 1, Figs 13 and S1 and S2 Figs. Abbreviations used (GenBank ID): CM—complete mitochondrial; COX1—cytochrome c oxidase subunit I; 12S - 12S ribosomal ribonucleic acid; 16S - 16S ribosomal ribonucleic acid.

https://doi.org/10.1371/journal.pone.0251358.s001

(PDF)

S1 Fig. TCS for the concatenated matrix (COX1+12S+16S) of Branchiostoma lanceolatum, showing relationships between the recorded haplotypes.

See S1 Table for codes. Circles representing haplotypes are scaled to their frequencies. Branch length connecting the different haplotypes is proportional to the number of mutations, with small transversal lines along the connecting branches representing mutational steps.

https://doi.org/10.1371/journal.pone.0251358.s003

(TIF)

S2 Fig. TCS for the COX1 matrix of Branchiostoma lanceolatum, showing relationships between the recorded haplotypes.

See S1 Table for codes. Abbreviations used: ATL–Atlantic Ocean; MED–Mediterranean Sea; NA–Napoli; SI–Siracusa. Circles representing haplotypes are scaled to their frequencies. Black dots represent missing intermediate haplotypes. Branch length connecting the different haplotypes is proportional to the number of mutations, with small transversal lines along the connecting branches representing mutational steps.

https://doi.org/10.1371/journal.pone.0251358.s004

(TIF)

S1 File. Alignment of the concatenated dataset (COX1+12S+16S).

https://doi.org/10.1371/journal.pone.0251358.s005

(FAS)

Acknowledgments

Agnes Boutet (France), Filipe Castro (Portugal), Hector Escriva (France), and Gianfranco Mazza (Italy) provided samples. Paula Carolina Rodríguez Flores (Spain) and Silvia Perea (Spain) advised on species delimitation analyses. Davide Caramiello (Italy) and Alberto Macina (Italy) maintained lancelet cultures at Stazione Zoologica Anton Dohrn. We thank John Kirwan for the critical reading of the manuscript.

References

  1. 1. Poss SG, Boschung HT. Lancelets (Cephalochordata: Branchiostomatidae): how many species are valid? Isr J Zool. 1996; 42: S13–S66.
  2. 2. Nishikawa T. A new deep-water lancelet (Cephalochordata) from off Cape Nomamisaki, SW Japan, with a proposal of the revised system recovering the genus Asymmetron. Zoolog Sci. 2004; 21: 1131–1136. pmid:15572865
  3. 3. Kon T, Nohara M, Yamanoue Y, Fujiwara Y, Nishida M, Nishikawa T. Phylogenetic position of a whale-fall lancelet (Cephalochordata) inferred from whole mitochondrial genome sequences. BMC Evol Biol. 2007; 7: 127–132. pmid:17663797
  4. 4. WoRMS (2020) Branchiostomatidae Bonaparte, 1846 [cited 2020-10-17]. Accessed at: http://www.marinespecies.org/aphia.php?p=taxdetails&id=196078.
  5. 5. Da Silva LF, Tavares M, Soares-Gomes A. Population structure of the lancelet Branchiostoma caribaeum (Cephalochordata: Branchiostomidae) in the Baía de Guanabara, Rio de Janeiro, southeastern Brazil. Rev Bras Zool. 2008; 25: 617–623.
  6. 6. Vergara M, Oliva ME, Riascos JM. Population dynamics of the amphioxus Branchiostoma elongatum from northern Chile. J Mar Biol Assoc U K. 2011; 92: 591–599.
  7. 7. Maghsoudlou A. Iranian Branchiostoma species (Cephalochordata, Branchiostomatidae) inhabiting Chabahar Bay (Gulf of Oman), with remarks on habitat preferences. Turk J Zool. 2018; 42: 179–186.
  8. 8. Laudien J, Rojo M, Oliva M, Arntz W, Thatje S. Sublittoral soft bottom communities and diversity of Mejillones Bay in northern Chile (Humboldt Current upwelling system). Helgol Mar Res. 2007; 61: 103–116.
  9. 9. Desdevises Y, Maillet V, Fuentes M, Escriva H. A snapshot of the population structure of Branchiostoma lanceolatum in the Racou Beach, France, during its spawning season. PLoS One. 2011; 6: e18520. pmid:21525973
  10. 10. Orton JH. The Ciliary Mechanisms on the Gill and the Mode of Feeding in Amphioxus, Ascidians, and Solenomya togata. J Mar Biol Assoc U K. 1913; 10: 19–49.
  11. 11. Richardson B, McKenzie A. Taxonomy and distribution of Australian Cephalochordates (Chordata: Cephalochordata). Invertebr. Taxon. 1994; 8: 1443–1459.
  12. 12. Nohara M, Nishida M, Miya M, Nishikawa T. Evolution of the mitochondrial genome in cephalochordata as inferred from complete nucleotide sequences from two Epigonichthys species. J Mol Evol. 2005a; 60: 526–37. pmid:15883887
  13. 13. Wang YQ, Fang SH. Taxonomic and molecular phylogenetic studies of amphioxus: a review and prospective evaluation. Zoolog Res. 2005; 26: 666–672.
  14. 14. Igawa T, Nozawa M, Suzuki DG, Reimer JD, Morov AR, Wang Y, et al. Evolutionary history of the extant amphioxus lineage with shallow-branching diversification. Sci Rep. 2017; 7: 1157. pmid:28442709
  15. 15. Gee H. Careful with that amphioxus. Nature. 2006; 439: 923–924. pmid:16495981
  16. 16. Bertrand S, Escriva H. Evolutionary crossroads in developmental biology: amphioxus. Development. 2011; 138: 4819–4830. pmid:22028023
  17. 17. Annona G, Holland ND, D’Aniello S. Evolution of the notochord. Evodevo. 2015; 5: 6:30. pmid:26446368
  18. 18. Shimeld SM, Holland PWH. Special Feature: Vertebrate innovations. Proc Natl Acad Sci USA 2000; 97: 4449–4452. pmid:10781042
  19. 19. Delsuc F, Brinkmann H, Chourrout D, Philippe H. Tunicates and not cephalochordates are the closest living relatives of vertebrates. Nature 2006; 439: 965–968. pmid:16495997
  20. 20. Pallas PS. Limax lanceolatus. Descriptio Limacis lanceolaris. In: spicilegia Zoologica, quibus novae imprimus et obscurae animalium species iconibus, descriptionibus. Gottlieb augusttus Lange, Berlin. 1774; 10: 9–14.
  21. 21. Gray J. Description of a new species of amphioxus from Borneo. J Nat Hist. 1847; 19: 463–464.
  22. 22. Stokes MD, Holland ND. The lancelet. Am Sci. 1998; 86: 552–560.
  23. 23. Schubert M, Escriva H, Xavier-Neto J, Laudet V. Amphioxus and tunicates as evolutionary model systems. Trends Ecol Evol. 2006; 21: 269–277. pmid:16697913
  24. 24. Bertrand S, Camasses A, Escriva H. Amphioxus: how to become a vertebrate. J Soc Biol. 2007; 201: 51–57. pmid:17762824
  25. 25. D’Aniello S, Irimia M, Maeso I, Pascual-Anaya J, Jiménez-Delgado S, Bertrand S, et al. Gene expansion and retention leads to a diverse tyrosine kinase superfamily in amphioxus. Mol Biol Evol. 2008; 25: 1841–54. pmid:18550616
  26. 26. Putnam NH, Thomas B, David E, Ferrier DEK, Furlong FR, Hellsten U, et al. The amphioxus genome and the evolution of the chordate karyotype. Nature. 2008; 453: 1064–1072. pmid:18563158
  27. 27. Huang S, Chen Z, Huang G, Yu T, Yang P, Li J, et al. HaploMerger: reconstructing allelic relationships for polymorphic diploid genome assemblies. Genome Res. 2012; 22: 1581–8. pmid:22555592
  28. 28. Marlétaz F, Firbas PN, Maeso I, Tena JJ, Bogdanovic O, Perry M, et al. Amphioxus functional genomics and the origins of vertebrate gene regulation. Nature. 2018; 564: 64–70. pmid:30464347
  29. 29. Nishikawa T. A Taxonomic Review of Lancelets (Cephalochordata) in Japanese Waters. In: Motokawa M, Kajihara H, (eds). Species Diversity of Animals in Japan. Diversity and Commonality in Animals. Springer, Tokyo; 2017. pp. 703–714. https://doi.org/10.1007/978-4-431-56432-4_27
  30. 30. Nishikawa T. Reinstatement of the Lancelet Name Asymmetron lucayanum, Recently Proposed as a Junior Synonym of Branchiostoma pelagicum (Cephalochordata). Species Divers. 2018; 23: 83–85.
  31. 31. Nohara M, Nishida M, Nishikawa T. New complete mitochondrial DNA sequence of the lancelet Branchiostoma lanceolatum (Cephalocordata) and the identity of this species’ sequences. Zool Sci. 2005b; 22: 61–64. pmid:15988162
  32. 32. Xu QS, Ma F, Wang YQ. Morphological and 12S rRNA gene comparison of two Branchiostoma species in Xiamen waters. J Exp Zool B Mol Dev Evol. 2005; 304: 259–267. pmid:15791653
  33. 33. Kon T, Nohara M, Nishida M, Sterrer W, Nishikawa T. Hidden ancient diversification in the circumtropical lancelet Asymmetron lucayanum complex. Mar Biol. 2006; 149: 875–883.
  34. 34. Zhang QJ, Zhong J, Fang SH, Wang YQ. Branchiostoma japonicum and B. belcheri are distinct lancelets (Cephalochordata) in Xiamen waters in China. Zoolog Sci. 2006; 23: 573–579. pmid:16849846
  35. 35. Xiao Y, Zhang Y, Gao T, Yabe M, Sakurai Y. Phylogenetic relationships of the lancelets of the genus Branchiostoma in China inferred from mitochondrial genome analysis. Afr J Biotechnol. 2008; 7: 3845–3852.
  36. 36. Subirana L, Farstey V, Bertrand S, Escriva H. Asymmetron lucayanum: How many species are valid? PLoS One. 2020; 15: e0229119. pmid:32130230
  37. 37. Vargas JA, Dean HK. On Branchiostoma californiense (Cephalochordata) from the Gulf of Nicoya estuary, Costa Rica. Revista de Biología Tropical. 2010; 58: 1143–1148. pmid:21246984
  38. 38. Sibaja-Cordero JA, Troncoso JS, Cortés J. The Lancelet Asymmetron lucayanum Complex in Cocos Island National Park, Pacific Costa Rica. Pac Sci. 2012; 66: 523–528.
  39. 39. Babu A, Sampathkumar P, Varadharajan D, Balasubramanian T. Branchiostoma malayanum—A Newly Recorded Amphioxus in Indian Coast. Int J Pharm Biol Arch. 2013; 4: 231–234.
  40. 40. Del Moral-Flores LF, Guadarrama-Martínez MÁ, Flores-Coto C. Composición taxonómica y distribución de los cefalocordados (Cephalochordata: Amphioxiformes) en México. Lat Am J Aquat Res. 2016; 44: 497–503.
  41. 41. Meerhoff E, Veliz D, Vega-Retter C, Yannicelli B. The amphioxus Epigonichthys maldivensis (Forster Cooper, 1903) (Cephalochordata Branchiostomatidae) larvae in the plankton from Rapa Nui (Chile) and ecological implications. Biodivers J. 2016; 7: 7–10.
  42. 42. Galván-Villa CM, Ríos-Jara E, Ayón-Parente M. New records of the Californian lancelet Branchiostoma californiense (Cephalochordata: Branchiostomidae) from the Pacific coast of Mexico. Rev Mex Biodiv. 2017; 88: 995–998.
  43. 43. Ishaq S, Siddiqui G. Amphioxus-Branchiostoma (Chordata: Cephalochordata): first report of its occurrence from the intertidal soft sediment benthic habitat of Clifton beach, Karachi, Pakistan (northern Arabian Sea). Int J Biol Biotech. 2017; 14: 419–422.
  44. 44. Nishikawa T, Namikawa H. New Japanese Localities for the Lancelets Asymmetron lucayanum complex and Epigonichthys cultellus [Cephalochordata], with Notes on their Northward Distribution Extensions. Bull Natl Mus Nat Sci, Ser A. 2017; 43: 93–99.
  45. 45. Campos-Dávila L, Pérez-Estrada CJ, Rodríguez-Estrella R, Morales-Bojórquez E, Brun-Murillo FG, Balart EF. Seagrass Halodule wrightii as a new habitat for the amphioxus Branchiostoma californiense (Cephalochordata, Branchiostomidae) in the southern Gulf of California, Mexico. Zookeys. 2019; 873:113–131. pmid:31534388
  46. 46. Rodríguez-Uribe MC, Chávez-Dagostino RM, Del Moral-Flores LF, Bravo-Olivas ML. First Record of Amphioxus Branchiostoma californiense (Amphioxiformes: Branchiostomatidae) Adjacent to a Shallow Submarine Hydrothermal System at Banderas Bay (Mexico). Diversity. 2019; 11: 227.
  47. 47. Costa OG. Cenni zoologici ossia descrizione sommaria delle specie nuove di animali discoperti in diverse contrade del regno nell’ anno 1834. Naples: Azzolino & Comp; 1834.
  48. 48. Willey A. The later larval development of amphioxus. QJ Microsc Sci. 1891; 82: 183–234.
  49. 49. Orton JH. On a hermaphrodite specimen of amphioxus with notes on experiments in rearing amphioxus. J Mar Biol Assoc U K. 1914; 10: 506–512.
  50. 50. Bone Q. The problem of the "Amphioxides" larva. Nature. 1957; 180: 1462–1464.
  51. 51. Webb JE. On the feeding and behaviour of the larva of Branchiostoma lanceolatum. Mar Biol. 1969; 3; 58–72.
  52. 52. Garcia-Fernàndez J, Jiménez-Delgado S, Pascual-Anaya J, Maeso I, Irimia M, Minguillón C, et al. From the American to the European amphioxus: towards experimental Evo-Devo at the origin of chordates. Int J Dev Biol. 2009; 53: 1359–66. pmid:19247934
  53. 53. Royo JL, Maeso I, Irimia M, Gao F, Peter IS, Lopes CS, et al. Transphyletic conservation of developmental regulatory state in animal evolution. Proc Natl Acad Sci USA. 2011; 108: 14186–91. pmid:21844364
  54. 54. Pascual-Anaya J, Adachi N, Álvarez S, Kuratani S, D’Aniello S, Garcia-Fernàndez J. Broken colinearity of the amphioxus Hox cluster. Evodevo. 2012; 3: 28. pmid:23198682
  55. 55. Somorjai IML, Martí-Solans J, Diaz-Gracia M, Nishida H, Imai KS, Escrivà H, et al. Wnt evolution and function shuffling in liberal and conservative chordate genomes. Genome Biol. 2018; 19: 98. pmid:30045756
  56. 56. Caccavale F, Coppola U, Vassalli QA, La Vecchia C, Palumbo A, D’Aniello E, et al. Transphyletic conservation of nitric oxide synthase regulation in cephalochordates and tunicates. Dev Genes Evol. 2020; 230: 329–338. pmid:32839880
  57. 57. Morgulis A, Coulouris G, Raytselis Y, Madden TL, Agarwala R, Schäffer AA. Database indexing for production MegaBLAST searches. Bioinformatics. 2008; 24: 1757–1764. pmid:18567917
  58. 58. Clark K, Karsch-Mizrachi I, Lipman DJ, Ostell J, Sayers EW. GenBank. Nucleic Acids Res. 2016; 44: D67–D72. pmid:26590407
  59. 59. Abascal F, Zardoya R, Telford MJ. TranslatorX: multiple alignment of nucleotidesequences guided by aminoacid translations. Nucleic Acids Res. 2010; 38: 7–13. pmid:20435676
  60. 60. Katoh K, Standley DM. MAFFT Multiple Sequence Alignment Software Version 7: Improvements in performance and usability. Molec. Bio. Evol. 2013; 30: 772–780. pmid:23329690
  61. 61. Glez-Peña D, Gómez-Blanco D, Reboiro-Jato M, Fdez-Riverola F, Posada D. ALTER: program-oriented format conversion of DNA and protein alignments. Nucleic Acids Res. 2010; 38: W14–W18. pmid:20439312
  62. 62. Felsenstein J. Evolutionary trees from DNA sequences: a Maximum Likelihood approach. J. Mol. Evol. 1981; 17: 368–376. pmid:7288891
  63. 63. Huelsenbeck J, Ronquist F. MRBAYES: bayesian inference of phylogenetic trees. Bioinformatics 2001; 17: 754–755. pmid:11524383
  64. 64. Stamatakis A. RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics. 2006; 22: 2688–2690. pmid:16928733
  65. 65. Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Höhna S, et al. MRBAYES 3.2: Efficient Bayesian phylogenetic inference and model selection across a large model space. Syst. Biol. 2012; 61:539–542. pmid:22357727
  66. 66. Miller MA, Pfeiffer W, Schwartz T. Creating the CIPRES Science Gateway for inference of large phylogenetic trees. 2010 Gateway Computing Environments Workshop (GCE), New Orleans, LA; 2010. pp. 1–8.
  67. 67. Lanfear R, Frandsen PB, Wright AM, Senfeld T, Calcott B. PartitionFinder 2: new methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analyses. Mol Biol Evol. 2016; 34:772–773. pmid:28013191
  68. 68. Akaike H. Information Theory and an Extension of the Maximum Likelihood Principle. In: Parzen E., Tanabe K, Kitagawa G, (eds). Selected Papers of Hirotugu Akaike. Springer Series in Statistics (Perspectives in Statistics). Springer, New York, NY; 1998. pp. 267–281. https://doi.org/10.1007/978-1-4612-1694-0_15
  69. 69. Felsenstein J. Confidence limits on phylogenies: an approach using the bootstrap. Evolution. 1985; 39: 783–791. pmid:28561359
  70. 70. Kimura M. A simple method for estimating evolutionary rate of base substitutions through comparative studies of nucleotide sequences. J Mol Evol. 1980; 16: 111–120. pmid:7463489
  71. 71. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: Molecular Evolutionary Genetics Analysis across computing platforms. Mol Biol Evol. 2018; 35: 1547–1549. pmid:29722887
  72. 72. Zhang J, Kapli P, Pavlidis P, Stamatakis A. A general species delimitation method with applications to phylogenetic placements. Bioinformatics. 2013; 29: 2869–2876. pmid:23990417
  73. 73. Puillandre N, Lambert A, Brouillet S, Achaz G. ABGD, automatic barcode gap discovery for primary species delimitation. Mol Ecol. 2012a; 21: 1864–1877. pmid:21883587
  74. 74. Posada D, Crandall KA. Intraspecific gene genealogies: trees grafting into networks. Trends Ecol Evol. 2001; 16: 37–45. pmid:11146143
  75. 75. Rodríguez-Flores PC, Machordom A, Abelló P, Cuesta JA, Macpherson E. Species delimitation and multi-locus species tree solve an old taxonomic problem for European squat lobsters of the genus Munida Leach, 1820. Mar Biodiv. 2019; 49: 1751–1773.
  76. 76. Puillandre N, Modica MV, Zhang Y, Sirovich L, Boisselier MC, Cruaud C, et al. Large-scale species delimitation method for hyperdiverse groups. Mol Ecol. 2012b; 21: 2671–2691. pmid:22494453
  77. 77. Blaxter ML. The promise of a DNA taxonomy. Philos Trans R Soc Lond B Biol Sci. 2004; 359: 669–679. pmid:15253352
  78. 78. Templeton AR. Using phylogeographic analyses of gene trees to test species status and processes. Mol. Ecol. 2001; 10: 779–791. pmid:11298987
  79. 79. Templeton AR, Crandall KA, Sing CF. A cladistic analysis of phenotypic associations with haplotypes inferred from restriction endonuclease mapping and DNA sequence data. III. Cladogram estimation. Genetics. 1992;132: 619–33. pmid:1385266
  80. 80. Clement M, Posada D, Crandall KA. TCS: a computer program to estimate gene genealogies. Mol Eco. 2000; 9:1657–1659. pmid:11050560
  81. 81. Hart MW, Sunday J. Things fall apart: biological species form unconnected parsimony networks. Biol Lett. 2007; 3: 509–512. pmid:17650475
  82. 82. Barco A, Houart R, Bonomolo G, Crocetta F, Oliverio M. Molecular data reveal cryptic lineages within the northeastern Atlantic and Mediterranean small mussel drills of the Ocinebrina edwardsii complex (Mollusca: Gastropoda: Muricidae). Zool J Linn Soc. 2013; 169: 389–407.
  83. 83. Zhong J, Zhang Q, Xu Q, Schubert M, Laudet V, Wang Y. Complete mitochondrial genomes defining two distinct lancelet species in the West Pacific Ocean. Mar. Biol. Res. 2009; 5: 278–285.
  84. 84. Zhang QL, Zhang GL, Yuan ML, Dong ZX, Li HW, Guo J, et al. A Phylogenomic Framework and Divergence History of Cephalochordata Amphioxus. Front Physiol. 2018; 9: 1833. pmid:30618839
  85. 85. Coll M, Piroddi C, Steenbeek J, Kaschner K, Ben Rais Lasram F, Aguzzi J, et al. The biodiversity of the Mediterranean Sea: estimates, patterns, and threats. PLoS One. 2010; 5: e11842. pmid:20689844
  86. 86. Bianchi CN, Morri C, Chiantore M, Montefalcone M, Parravicini V, Rovere A. 2012 Mediterranean Sea biodiversity between the legacy from the past and a future of change. In: Life in the Mediterranean Sea: a look at habitat changes (Stambler N., ed.), Nova Science Publishers, New York; 2012. pp. 1–55.
  87. 87. Bo M, Tazioli S, Spanò N, Bavestrello G. Antipathella subpinnata (Antipatharia, Myriopathidae) in Italian seas. Italian Journal of Zoology. 2008; 75: 185–195.
  88. 88. García-Merchán VH, Robainas-Barcia A, Abelló P, Macpherson E, Palero F, García-Rodríguez M, et al. Phylogeographic patterns of decapod crustaceans at the Atlantic-Mediterranean transition. Mol Phylogenet Evol. 2012; 62: 664–72. pmid:22138160
  89. 89. Vecchioni L, Marrone F, Deidun A, Adepo-Gourene B, Froglia C, Sciberras A, et al. DNA Taxonomy Confirms the Identity of the Widely-Disjunct Mediterranean and Atlantic Populations of the Tufted Ghost Crab Ocypode cursor (Crustacea: Decapoda: Ocypodidae). Zoolog Sci. 2019; 36(4), 322–329.
  90. 90. Crocetta F, Caputi L, Paz-Sedano S, Tanduo V, Vazzana A, Oliverio M. High genetic connectivity in a gastropod with long-lived planktonic larvae. J Molluscan Stud. 2020; 86: 42–55.
  91. 91. Patarnello T, Volckaert FA, Castilho R. Pillars of Hercules: is the Atlantic-Mediterranean transition a phylogeographical break? Mol Ecol. 2007; 16: 4426–44. pmid:17908222
  92. 92. González-Castellano I, González-López J, González-Tizón AM, Martínez-Lage A. Genetic diversity and population structure of the rockpool shrimp Palaemon elegans based on microsatellites: evidence for a cryptic species and differentiation across the Atlantic–Mediterranean transition. Sci Rep. 2020; 10: 10784. pmid:32612210
  93. 93. Mastrototaro F, Montesanto F, Salonna M, Viard F, Chimienti G, Trainito E, et al. An integrative taxonomic framework for the study of the genus Ciona (Ascidiacea) and description of a new species, Ciona intermedia. Zool J Linn Soc. 2020; 190: 1193–1216.
  94. 94. Villamor A, Signorini LF, Costantini F, Terzin M, Abbiati M. Evidence of genetic isolation between two Mediterranean morphotypes of Parazoanthus axinellae. Sci Rep. 2020; 10: 13938. pmid:32811877
  95. 95. Palumbi SR. Genetic Divergence, Reproductive Isolation, and Marine Speciation. Annu Rev Ecol Syst. 1994; 25: 547–572.
  96. 96. Miglietta MP, Faucci A, Santini F. Speciation in the Sea: Overview of the Symposium and Discussion of Future Directions. Integr Comp Biol. 2011; 51: 449–455. pmid:21593140
  97. 97. Potkamp G, Fransen CHJM. Speciation with gene flow in marine systems. Contrib Zool. 2019; 88: 133–12.
  98. 98. Pascual M, Rives B, Schunter C, Macpherson E. Impact of life history traits on gene flow: A multispecies systematic review across oceanographic barriers in the Mediterranean Sea. PLoS One. 2017; 12(5): e0176419. pmid:28489878
  99. 99. Barco A, Aissaoui C, Houart R, Bonomolo G, Crocetta F, et al. Revision of the Ocinebrina aciculata species complex (Mollusca: Gastropoda: Muricidae) in the northeastern Atlantic Ocean and Mediterranean Sea. J Molluscan Stud. 2018; 84: 19–29.
  100. 100. Pola M, Paz-Sedano S, Macali A, Minchin D, Marchini A, Vitale F, et al. What is really out there? Review of the genus Okenia Menke, 1830 (Nudibranchia: Goniodorididae) in the Mediterranean Sea with description of two new species. PLoS One. 2019; 14: e0215037. pmid:31042722
  101. 101. Voultsiadou E, Vafidis D. Marine invertebrate diversity in Aristotle’s zoology. Zootaxa. 2007; 76: 103–120.
  102. 102. Fasulo G, Duraccio S, Federico A, Crocetta F. The (almost) unknown Italian naturalist Raffaello Bellini (1874–1930): biography, malacological publications, and status of his recent molluscan taxa. Zootaxa. 2019; 4668: 343–369. pmid:31716617
  103. 103. Somorjai IML, Bertrand S, Camasses A, Haguenauer A, Escriva H. Evidence for stasis and not genetic piracy in developmental expression patterns of Branchiostoma lanceolatum and Branchiostoma floridae, two amphioxus species that have evolved independently over the course of 200 Myr. Dev Genes Evol. 2008; 218: 703–13. pmid:18843503
  104. 104. Bányai L, Kerekes K, Trexler M, Patthy L. Morphological Stasis and Proteome Innovation in Cephalochordates. Genes. 2018; 9: 353. pmid:30013013
  105. 105. Bellan-Santini D, Lacaze JC, Poizat C. Les Biocénoses marines et littorales de Méditerranée. Synthèse, menaces et perspectives. Museum National d’Histoire Naturelle. Paris. 1994.
  106. 106. Zhao Q, Zhu Q. Taxonomic and genetic status of lancelet in Weihai coastal waters based on mitochondrial DNA sequence. Chin J Ocean Limnol. 2011; 29: 623.
  107. 107. Li W, Zhong J, Wang Y. Genetic diversity and population structure of two lancelets along the coast of China. Zoolog Sci. 2013; 30(2):83–91. pmid:23387841
  108. 108. Webb JE. The ecology of Lagos Lagoon III. The life history of Branchiostoma nigeriense Webb. Philos Trans R Soc Lond B Biol Sci. 1958; 241: 335–353.
  109. 109. Stokes MD, Holland ND. Embryos and Larvae of a Lancelet, Branchiostoma floridae, from Hatching through Metamorphosis: Growth in the Laboratory and External Morphology. Acta zool. 1995; 76: 105–120.
  110. 110. Urata M, Yamaguchi N, Henmi Y, Yasui K. Larval development of the oriental lancelet, Branchiostoma belcheri, in laboratory mass culture. Zoolog Sci. 2007; 24: 787–97. pmid:18217485
  111. 111. Spruyt N, Delarbre C, Gachelin G, Laudet V. Complete sequence of the amphioxus (Branchiostoma lanceolatum) mitochondrial genome: relation to vertebrates. Nucleic Acids Res. 1998; 26:3279–3285. pmid:9628930
  112. 112. Boore JL, Daehler LL, Brown W. Complete sequence, gene arrangement, and genetic code of mitochondrial DNA of the Cephalochordata Branchiostoma floridae (Amphioxus). Mol Biol Evol. 1999; 16:410–418. pmid:10331267
  113. 113. Yue JX, Yu JK, Putnam NH, Holland LZ. The transcriptome of an amphioxus, Asymmetron lucayanum, from the Bahamas: a window into chordate evolution. Genome Biol Evol. 2014; 6(10):2681–96. pmid:25240057