Research Article
Print
Research Article
Molecular systematics of Perinereis and an investigation of the status and relationships of the cultured species Perinereis wilsoni Glasby & Hsieh, 2006 (Annelida, Nereididae)
expand article infoDeyuan Yang§, Sheng Zeng§, Zhi Wang§, Yanjie Zhang|, Dazuo Yang, Christopher J. Glasby#¤, Jiang‐Shiou Hwang, Lizhe Cai§
‡ National Taiwan Ocean University, Keelung, Taiwan
§ Xiamen University, Xiamen, China
| Hainan University, Haikou, China
¶ Dalian Ocean University, Dalian, China
# Natural Sciences, Museum & Art Gallery Northern Territory, Darwin, Australia
¤ Australian Museum Research Institute, Sydney, Australia
Open Access

Abstract

In this study, we conducted morphological and molecular analyses of Perinereis wilsoni, a species being considered for aquaculture in China. We found this species difficult to identify because of its close morphological similarity to the sympatric P. mictodonta and thus sought genetic markers to more easily distinguish it and to investigate its phylogenetic relationship to P. mictodonta and other nereidids. For the first time, we sequenced, assembled, and annotated the complete mitochondrial genome, nuclear ribosomal sequences (18S-ITS1-5.8S-ITS2-28S), and four nuclear histone genes (H3-H2A-H2B-H4) of P. wilsoni. Comprehensive bioinformatics methods were employed to assemble the genome-skimming data of P. wilsoni to ensure assembly quality. Phylogenetic analyses based on five datasets of the available mitochondrial genomes (32 taxa in Nereididae, including 8 taxa in Perinereis), using maximum likelihood and Bayesian analyses, provide support for the monophyly of the genus Perinereis. In contrast, the P. nuntia species group, a subgroup within Perinereis, is nonmonophyletic. Perinereis wilsoni has a closer phylogenetic relationship with P. vancaurica and P. nuntia. Our study serves as a baseline for future work on the cultivation, reproductive biology, and phylogeny of P. wilsoni.

Key Words

Bioinformatic analyses, genome skimming, mitogenomes, Perinereis

Introduction

Recently, a local Perinereis species intended for large-scale aquaculture in China by the Key Laboratory of Marine Bio-resource Restoration and Habitat Reparation in Liaoning Province, Dalian Ocean University, was identified as P. wilsoni Glasby & Hsieh, 2006. The genus Perinereis Kinberg, 1865, belongs to the Nereididae. This family comprises over 700 described species and 45 genera (Wilson et al. 2023). Many nereidid species, particularly Perinereis, are of significant commercial and ecological importance for fishing bait, aquaculture feed, and wastewater treatment (Palmer 2010; Arias et al. 2013). In China, Perinereis aibuhitensis (Grube, 1878) is farmed and exported worldwide as fish bait, and Tylorrhynchus heterochetus (Quatrefages, 1866) is consumed as food by people in Guangdong Province and other Southeast Asian countries (Glasby and Hsieh 2006; Wilson et al. 2023).

Perinereis represents the second-largest genus in Nereididae, with approximately 100 species (Mahcene et al. 2023; Prajapat et al. 2023; Wilson et al. 2023; Teixeira et al. 2024). Bakken and Wilson (2005) conducted a phylogenetic analysis based on morphological evidence, suggesting that Perinereis may be polyphyletic. This was supported by subsequent molecular studies (Glasby et al. 2013; Alves et al. 2020; Elgetany et al. 2022). For practical identification purposes, Hutchings et al. (1991) first divided Perinereis into three species groups based on the number of pharyngeal Area VI paragnaths and further divided each group based on parapodial characters. The Perinereis nuntia species group was characterized by a distinctive arc of bar-shaped paragnaths on area VI (usually numbering 6–14 on each side) (Hutchings et al. 1991; Wilson and Glasby 1993; Glasby and Hsieh 2006). This group has been reviewed by Wilson and Glasby (1993), Glasby and Hsieh (2006), and Villalobos-Guerrero (2019). Currently, it comprises approximately 20 recognized species (Villalobos-Guerrero 2019; Wilson et al. 2023), with the type localities of about 13 species in the Indo-West Pacific. These studies have significantly improved our understanding of this group. However, relying solely on a morphology-based approach may not be sufficient to resolve the taxonomic status of these species because some key taxonomic characters used to distinguish species often overlap, as they show ontogenetic and intraspecific variation (Tosuji et al. 2019; Tosuji et al. 2023). These morphology-defined species need to be re-evaluated using integrated taxonomic methods by taxonomic specialists. Recently, studies based on integrative taxonomy have revealed more new species in Nereididae (see Glasby et al. 2013; Teixeira et al. 2022a; Teixeira et al. 2022b).

Perinereis wilsoni, a member of the P. nuntia species group, is morphologically very similar to Perinereis mictodonta (Marenzeller, 1879), with no distinct morphological differences, although slight but statistically significant morphometric differences were found with respect to paragnath numbers and the relative length of the dorsal cirri (Glasby and Hsieh 2006). For this reason, these two species were primarily established based on the differences in ITS (internal transcribed spacer) genes (Chen et al. 2002; Glasby and Hsieh 2006). Therefore, identifying these two species based solely on morphology presents a significant challenge. To date, only two studies have contributed molecular data: Chen et al. (2002) provided ITS genes. Tosuji et al. (2019) provided sequences for both the partial mitochondrial 16S ribosomal RNA (16S) and ITS genes. Although Park and Kim (2007) researched P. wilsoni using the partial mitochondrial cytochrome oxidase I (COX1) gene, these sequences were not made publicly available. Moreover, there are some COX1 sequences labeled P. wilsoni in the NCBI GenBank database, but these sequences remain pending verification.

Given the considerable economic importance of Perinereis species, accurate species identification is crucial because correct scientific names can facilitate the linking of subsequent physiological or reproductive studies, thus ensuring the reproducibility of these research efforts (Glasby and Hsieh 2006; Hutchings and Lavesque 2020). Therefore, we have focused on several key questions regarding taxonomy: 1) Is the identification of P. wilsoni from Liaoning Province accurate? 2) Which molecular markers are suitable for identifying this species? 3) What is the phylogenetic relationship of P. wilsoni to other nereidids?

In this study, we used low-coverage whole genome sequencing, also known as genome skimming (Straub et al. 2012), which has been widely used in polychaete phylogeny studies because it is cost-effective and does not require fresh tissue (Richter et al. 2015; Coissac et al. 2016; Alves et al. 2020; Hektoen et al. 2024). Utilizing this strategy, we provided more molecular information, including high-copy regions: mitochondrial genome (mitogenome), nuclear ribosomal sequences (18S rRNA-ITS1-5.8S rRNA-ITS2-28S rRNA), and nuclear histone genes (H3-H2A-H2B-H4).

Materials and methods

Sample collection, identification, and sequencing

The specimens of P. wilsoni were sampled from Dalian, Liaoning Province, China (38.8732°N, 121.6767°E) and identified by Deyuan Yang. All specimens were fixed directly in 95% ethanol. Two specimens with a fully-everted pharynx were used for morphological and molecular studies. They were deposited at Xiamen University (XMU) under voucher numbers 23007-1 and 23007-2, respectively. Specimens were identified based on the key in Glasby and Hsieh (2006). Methods used in the morphological study were followed by Yang et al. (2022). In summary, whole worms were photographed with an Olympus E-M1 Mark II camera with a 60 mm macro lens, detailed structures with a Zeiss Discovery V20 stereomicroscope, and a Nikon 80i compound microscope. Image stacks were obtained using Helicon Focus 7 (https://www.heliconsoft.com/heliconsoft-products/helicon-focus/) and post-processed using Adobe Photoshop. The terminology of nereidid followed Wilson et al. (2023) generally. To better describe areas II, III, and VI of paragnaths, we introduced the notations ‘L:R’ and ‘L:M:R’. ‘L:R’ is for areas II and VI, and ‘L:M:R’ to describe the paragnath patch in area III, where ‘L’ represents the number of paragnaths on the left, ‘M’ in the middle, and ‘R’ on the right. For example, for area II, ‘2:3’ signifies 2 paragnaths on the left and 3 on the right; for area III, ‘3:5,2R:7’ signifies 3 paragnaths on the left, a central patch with 5 paragnaths in 2 rows, and 7 paragnaths on the right.

Before DNA extraction, each individual was cleaned with 95% ethanol. To avoid gut contamination, two to seven parapodia were clipped from the specimens. Whole genomic DNA was extracted using the TIANamp Genomic DNA Kit (TIANGEN, Beijing, China). Genome skimming was conducted on the Illumina NovaSeq X Plus with a PE150 strategy. DNA extraction and sequencing were performed at Novogene Bioinformatics Technology Co., Ltd. (Beijing, China). Our initial sequencing effort was to obtain 5 Gb of raw data for each sample. Voucher number 23007-2 was increased to 25 Gb to explore more universal single-copy ortholog genes.

Assembly and annotation of mitochondrial and nuclear sequences

Raw paired-end reads were removed from sequence adapters, and low-quality regions were trimmed using Fastp v.0.23.4 (Chen et al. 2018). The quality of all cleaned reads was checked with FastQC v.0.12.1 (http://www.bioinformatics.babraham.ac.uk/projects/fastqc/), and subsequently, FastQC results were summarized using MultiQC v.1.8 (Ewels et al. 2016) (Suppl. material 1).

To ensure reliable assembly for the present data, various assemblers were used: 1) SPAdes v.3.15.5 (Bankevich et al. 2012), Ray v.2.3.1 (Boisvert et al. 2010), Megahit v.1.2.9 (Li et al. 2015), and IDBA-UD v.1.1.3 (Peng et al. 2012); 2) GetOrganelle v.1.7.6.1 (Jin et al. 2020); 3) MITObim v.1.9 (Hahn et al. 2013) and NovoPlasty v.4.3.1 (Dierckxsens et al. 2017), with 16S (GenBank accession numbers: LC482188) as the seed; 4) our new assembly pipeline, FastMitoAssembler (original version at https://github.com/suqingdong/FastMitoAssembler). The pipeline implemented in FastMitoAssembler involved the following steps: First, MEANGS v.1.2.1 (Song et al. 2022) was used to obtain and assemble mitochondrial genes, which then served as a seed input for NovoPlasty. Second, the results of NovoPlasty served as a seed input for GetOrganelle. Third, MitoZ v3.6 (Meng et al. 2019) was used for annotation. This workflow was managed using Snakemake (Köster and Rahmann 2012), which can handle large-scale samples with minimal input–only raw data.

For de novo assembly results (SPAdes, Megahit, Ray, and IDBA-UD), we used MitoFinder v.1.4 (Allio et al. 2020) to extract mitochondrial contigs (mt contigs) and annotations, using the Platynereis dumerilii mitogenome (GenBank: NC 000931) as a reference. Unfortunately, no mt contigs were found from the Ray assembler. If the circularization of a mt contig was not automatically completed by MitoFinder, the circules.py script (https://github.com/chrishah/MITObim/tree/master/misc_scripts) was used to assess the circularization of the contig. After obtaining mitogenomes from various assemblers, Quast v.5.2.0 (Gurevich et al. 2013) was employed for assembly quality evaluation (Suppl. material 2). A BLASTn query against GenBank in the NCBI (NCBI BLAST Analysis, https://BLAST.ncbi.nlm.nih.gov/BLAST.cgi) was also performed to check for contamination in the mitogenomes (Suppl. material 3: table S1). The GC content and read (per-base) coverage depth of mitogenome sequences were calculated using the visualize subcommand in MitoZ v.3.6 with default parameters.

The mitogenome was annotated using MITOS2 (Donath et al. 2019), MitoFinder v.1.4.1, and MitoZ v.3.6. All annotation files (in .gb or .gbf formats) were reordered with COX1 as the start gene using PhyloSuite v.1.2.3 (Zhang et al. 2020). The reordered GenBank format (gb) file was converted to a fasta file and uploaded to GeSeq (https://chlorobox.mpimp-golm.mpg.de/geseq.html) to combine the results of ARWEN v.1.2.3 and tRNAscan-SE v.2.0.7 into a single gb file. This approach allowed us to compare tRNA annotations generated from three common methods (ARWEN, tRNAscan-SE, and MiTFi). All annotations were then loaded into Geneious Prime v.2022.2.2 for manual curation. First, we utilized the MAFFT plugin in Geneious to align these gb files. Then, we checked the coding region annotations with the assistance of the ORFs function based on the invertebrate mitochondrial genetic code table 5. Second, tRNAs were evaluated using the following criteria: tRNAscan-SE was given precedence over ARWEN and MiTFi, while MiTFi had precedence over ARWEN. Thus, if tRNAscan-SE identified a tRNA, its result was chosen. In cases where both MiTFi and ARWEN identified a tRNA, the result from MiTFi was selected. Subsequently, we imported other gb files of Perinereis into Geneious to compare each gene and manually edit them.

The boundaries of each gene were determined by the following rules: 1) Protein-coding genes can overlap with each other but cannot overlap with tRNA genes. In Nereididae, only the ND4 and ND4L coding genes are allowed to overlap. 2) The boundaries of mitochondrial rRNA genes (12S and 16S) were defined by flanking genes. 3) The large non-coding region (or putative control region) was determined by the boundaries of neighboring genes and a low GC region.

The CGView Comparison Tool (CCT) (https://github.com/paulstothard/cgview_comparison_tool) (Grant et al. 2012) was employed to compare all accessible mitogenomes of Nereididae in NCBI (as of 15 January 2024) to evaluate the quality of annotations further.

The high-copy nuclear markers (18S to 28S genes and histone genes) were assembled using GetOrganelle v.1.7.6.1. Considering the lack of effective software or pipelines to annotate these genes, we manually annotated nuclear rRNA genes in Geneious using primer pairs to define gene boundaries: 18SA and 18SB (Medlin et al. 1988) for 18S, F63.2 and R3264.2 (Struck et al. 2006) for 28S, and ITS18SFPOLY and ITS28SRPOLY (Nygren et al. 2009) for the ITS1-5.8S-ITS2 region. After completing one sequence annotation, we utilized the “Transfer annotations” function within Geneious Prime v.2022.2.2 to annotate other sequences. The histone genes were annotated using the same function, with GenBank accession number X58895 (Urechis caupo) as a reference at an 85% similarity threshold. If the number of gene annotations was incomplete in one sequence, we imported the corresponding cleaned sequencing reads into Geneious and reassembled the sequence employing the “map to reference” function in Geneious Prime v.2022.2.2 (mapper: Geneious, fine-tuning: iterate up to 25 times) and then re-annotated it. Once annotations were completed, each gene was extracted for NCBI BLAST analysis. The read (per-base) coverage depth of nuclear genes was calculated using a custom script, which used BWA (Li 2013) and SAMtools (Li et al. 2009).

BUSCO v.5.6 was employed to generate universal single-copy orthologs (USCOs) (Simão et al. 2015) using only the assembly results (contigs) from SPAdes with the following parameters: -l metazoan_odb10, -m genome, --augustus, -c 40. A total of 618 single-copy genes were searched.

Phylogenetic analyses

The phylogenetic analysis utilized two types of datasets: complete or near-complete mitogenomes of Nereididae and DNA barcoding sequences (COX1, 16S, and ITS) of Perinereis. All publicly available Nereididae mitogenomes from GenBank (as of 15 January 2024), along with two outgroup species (Craseoschema thyasiricola NC 060815 and Leocrates chinensis NC 066969, belonging to Chrysopetalidae and Hesionidae, respectively), were included in the phylogenetic analysis. The outgroup taxa were selected based on previous hypotheses that Chrysopetalidae and Hesionidae are sister groups to Nereididae (Glasby 1993; Pleijel and Dahlgren 1998; Tilic et al. 2022). To ensure consistent annotation criteria and to correct potential errors in previous annotations of mitogenomes, we re-annotated all available Nereididae mitogenomes using the methods described above. Ultimately, Platynereis massiliensis (NC 051996) and Alitta succinea (NC 051993) from the Nereididae were excluded from this study (Suppl. material 3: table S2).

To identify potential taxonomically mislabeled sequences in the mitogenomes dataset, we extracted the COX1 and 16S genes and conducted NCBI BLAST analysis. The results were downloaded as CSV files. For each sequence, the top 10 matching sequences were sorted based on identity using WEKits v.1.0 (https://github.com/GP-sir/wekits/releases). Next, we retrieved the source information (subjectAcc) for these matched sequences from NCBI. The species identification of each sequence was then manually verified in Microsoft Excel. Specifically, we focused on matched sequences with identity over 97% and checked whether they were supported by reliable taxonomic references (Suppl. material 3: table S3, S4).

All manually curated mitogenomes were imported into PhyloSuite v.1.2.3. The extracted protein-coding genes (PCGs) and two rRNAs (2R) from these sequences were aligned using MAFFT in normal mode. MACSE v.2 (Ranwez et al. 2018) was used to improve the multiple sequence alignment of 13 PCGs in the “refinement strategy.” After completing the alignment tasks, each alignment was manually checked in Geneious. Ambiguously aligned regions were removed using Gblocks with the default settings or trimAl v.1.2 (Capella-Gutiérrez et al. 2009) with the automated1 setting (Suppl. material 3: table S5). Five mitogenomic datasets were used in the phylogenetic analyses: (i) 13PCGs: PCGs with all three codon positions; (ii) 13PCGs12: PCGs without 3rd codon positions; (iii) 13PCGs_2R: PCGs dataset plus two rRNA genes; (iv) 13PCG12_2R: PCG12 dataset plus two rRNA genes; and (v) 13PCGsAA: amino acid sequences translated from 13PCGs. The heterogeneity of each dataset was assessed using AliGROOVE v.1.08 (https://github.com/PatrickKueck/AliGROOVE) with the default settings.

ModelFinder v.2.2.0 was used to select the best substitution models for each partition of the maximum likelihood (ML) and Bayesian inference (BI) analyses (Suppl. material 3: table S6). ML analysis was carried out in IQ-TREE v.2.2.2 under the edge-linked partition model with 20,000 ultrafast bootstraps. BI analysis was performed using MrBayes v.3.2.7a (Ronquist et al. 2012), with two parallel runs for at least 2,000,000 generations, ensuring the average standard deviation of split frequencies was below 0.01 (Ronquist et al. 2012).

Considering the mitogenomes of Perinereis are limited, shorter DNA sequences (DNA barcodes) were also used to explore the phylogenetic position of Perinereis wilsoni. All available Perinereis genes from NCBI (https://www.ncbi.nlm.nih.gov/nuccore/?term=Perinereis) were downloaded. The COX1, 16S, 18S, 28S, H3, and ITS genes were further extracted because these molecular markers are widely employed in Annelida. To organize the data more efficiently, we used a custom script (Suppl. material 4) to categorize these data based on voucher numbers (treating isolates, clones, strains, and haplotypes as synonymous with voucher numbers). Ultimately, the COX1, 16S, and ITS genes were selected for phylogenetic analysis due to their higher representation among taxa, and new sequences from this study were added to explore the phylogenetic position of Perinereis wilsoni (Suppl. material 3: table S7). The trees constructed with the COX1, 16S, and ITS genes used the same methods as the mitogenome data.

Multiple alignment summary statistics were generated using the alignment summary function in BioKIT (Steenwyk et al. 2022), including the number of taxa (sequences), alignment lengths, number of constant sites, number of parsimony informative sites and variable sites, and the frequency of all character states.

Pairwise tree structure comparison was conducted using the all.equal.phylo function in the ape v.5.7.1 package (Paradis and Schliep 2019), and topological differences were assessed using TreeSpace (Jombart et al. 2017), both implemented in R v.4.3.1 (R Core Team 2023). Finally, iTOL v.6 (Letunic and Bork 2021) was used to visualize the trees.

Sequence analyses

Strand asymmetries were calculated using the following formulae (Perna and Kocher 1995): AT-skew = (A - T) / (A + T); GC-skew = (G - C) / (G + C). Codon usage and relative synonymous codon usage (RSCU) of 13 PCGs were computed in PhyloSuite and visualized using the ‘ggplot2’ package (Wickham 2016) in R v.4.1.3 (R Core Team 2023). DnaSP v.6.0 (Rozas et al. 2017) was used to calculate the non-synonymous (Ka)/synonymous (Ks) substitution rates among the 13 PCGs of Nereididae and the nucleotide diversity (Pi) with a sliding window of 100 bp and a step size of 25 bp. A Ka/Ks ratio < 1 indicates purifying selection, while a ratio > 1 indicates positive selection.

To determine which molecular marker is reliable for accurate species delimitation, all labeled P. wilsoni and P. mictodonta COX1, 16S, and ITS genes in NCBI were used. We used the following criteria: 1) whether the presence of a barcode gap: the minimum interspecific genetic distance is greater than the maximum intraspecific genetic distance; 2) whether each species is recovered as monophyletic; and 3) whether there is a small overlap between intra- and interspecific distances and a large barcode gap, as recommended by dos Santos Vieira et al. (2020). Genetic distances (p-distance and Kimura 2-parameter) over sequence pairs were calculated in MEGA X (Kumar et al. 2018). The heat maps were generated in TBtools v.2.080 (Chen et al. 2020).

Results

Morphological analyses

Perinereis wilsoni Glasby & Hsieh, 2006

Fig. 1A–K

Material examined

• 23007-1 and 23007-2, both collected from Dalian, Liaoning, China (38.87315°N, 121.676671°E), 08 August 2023, preserved in 95% ethanol.

Description

Description based on 23007-1 and 23007-2. 23007-1 complete, 93 chaetigers, 7.0 cm in length, 2.9 mm wide at chaetiger 10 (excluding parapodia); and 4.0 mm wide at chaetiger 10 (including parapodia) (Fig. 1A). 23007-2 complete, 6.5 cm in length, 98 chaetigers, 2.3 mm wide at chaetiger 10 (excluding parapodia), 3.7 mm wide at chaetiger 10 (including parapodia).

Figure 1. 

Perinereis wilsoni Glasby & Hsieh, 2006; A–I. (except J, K) 23007-1; J, K. 23007-2; A. Entire body in dorsal view; B. Right jaw, ventral, and dorsal view; C. Anterior region with pharynx everted, dorsal view; D. Maxillary ring, frontal view; E. Anterior end with pharynx everted, ventral view; F. Left parapodium, posterior view, chaetiger 10; G. Right parapodium, posterior view, chaetiger 30; H. Sub-acicular neuropodial heterogomph falciger, chaetiger 30; I. Right parapodium, posterior view, chaetiger 80; J. Anterior view with pharynx everted, dorsal view; K. Anterior view with pharynx everted, ventral view. All photos were taken by Deyuan Yang.

Prostomium and anterior dorsum with dark brown pigmentation. Prostomium anterior margin entire, pear-shaped, wider than long, with shallow longitudinal groove in central area. Antennae conical, about one-third length relative to prostomium length. Palps longer than prostomium, biarticulate, with palpophores and palpostyles (Fig. 1A, C, J). Palpophores barrel-shaped, slightly wider basally, palpostyles spherical (or globular). Eyes black (Fig. 1A).

Pharynx fully everted (Fig. 1A, C–F, J–K). Jaws brown, with 7 teeth (based on 23007-1) (Fig. 1B). Tentacular cirri extend back 3–12 setigers; posterodorsal one extending to chaetiger 8–12.

Paragnath counts (Fig. 1C–E, J–K): area I with 1 conical paragnath; area II with 9 or 11 conical paragnaths on left, 11 or 14 conical paragnaths on right (23007-1, II = 9:11; 23007-2, II = 11:14); area III with 14–17 conical paragnaths, central patch with 9–12 in 2 rows, 2–3 laterally on either side (23007-1, III = 3:9,2R:2; 23007-2, III = 2:12,2R:3); area IV with 18–22 conical paragnaths on each side (23007-1, IV = 18:25; 23007-2, III = 22:22), bars absent; area V with 1 conical paragnath; area VI with 3–5 shield-shaped bars on each side (23007-1, VI = 5:3; 23007-2, VI = 5:5); area VII–VIII with 19 conical paragnths in 2 rows.

Notopodia with 2 lobes, prechaetal lobe absent. Dorsal cirri longer than notopodial ligule, about 1.5 times length of the notopodial ligule throughout. Notopodial ligule similar to median ligule throughout. Neuropodial postchaetal lobe rounded, not projecting beyond acicular lobe. Ventral ligule similar in length to acicular ligule in all chaetigers (Fig. 1F, G, I).

Notopodia homogomph spinigers only. Neuropodial heterogomph spinigers present throughout. At chaetiger 10, lower neurochaetae all heterogomph falcigers, upper neurochaetae with heterogomph falcigers, and heterogomph spinigers. At chaetiger 30 and following chaetigers, lower neurochaetae present heterogomph spinigers.

Remarks

Perinereis wilsoni was established by Glasby and Hsieh (2006), who provided a comprehensive description and discussion. In this study, we initially identified our specimens as P. wilsoni based on the key and description provided by Glasby and Hsieh (2006). Specifically, in P. wilsoni, the dorsal cirri are approximately 1.5 times the length of the dorsal notopodial lobe anteriorly and 2–3 times longer posteriorly; Area IV has 29.8 (± 3.6) conical paragnaths, and Area V has 1–3 conical paragnaths, usually in a longitudinal line. In contrast, in P. mictodonta, the dorsal cirri are either equal to or only slightly longer than the dorsal notopodial lobe throughout the body; Area IV has 35.3 (± 6.8) conical paragnaths, and Area V has 3 conical paraganths, usually in a triangle arrangement. Our specimens more closely resemble P. wilsoni, having dorsal cirri about 1.5 times the length of the dorsal notopodial lobe throughout the body (or only increasing slightly in length posteriorly), Area IV with 18–22 conical paragnaths, and Area V with 1 conical paragnath. However, despite the statistically significant difference in these morphometric characters (see pp. 573, Glasby and Hsieh 2006), the characters all show overlap (see pp. 560 & 572, Glasby and Hsieh 2006), and thus, morphology alone may not be sufficient to distinguish all specimens belonging to P. wilsoni and P. mictodonta.

The ITS genes may presently be the most effective and reliable method for accurately identifying these species, as these sequences were generated from the paratypes of P. wilsoni (Chen et al. 2002; Glasby and Hsieh 2006). DNA sequence-based NCBI BLAST and phylogenetic analyses of ITS sequences also support our identification (see next sections).

Distribution

Japan; China (based on ITS genes). Other records require validation using at least ITS data.

Characteristics of the Perinereis wilsoni mitochondrial genome and nuclear genes

The mitogenomes (23007-1 and 23007-2) generated from various assemblers yielded consistent results, except for the Ray assembler. They are 15,817 bp with an average coverage depth of 315×, and read mapping is 0.09% (Suppl. material 2). The genomes contain 13 protein-coding genes (PCGs), two ribosomal RNA (rRNA) genes, 22 transfer RNA (tRNA) genes, and one putative control region measuring 1160 bp. All genes are distributed on the heavy (H-) strand, similar to other Nereidinae species (Fig. 2A, Suppl. material 3: table S8). The nucleotide identity of 13 PCGs between P. wilsoni and the other 31 Nereididae species showed that the other species of Perinereis and Neanthes acuminata (OQ729916) have a high nucleotide identity with P. wilsoni (Fig. 2D).

Figure 2. 

The gene map of Perinereis wilsoni (23007-1), 23007-2, is the same as 23007-1 but only shows one; A. Mitogenome; B. Nuclear rRNA cluster; C. And histone genes; D. CCT (CGView Comparison Tool) map and sequence identity compare the mitogenome between P. wilsoni and the other nereidids. A–C. The green lines depict the distribution of coverage depth; D. Starting from the outermost ring, the feature rings depict: 1. COG (Clusters of Orthologous Groups of Proteins) functional categories for forward strand coding sequences; 2. Forward strand sequence features; 3. The remaining rings show regions of sequence similarity detected by BLAST comparisons between CDS translations from the reference genome and 31 comparison genomes. BLAST identities are organized from high to low, with higher values closer to the outer ring.

The NCBI BLAST results of the mitogenome revealed that the sequence identity with the published 16S sequences of P. wilsoni (LC482171LC482183, Tosuji et al. 2019) ranged from 98.34% to 99.52%. However, the NCBI BLAST results for the full-length 16S indicated over 98% sequence identity with both P. mictodonta (e.g., LC482161LC482168, Tosuji et al. 2019) and P. wilsoni (e.g., LC482171LC482183, Tosuji et al. 2019). These results show that the 16S marker is not suitable for distinguishing these two species.

The nuclear rRNA contigs for the two specimens (23007-1 and 23007-2) yielded 10,983 bp and 10,820 bp with average coverages of 218.3× and 306×, respectively. Both results contained the full 18S, ITS1, 5.8S, ITS2, and 28S regions, with gene lengths of 1,849 bp, 358 bp, 160 bp, 305 bp, and 3,840 bp, respectively (Fig. 2B). The nuclear histone genes for both specimens were 6,399 bp, with coverage depths of 1,547.5× and 2,620×, respectively, incorporating complete sequences of H3 (411 bp), H2A (375 bp), H2B (372 bp), and H4 (312 bp) genes (Fig. 2C).

The NCBI BLAST analysis of nuclear rRNA and histone genes showed that the 18S sequence had over 99% sequence identity with eight species belonging to five genera, indicating that 18S may not be suitable for species identification: 91.43% for 28S with Alitta virens (OW028578); 98.48%-98.54% for ITS with P. wilsoni (AF332158AF332162, Chen et al. 2002); 96%-98% for H3 genes with Perinereis sp. A ZW-2022 (OL546356), P. aibuhitensis (MW622055), and P. suluana (JX443591); approximately 92% for H2A and H2B with A. virens (OW028584), P. dumerilii (X53330); and 89.4% for H4 (Suppl. material 3: table S1).

Mitochondrial genes and codon usage

The mitochondrial genes in P. wilsoni exhibit a high A + T content of 64.6% (34.3% T and 30.3% A) and lower levels of C and G at 21.6% and 13.8%, respectively (Suppl. material 3: table S8). AT-skew and GC-skew are negative for the mitogenome (-0.062 and -0.221) and PCGs except for COX2 (-0.084 and -0.250), and AT-skew is positive for tRNAs and rRNAs (0.041 and 0.043) (Suppl. material 3: table S8). The mitogenome is compact, with a total of three gene overlaps ranging in length from 2 to 7 bp. There are 12 gene spacers (259 bp in total) from 1 bp to 78 bp. All 13 PCGs start with the ATG codon and stop with T, TAA, or TAG (Suppl. material 3: table S9). A total of 22 tRNA genes with lengths ranging from 53 to 68 bp were identified in the mitogenome of P. wilsoni.

The most frequently utilized amino acids in the mitogenome of P. wilsoni are Leu (14.83%), Ile (9.94%), Ser (8.60%), and Ala (7.84%). The least common amino acids are Cys (1.01%), Arg (1.78%), Asp (1.78%), and Gln (1.94%) (Fig. 3, Suppl. material 3: table S10). Relative synonymous codon usage (RSCU) values for the 13 PCGs showed that UCU (Ser) and UUA (Leu) are the two most frequent codons, whereas UGG (Trp) and CCG (Pro) have the lowest frequencies (Fig. 3).

Figure 3. 

The relative synonymous codon usage (RSCU) in the mitogenome of Perinereis wilsoni. The numbers on the bars represent amino acid composition.

Phylogenetic analyses and genetic distances

Summary statistics for multiple alignments of various datasets are available in Suppl. material 3: table S11. The heterogeneity analysis of five mitogenome datasets revealed that datasets excluding the third codon position (13PCGs12, 13PCGs12_2R) exhibited lower heterogeneity compared to those including the third codon position (13PCGs, 13PCGs_2R). The lowest heterogeneity was observed in the 13PCGsAA dataset (see Suppl. material 5).

A total of ten ML and BI trees were inferred from five mitogenome datasets of Nereididae (see Suppl. material 6). These trees exhibit eight different topologies based on pairwise tree comparisons, indicating the significant influence of the chosen mitogenomic dataset and tree-building method on the inferred phylogenetic relationships. In general, ML and BI trees inferred from the same dataset are generally consistent, except for some obvious differences observed in the 13PCGs_2R and 13PCGs12_2R datasets (see Suppl. material 6). In the 13PCGs12 dataset, we could not re-root the ML and BI trees using the outgroup, which was excluded from TreeSpace analysis. The remaining eight trees were clustered into four groups, representing four types of trees (see Fig. 4B–F). The most common topology is depicted in Fig. 4A, D.

Figure 4. 

An analysis of phylogenies from five mitogenome datasets. A. Maximum likelihood (ML) and Bayesian inference (BI) tree of Nereididae based on the dataset 13PCGs. The GenBank accession numbers used are listed after the species names. The scale bar (0.5) corresponds to the estimated number of substitutions per site. Numbers at nodes are statistical support values for ML bootstrap support. Asterisks denote 100% bootstrap support. “-” indicates no support value. Color-coded clades are four subfamilies within Nereididae. The gene order is shown to the right. “?” indicates the deletion of a gene; B. A two-dimensional MDS plot of eight trees (excluding 13PCGs12_ML and 13PCGs12_BI), colored by different clusters. C–F. For each cluster identified in B., a representative tree was selected. Note: Neanthes glandicincta (NC 035893) is an incorrectly identified taxon, which should be the genus Dendronereis. Neanthes acuminata (OQ729916) and Perinereis fayedensis (OQ729919) should be Perinereis suezensis and Perinereis damietta, respectively. Perinereis aibuhitensis (NC 023943) should be Perinereis linea (NC 063944).

Phylogenetic analyses based on five datasets of the available mitochondrial genomes (32 taxa in Nereididae, including 8 taxa in Perinereis) provide support for the monophyly of the genus Perinereis. Perinereis was either sister to the genus Nereis and the species Cheilonereis cyclurus (MF538532) (Fig. 4C, E, F) or sister to the genera Platynereis, Hediste, Alitta, Nectoneanthes, and the species Laeonereis culveri (KU992689) (Fig. 4D). In all trees, P. wilsoni was sister to P. vancaurica and P. nuntia with high nodal support values (BS ≥ 94%, PP = 1) (Fig. 4 and Suppl. material 6). The P. nuntia species group was found not to be monophyletic. Specifically, P. linea (NC 063944), P. aibuhitensis (NC 023943), and P. vancaurica (ON611802) have two or three paragnaths on Area VI of the pharynx, which belong to Perinereis Group 2 (Hutchings et al. 1991); all other Perinereis species (except Perinereis cultrifera) in the trees belong to the P. nuntia group. Yet, the species of Group 2 were nested within the P. nuntia group.

Phylogenetic trees (ML and BI) based on COX1 genes recovered that P. wilsoni (23007-1, 23007-2) have a closer relationship with P. mictodonta (KC800632, KC800630, KC800628), with nodal support values (BS = 83%, PP = 0.94). All taxa labeled P. wilsoni and P. mictodonta did not each form a monophyletic group, respectively; instead, they were divided into five distinct clades. This suggests the potential of cryptic species present within P. wilsoni and P. mictodonta or that the specimens were sampled from geographically distant localities with some degree of isolation. These five clades were also supported by genetic distances, which have a distinct barcode gap from each other (Fig. 5A, D, and Suppl. material 7). The phylogenetic trees (ML and BI) based on 16S genes suggested that P. wilsoni and P. mictodonta were each not monophyletic (Fig. 5B, E, and Suppl. material 8). Genetic distances also showed that the sequences of P. wilsoni and P. mictodonta had no barcode gap (Fig. 5B). The phylogenetic trees (ML and BI) based on ITS genes supported P. wilsoni as a sister to P. mictodonta (BS = 96%, PP = 1). Although P. wilsoni and P. mictodonta formed a monophyletic cluster with each other, the genetic distance analyses showed that these two species have no distinct barcode gap (Fig. 5C, F, and Suppl. material 9). Based on the dos Santos Vieira et al. (2020) methodology, optimal markers were selected based on an overlap criterion of less than 20%. However, the overlap between P. wilsoni and P. mictodonta exceeded 35%, indicating that COX1, 16S, and ITS genes are not optimal molecular markers (Fig. 5A–C).

Figure 5. 

The heatmap of COX1 A, 16S B., and ITS C. p-distance for Perinereis wilsoni and P. mictodonta, with the barcode gap and distance overlap of them on the left. Bayesian inference (BI) phylogenetic trees for the two species, based on sequences of COX1 D, 16S E, and ITS F, are excerpted from Suppl. materials 79. Sequences from P. wilsoni are depicted in black, those from P. mictodonta in green, with sequences from this study highlighted in bold.

The positions of Paraleonnates uschakovi (NC 032361) and Laeonereis culveri (KU992689) are unstable, jumping across different phylogenies (Fig. 4, Suppl. material 6). Laeonereis culveri (KU992689) is closer to P. uschakovi (NC 032361) in four trees (13PCGs_2R_ML, 13PCGs12_ML, 13PCGs12_2R_ML, and 13PCGsAA_BI in Suppl. material 6), all with high nodal support. In contrast, in six other trees, L. culveri (KU992689) is more closely related to the genera Hediste, Alitta, and Nectoneanthes, also with high nodal support, except in the 13PCGs12_ML and 13PCGs12_BI trees (Fig. 4, Suppl. material 6). Paraleonnates uschakovi is closer to the subfamilies Gymnonereidinae, Dendronereinae, and Nereidinae (8 of 10 trees), with low nodal support in ML trees but with high nodal support in BI trees. In 13PCGs12_ML and 13PCGs12_BI trees, P. uschakovi was clustered with the outgroup, all with low nodal support (Suppl. material 6).

There are two primary types of mitochondrial gene order in the known mitochondrial genomes of Nereididae, except for L. culveri (KU992689). The first type of gene order is observed in the subfamilies Gymnonereidinae, Dendronereinae, and P. uschakovi. The second type is found in the subfamily Nereidinae, except for L. culveri (see Fig. 4).

Nucleotide diversity and evolutionary rate analyses

The nucleotide diversity (Pi) analysis was conducted using concatenated alignments of 13 PCGs and two rRNAs of 32 Nereididae species. The sequence variation ratio exhibits variable nucleotide diversity between the Nereididae, with Pi values for the 100 bp windows ranging from 0.061 to 0.563 (Fig. 6A). COX1 (Pi = 0.211), COX3 (0.235), CYTB (0.238), and COX2 (0.239) exhibit a comparatively low sequence variability, whereas ATP8 (0.405), ND2 (0.386), and ND6 (0.378) have a comparatively high sequence variability. This is corroborated by the non-synonymous/synonymous (Ka/Ks) ratio analysis, which shows that COX1 (Ka/Ks = 0.036), COX3 (0.038), CYTB (0.053), and COX2 (0.065) are evolving comparatively slowly, whereas ATP8 (0.270), ND2 (0.196), and ND6 (0.173) are evolving comparatively fast (Fig. 6B). All genes are under purifying selection.

Figure 6. 

Nucleotide diversity analysis: A. Of 13 PCGs + two rRNAs and Ka/Ks rates; B. Of 13 PCGs based on 32 Nereididae species. The Pi values for the 13 PCGs + two rRNAs are shown on the graph. The red line represents the value of nucleotide diversity (Pi) (window size = 100 bp, step size = 20 bp). The pink, purple, and green columns represent the values of Ka, Ks, and Ka/Ks, respectively.

Discussion

The taxonomic status of Perinereis wilsoni

Before discussing this topic, it is crucial to specify the species concept employed in this study. Despite approximately 30 species concepts having been proposed (Hong 2020), the scientific community has not reached a consensus on defining a species. Our study adopts the Gen-morph species concept, initially proposed by Deyuan Hong (Hong 2020), which requires at least two each of independent morphological characteristics and genetic markers for species definition. According to this concept, morphological characters include quantitative morphological features that have been shown to be statistically different between species.

Perinereis wilsoni and P. mictodonta were initially established as separate species based on statistically validated morphometric differences and the results of the ITS genes (Glasby and Hsieh 2006). Subsequent studies by Park and Kim (2007) and Tosuji et al. (2019) demonstrated that COX1, 16S, and ITS genes can distinguish these two species, respectively. However, the COX1 sequences used in Park and Kim (2007) have not been made publicly available. Moreover, their study relied solely on morphological characteristics for species identification and did not validate these identifications with ITS gene analysis, casting doubt on the reliability of their findings. Our re-analysis of sequences from (Tosuji et al. 2019) showed that the 16S was insufficient to distinguish the two species. To date, only ITS genes have proven effective in distinguishing P. wilsoni and P. mictodonta (Chen et al. 2002; Glasby and Hsieh 2006; Tosuji et al. 2019). Considering the Gen-morph species concept, P. wilsoni may not be a “good” species until a second genetic marker (and a more optimal one than ITS; see below) can be found. Therefore, comprehensive specimen sampling from various geographical locations is still required to assess its taxonomic status thoroughly (Deyuan Yang et al. in preparation).

Which molecular markers suit Perinereis wilsoni Glasby & Hsieh, 2006 identification?

Partial mitochondrial genes, like COX1 and 16S, and nuclear genes, like 18S, 28S, ITS, and H3, are widely used in Nereididae for species discovery and phylogenetic studies (Glasby et al. 2013; Elgetany et al. 2022; Teixeira et al. 2022a; Teixeira et al. 2022b; Teixeira et al. 2024). Among these genes, the 18S and 28S genes are typically used in higher-taxon phylogenetic analyses. A comprehensive assessment of the effectiveness of these molecular markers in differentiating species, both closely and distantly related across various genera and families, remains limited (Halanych and Janosik 2006).

Given that species within the P. nuntia complex are morphologically similar and not easily distinguishable based on morphology alone, molecular-based identification can offer a faster and more reliable method for species identification when reliable molecular references are available. Our analyses were unable to confirm the discriminative capability of the COX1 gene between P. wilsoni and P. mictodonta, as the sequences sourced from public databases were not linked to morphological and ITS gene data, thus limiting our further study. Additionally, the 16S gene was proven to be insufficient to distinguish these two species. Although ITS genes have been found effective in distinguishing cryptic species (Chen et al. 2002; Pleijel et al. 2009; Nygren and Pleijel 2011; Nygren 2014), they may not be the optimal molecular marker in distinguishing P. wilsoni and P. mictodonta due to the lack of a distinctive barcoding gap between these two species. Furthermore, there are only a few available sequences of Polychaeta ITS genes in the public database, possibly due to difficulties in amplifying these sequences by PCR. We found that the primer sets ITS18SFPOLY and ITS28SRPOLY (Nygren et al. 2009) may be more effective than the one provided by Chen et al. (2002) for amplifying ITS genes in Perinereis.

With advances in sequencing technologies, high-throughput sequencing (HTS) is more efficient than PCR-based Sanger sequencing in obtaining molecular markers. Recently, some new molecular markers have been proposed, such as nearly universal single-copy nuclear protein-coding genes (Eberle et al. 2020) and organelle genomes (Margaryan et al. 2021). In this paper, we found the genome-skimming approach makes it easier to explore the mitogenome, complete nuclear ribosomal DNA (18S-ITS1-5.8S-ITS2-28S), and complete histone genes. However, it is more difficult to explore universal single-copy nuclear protein-coding genes, even with sequencing data increased to 25 Gb.

The phylogeny of Perinereis and the systematic position of Perinereis wilsoni

Currently, a robust phylogenetic backbone of the genus Perinereis, based on phylogenomic methods and extensive taxon sampling encompassing major species in all five informal grouping schemes proposed by Hutchings et al. (1991), is still lacking. Traditionally, the genus Perinereis has been considered polyphyletic (Bakken and Wilson 2005; Glasby et al. 2013; Alves et al. 2020; Elgetany et al. 2022). However, this view is largely based on morphological evidence or a limited number of loci and/or may include assembly errors in their datasets. In this study, although our curated mitogenome datasets support the monophyly of the genus Perinereis, the dataset includes only 8 species, whereas the genus currently comprises 103 species (Prajapat et al. 2024). Furthermore, morphological diversity within the genus has been unevenly sampled. Hutchings et al. (1991) proposed an informal grouping scheme with five groups. Our mitogenome dataset includes Group 1A, represented by 1 taxon (P. cultrifera); Group 2A, represented by 4 taxa (P. linea, P. aibuhitensis (should be P. linea), P. camiguinoides, and P. vancaurica); and Group 3A, represented by 4 taxa (P. fayedensis, N. acuminata (=P. suezensis), P. wilsoni, and P. nuntia). However, Groups 1B and 3B are not represented. To confidently establish the monophyly of the genus, future analyses should include multiple representatives from each of these informal morphological groupings.

Based on the available mitogenome datasets, Perinereis wilsoni is a sister group to P. vancaurica and P. nuntia, with high nodal support in all phylogenetic trees. In contrast, single-gene phylogenetic trees suggest that P. wilsoni is more closely related to P. mictodonta, with low nodal support. Although including more taxa, single-gene trees do not provide sufficient resolution in phylogeny. In summary, the phylogenetic relationships of P. wilsoni remain poorly understood due to the limited number of Perinereis species or other Nereididae for which genomic data are available.

The phylogeny of Nereididae

A deep discussion of the phylogeny of Nereididae is beyond the scope of our current work. Here, we provide a brief discussion. The positions of P. uschakovi and L. culveri (KU992689) were observed to be unstable in the trees, inferred from different mitogenome datasets in this study.

In detail, L. culveri always nested within the subfamily Nereidinae, which was also found in previous studies using different datasets, such as COX1, 16S, and 18S (Wang et al. 2021; Alves et al. 2023; Villalobos-Guerrero et al. 2024), as well as mitogenome datasets (Alves et al. 2020). L. culveri is also a sister group to the Gymnonereidinae, Dendronereinae, and Nereidinae subfamilies, along with P. uschakovi (this study; Villalobos-Guerrero et al. 2024). P. uschakovi was always found as a sister group to other subfamilies (Alves et al. 2020; Wang et al. 2021; Alves et al. 2023; Villalobos-Guerrero et al. 2024; this study). However, it occasionally grouped with the outgroup Craseoschema thyasiricola (NC 060815, see Suppl. material 6: 13PCGs12_ML tree and 13PCGs12_BI tree) or formed a single clade with the Gymnonereidinae and Dendronereinae subfamilies (see Suppl. material 6: 13PCGsAA_ML tree), which appears to be supported by mitochondrial genome gene order. Alves et al. (2020) also found this situation when removing outgroups. Inspired by a morphological phylogenetic tree proposed by Wu et al. (1981), which is based on the characteristics of pharyngeal armature, we propose a hypothesis that Paraleonnates and Laeonereis (KU992689) may represent early branching members of Nereidinae. The lack of sufficient data on such taxa may be causing their uncertain placements within the phylogenetic framework.

In this study, we uncovered potential errors in assembly and annotation within GenBank. However, the absence of corresponding Sequence Read Archive (SRA) data in public databases hampers accurate confirmation of these errors. Consequently, we filtered these sequences based solely on our expertise. Therefore, I (Deyuan Yang) advocate for the uploading of original data (raw data) to public databases, such as NCBI. Even if some authors are unwilling to upload their data, various assembly methods should be employed to ensure the accuracy of the assemblies.

Additionally, we emphasize that carefully curated datasets (verifying the taxonomic identification of the species used in the phylogenetic study), especially those from public databases, are crucial before conducting phylogenetic studies, as taxonomic misidentifications can lead to incorrect conclusions. For example, Alves et al. (2020) concluded that Gymnonereidinae was non-monophyletic, a finding that was impacted by the inclusion of an incorrectly identified taxon, Neanthes glandicincta (NC 035893), which should have been classified under the genus Dendronereis Peters, 1854 (Zhen et al. 2022; this study). Additionally, in this study, if we had not questioned the name Neanthes acuminata (OQ729916), we would have concluded that Perinereis is non-monophyletic. However, N. acuminata should be re-identified as Perinereis suezensis (Elgetany et al. 2022).

Author Contribution

D.Y.Y. and S.Z. wrote the manuscript, and S.Z. wrote the part on mitochondrial genes, codon usage, nucleotide diversity, and evolutionary rate analyses. D.Y.Y. and S.Z. analyzed the data. Z.W., Y.J.Z., and D.Z.Y. participated in the discussion and reviewed the manuscript. C.J.G., J.S.H., and L.Z.C. conceived and designed, supervised the work, and reviewed drafts of the paper. S.Z. and D.Y.Y. contributed equally to this manuscript. All authors have read and agreed to the published version of the manuscript.

Acknowledgments

Many thanks to Yuanzheng Meng for helping us organize and format the literature for this paper. Thanks to Grammarly (https://www.grammarly.com/) for grammar correction while writing the first manuscript and ChatGPT 4.0 for generating some Python and R scripts during this study. Thanks to reviewer Robin Wilson for his constructive comments, especially focused on the phylogeny of the genus Perinereis, which have greatly contributed to the revision of our article. This work was supported by the Youth Fund of the National Natural Science Foundation of China (42306107) and the China Postdoctoral Science Foundation (2021M691866).

References

  • Allio R, Schomaker‐Bastos A, Romiguier J, Prosdocimi F, Nabholz B, Delsuc F (2020) MitoFinder: Efficient automated large‐scale extraction of mitogenomic data in target enrichment phylogenomics. Molecular Ecology Resources 20(4): 892–905. https://doi.org/10.1111/1755-0998.13160
  • Alves PR, Halanych KM, Santos CSG (2020) The phylogeny of Nereididae (Annelida) based on mitochondrial genomes. Zoologica Scripta 49(3): 366–378. https://doi.org/10.1111/zsc.12413
  • Arias A, Richter A, Anadón N, Glasby CJ (2013) Revealing polychaetes invasion patterns: Identification, reproduction and potential risks of the Korean ragworm, Perinereis linea (Treadwell), in the Western Mediterranean. Estuarine, Coastal and Shelf Science 131: 117–128. https://doi.org/10.1016/j.ecss.2013.08.017
  • Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, Lesin VM, Nikolenko SI, Pham S, Prjibelski AD (2012) SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. Journal of Computational Biology 19(5): 455–477. https://doi.org/10.1089/cmb.2012.0021
  • Boisvert S, Laviolette F, Corbeil J (2010) Ray: simultaneous assembly of reads from a mix of high-throughput sequencing technologies. Journal of Computational Biology 17(11): 1519–1533. https://doi.org/10.1089/cmb.2009.0238
  • Capella-Gutiérrez S, Silla-Martínez JM, Gabaldón T (2009) trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25(15): 1972–1973. https://doi.org/10.1093/bioinformatics/btp348
  • Chen CA, Chen C-P, Fan T-Y, Yu J-K, Hsieh H-L (2002) Nucleotide sequences of ribosomal internal transcribed spacers and their utility in distinguishing closely related Perinereis polychaets (Annelida; Polychaeta; Nereididae). Marine Biotechnology 4: 17–29. https://doi.org/10.1007/s10126-001-0069-3
  • Chen C, Chen H, Zhang Y, Thomas HR, Frank MH, He Y, Xia R (2020) TBtools: an integrative toolkit developed for interactive analyses of big biological data. Molecular Plant 13(8): 1194–1202. https://doi.org/10.1016/j.molp.2020.06.009
  • Coissac E, Hollingsworth PM, Lavergne S, Taberlet P (2016) From barcodes to genomes: extending the concept of DNA barcoding. Molecular Ecology 25(7): 1423–1428. https://doi.org/10.1111/mec.13549
  • Dierckxsens N, Mardulyn P, Smits G (2017) NOVOPlasty: de novo assembly of organelle genomes from whole genome data. Nucleic Acids Research 45(4): e18-e18. https://doi.org/10.1093/nar/gkw955
  • Donath A, Jühling F, Al-Arab M, Bernhart SH, Reinhardt F, Stadler PF, Middendorf M, Bernt M (2019) Improved annotation of protein-coding genes boundaries in metazoan mitochondrial genomes. Nucleic Acids Research 47(20): 10543–10552. https://doi.org/10.1093/nar/gkz833
  • dos Santos Vieira WA, Bezerra PA, da Silva AC, Veloso JS, Câmara MPS, Doyle VP (2020) Optimal markers for the identification of Colletotrichum species. Molecular Phylogenetics and Evolution 143: 106694. https://doi.org/10.1016/j.ympev.2019.106694
  • Eberle J, Ahrens D, Mayer C, Niehuis O, Misof B (2020) A plea for standardized nuclear markers in metazoan DNA taxonomy. Trends in Ecology & Evolution 35(4): 336–345. https://doi.org/10.1016/j.tree.2019.12.003
  • Glasby CJ (1993) Family revision and cladistic analysis of the Nereidoidea (Polychaeta: Phyllodocida). Invertebrate Systematics 7(6): 1551–1573. https://doi.org/10.1071/IT9931551
  • Glasby CJ, Wei N-WV, Gibb KS (2013) Cryptic species of Nereididae (Annelida: Polychaeta) on Australian coral reefs. Invertebrate Systematics 27(3): 245–264. https://doi.org/10.1071/IS12031
  • Hahn C, Bachmann L, Chevreux B (2013) Reconstructing mitochondrial genomes directly from genomic next-generation sequencing reads—a baiting and iterative mapping approach. Nucleic Acids Research 41(13): e129-e129. https://doi.org/10.1093/nar/gkt371
  • Halanych KM, Janosik AM (2006) A review of molecular markers used for Annelid phylogenetics. Integrative and Comparative Biology 46(4): 533–543. https://doi.org/10.1093/icb/icj052
  • Hektoen MM, Bakken T, Ekrem T, Radashevsky VI, Dunshea G (2024) Species delimitation and phylogenetic relationships of the Prionospio complex (Annelida, Spionidae) in the Northeast Atlantic. Zoologica Scripta: 358–375. https://doi.org/10.1111/zsc.12648
  • Hong DY (2020) Gen‐morph species concept—A new and integrative species concept for outbreeding organisms. Journal of systematics and evolution 58(5): 725–742. https://doi.org/10.1111/jse.12660
  • Hutchings P, Reid A, Wilson R (1991) Perinereis (Polychaeta, Nereididae) from Australia, with redescriptions of six additional species. Records of the Australian Museum 43(3): 241–274. https://doi.org/10.3853/j.0067-1975.43.1991.47
  • Jin J-J, Yu W-B, Yang J-B, Song Y, DePamphilis CW, Yi T-S, Li D-Z (2020) GetOrganelle: a fast and versatile toolkit for accurate de novo assembly of organelle genomes. Genome Biology 21: 1–31. https://doi.org/10.1186/s13059-020-02154-5
  • Jombart T, Kendall M, Almagro‐Garcia J, Colijn C (2017) treespace: Statistical exploration of landscapes of phylogenetic trees. Molecular Ecology Resources 17(6): 1385–1392. https://doi.org/10.1111/1755-0998.12676
  • Kumar S, Stecher G, Li M, Knyaz C, Tamura K (2018) MEGA X: molecular evolutionary genetics analysis across computing platforms. Molecular Biology and Evolution 35(6): 1547. https://doi.org/10.1093/molbev/msy096
  • Letunic I, Bork P (2021) Interactive Tree Of Life (iTOL) v5: an online tool for phylogenetic tree display and annotation. Nucleic Acids Research 49(W1): W293-W296. https://doi.org/10.1093/nar/gkab301
  • Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, Marth G, Abecasis G, Durbin R, Subgroup GPDP (2009) The sequence alignment/map format and SAMtools. Bioinformatics 25(16): 2078–2079. https://doi.org/10.1093/bioinformatics/btp352
  • Li D, Liu C-M, Luo R, Sadakane K, Lam T-W (2015) MEGAHIT: an ultra-fast single-node solution for large and complex metagenomics assembly via succinct de Bruijn graph. Bioinformatics 31(10): 1674–1676. https://doi.org/10.1093/bioinformatics/btv033
  • Mahcene HR, Villalobos-Guerrero TF, Kurt G, Denis F, Daas T (2023) A new species of Perinereis Kinberg, 1865 (Annelida: Nereididae) from the Western Mediterranean Sea revealed by morphological and molecular approaches. Mediterranean Marine Science 24(2): 454–460. https://doi.org/10.12681/mms.33969
  • Margaryan A, Noer CL, Richter SR, Restrup ME, Bülow‐Hansen JL, Leerhøi F, Langkjær EMR, Gopalakrishnan S, Carøe C, Gilbert MTP (2021) Mitochondrial genomes of Danish vertebrate species generated for the national DNA reference database, DNAmark. Environmental DNA 3(2): 472–480. https://doi.org/10.1002/edn3.138
  • Meng G, Li Y, Yang C, Liu S (2019) MitoZ: a toolkit for animal mitochondrial genome assembly, annotation and visualization. Nucleic Acids Research 47(11): e63. https://doi.org/10.1093/nar/gkz173
  • Nygren A, Pleijel F (2011) From one to ten in a single stroke–resolving the European Eumida sanguinea (Phyllodocidae, Annelida) species complex. Molecular Phylogenetics and Evolution 58(1): 132–141. https://doi.org/10.1016/j.ympev.2010.10.010
  • Park T-S, Kim W (2007) A taxonomic study on Perinereis nuntia species group (Polychaeta: Nereididae) of Korea. Animal Systematics, Evolution and Diversity 23(1): 75–85. https://doi.org/10.5635/KJSZ.2007.23.1.075
  • Peng Y, Leung HC, Yiu S-M, Chin FY (2012) IDBA-UD: a de novo assembler for single-cell and metagenomic sequencing data with highly uneven depth. Bioinformatics 28(11): 1420–1428. https://doi.org/10.1093/bioinformatics/bts174
  • Perna NT, Kocher TD (1995) Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes. Journal of molecular evolution 41: 353–358. https://doi.org/10.1007/BF01215182
  • Pleijel F, Rouse G, Nygren A (2009) Five colour morphs and three new species of Gyptis (Hesionidae, Annelida) under a jetty in Edithburgh, South Australia. Zoologica Scripta 38(1): 89–99. https://doi.org/10.1111/j.1463-6409.2008.00356.x
  • Prajapat V, Villalobos-Guerrero TF, Vachhrajani KD (2023) A new species of Perinereis Kinberg, 1865 (Annelida: Nereididae) and invalidation of two congeners from Western India. Zootaxa 5330(3): 398–412. https://doi.org/10.11646/zootaxa.5330.3.4
  • Prajapat V, Villalobos-Guerrero TF, Vachhrajani KD (2024) Two new species of Perinereis Kinberg, 1865 (Annelida: Nereididae) from the rocky shore of Maharashtra, India, including notes and an identification key to Group 1. European Journal of Taxonomy 935: 256–282. https://doi.org/10.5852/ejt.2024.935.2561
  • R Core Team (2023) R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. https://www.R-project.org/
  • Ranwez V, Douzery EJ, Cambon C, Chantret N, Delsuc F (2018) MACSE v2: toolkit for the alignment of coding sequences accounting for frameshifts and stop codons. Molecular Biology and Evolution 35(10): 2582–2584. https://doi.org/10.1093/molbev/msy159
  • Richter S, Schwarz F, Hering L, Böggemann M, Bleidorn C (2015) The utility of genome skimming for phylogenomic analyses as demonstrated for glycerid relationships (Annelida, Glyceridae). Genome Biology and Evolution 7(12): 3443–3462. https://doi.org/10.1093/gbe/evv224
  • Ronquist F, Teslenko M, Van Der Mark P, Ayres DL, Darling A, Höhna S, Larget B, Liu L, Suchard MA, Huelsenbeck JP (2012) MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Systematic Biology 61(3): 539–542. https://doi.org/10.1093/sysbio/sys029
  • Rozas J, Ferrer-Mata A, Sánchez-DelBarrio JC, Guirao-Rico S, Librado P, Ramos-Onsins SE, Sánchez-Gracia A (2017) DnaSP 6: DNA sequence polymorphism analysis of large data sets. Molecular Biology and Evolution 34(12): 3299–3302. https://doi.org/10.1093/molbev/msx248
  • Simão FA, Waterhouse RM, Ioannidis P, Kriventseva EV, Zdobnov EM (2015) BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics 31(19): 3210–3212. https://doi.org/10.1093/bioinformatics/btv351
  • Song M-H, Yan C, Li J-T (2022) MEANGS: an efficient seed-free tool for de novo assembling animal mitochondrial genome using whole genome NGS data. Briefings in Bioinformatics 23(1): bbab538. https://doi.org/10.1093/bib/bbab538
  • Steenwyk JL, Buida III TJ, Gonçalves C, Goltz DC, Morales G, Mead ME, LaBella AL, Chavez CM, Schmitz JE, Hadjifrangiskou M (2022) BioKIT: a versatile toolkit for processing and analyzing diverse types of sequence data. Genetics 221(3): iyac079. https://doi.org/10.1093/genetics/iyac079
  • Straub SC, Parks M, Weitemier K, Fishbein M, Cronn RC, Liston A (2012) Navigating the tip of the genomic iceberg: Next‐generation sequencing for plant systematics. American Journal of Botany 99(2): 349–364. https://doi.org/10.3732/ajb.1100335
  • Struck TH, Purschke G, Halanych KM (2006) Phylogeny of Eunicida (Annelida) and exploring data congruence using a partition addition bootstrap alteration (PABA) approach. Systematic Biology 55(1): 1–20. https://doi.org/10.1080/10635150500354910
  • Teixeira MA, Bakken T, Vieira PE, Langeneck J, Sampieri BR, Kasapidis P, Ravara A, Nygren A, Costa FO (2022a) The curious and intricate case of the European Hediste diversicolor (Annelida, Nereididae) species complex, with description of two new species. Systematics and Biodiversity 20(1): 1–39. https://doi.org/10.1080/14772000.2022.2116124
  • Teixeira MA, Langeneck J, Vieira PE, Hernández JC, Sampieri BR, Kasapidis P, Mucciolo S, Bakken T, Ravara A, Nygren A (2022b) Reappraisal of the hyperdiverse Platynereis dumerilii (Annelida: Nereididae) species complex in the Northern Atlantic, with the description of two new species. Invertebrate Systematics 36(11): 1017–1061. https://doi.org/10.1071/IS21084
  • Teixeira MA, Fourreau CJL, Sampere-Valverde J, Carvalho S (2024) Two new records and description of a new Perinereis (Annelida, Nereididae) species for the Saudi Arabian Red Sea region. ZooKeys 1196: 331–354. https://doi.org/10.3897/zookeys.1196.115260
  • Tilic E, Stiller J, Campos E, Pleijel F, Rouse GW (2022) Phylogenomics resolves ambiguous relationships within Aciculata (Errantia, Annelida). Molecular Phylogenetics and Evolution 166: 107339. https://doi.org/10.1016/j.ympev.2021.107339
  • Tosuji H, Nishinosono K, Hsieh H-L, Glasby CJ, Sakaguchi T, Sato M (2019) Molecular evidence of cryptic species diversity in the Perinereis nuntia species group (Annelida: Nereididae) with first records of P. nuntia and P. shikueii in southern Japan. Plankton and Benthos Research 14(4): 287–302. https://doi.org/10.3800/pbr.14.287
  • Tosuji H, Park T, Goryo Y, Kan K, Abe H, Sato M (2023) Molecular method to identify the morphologically similar four species of the Perinereis nuntia species complex (Annelida: Nereididae) based on PCR-RFLP analysis of nuclear ribosomal ITS, with new distributional records of the two forms of Perinereis shikueii. Plankton and Benthos Research 18(3): 124–130. https://doi.org/10.3800/pbr.18.124
  • Villalobos-Guerrero TF (2019) Redescription of two overlooked species of the Perinereis nuntia complex and morphological delimitation of P. nuntia (Savigny in Lamarck, 1818) from the Red Sea (Annelida, Nereididae). Zoosystema 41(1): 465–496. https://doi.org/10.5252/zoosystema2019v41a24
  • Wang Y, Cheng H, Wang C (2021) A new eyeless species of Nicon (Annelida: Nereididae) from the deep Northwest Pacific Ocean. Acta Oceanologica Sinica 40(12): 20–26. https://doi.org/10.1007/s13131-021-1886-z
  • Wu B, Sun R, Yang D (1981) The Nereidae (Poluchaetous Annelids) of the Chinese Coast. Ocean Press, Beijing, 228 pp.
  • Yang D, Wu X, Wang Z, Zhao X, Hwang J, Cai L (2022) Redescription of a rarely encountered species Travisachinensis Grube, 1869 (Annelida, Travisiidae), including a description of a new species of Travisa from Amoy, China. ZooKeys 1128: 1–17. https://doi.org/10.3897/zookeys.1128.90020
  • Zhang D, Gao F, Jakovlić I, Zou H, Zhang J, Li WX, Wang GT (2020) PhyloSuite: An integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Molecular ecology resources 20(1): 348–355. https://doi.org/10.1111/1755-0998.13096
  • Zhen W, Wu X, Hao E, Xu W, Deng J, Zhu J, Xu Y (2022) The first complete mitochondrial genome of the genus Dendronereis, represented by D. chipolini Hsueh, 2019 (Annelida, Nereididae) from Beibu Gulf, China. Mitochondrial DNA Part B 7(2): 393–395. https://doi.org/10.1080/23802359.2022.2040389

Supplementary materials

Supplementary material 1 

The reports from FastQC

Deyuan Yang, Sheng Zeng, Zhi Wang, Yanjie Zhang, Dazuo Yang, Christopher J. Glasby, Jiang-Shiou Hwang, Lizhe Cai

Data type: pdf

Explanation note: (a) Basic information on clean data, including duplicate reads (%, Dups), average GC content (%, GC), and total sequences (millions, M Seqs). (b) Sequence counts for each sample. Duplicate read counts are an estimate only.

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (498.83 kb)
Supplementary material 2 

The report of Quast

Deyuan Yang, Sheng Zeng, Zhi Wang, Yanjie Zhang, Dazuo Yang, Christopher J. Glasby, Jiang-Shiou Hwang, Lizhe Cai

Data type: pdf

Explanation note: (a) Basic information of various assemblers. (b) The cumulative length of each assembler. All statistics are based on contigs of size >= 500 bp, unless otherwise noted (e.g., "# contigs (>= 0 bp)" and "Total length (>= 0 bp)" include all contigs).

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (1.08 MB)
Supplementary material 3 

Additional information

Deyuan Yang, Sheng Zeng, Zhi Wang, Yanjie Zhang, Dazuo Yang, Christopher J. Glasby, Jiang-Shiou Hwang, Lizhe Cai

Data type: xlsx

Explanation note: table S1. The Blast results of mitochondrial gene, 18S, 28S, ITS, and histone genes (Only 20 sequences are shown). table S2. List of 32 species and two outgroups used in this paper. table S3. Sequence Information from NCBI BLAST Analysis of the COX1 Gene. table S4. Sequence Information from NCBI BLAST Analysis of the 16S Gene. table S5. Original and Gblock lengths of the PCGs and PCGsAA sequences. table S6. Best partitioning schemes and models based on different datasets for maximum likelihood and Bayesian inference analysis. table S7. COI, 16S, 18S, 28S, ITS, and H3 gene sequences information of Perinereis. table S8. Nucleotide composition and skewness comparison of different elements of the mitochondrial genomes of P. wilsoni. table S9. Features of the P. wilsoni mitogenome. table S10. Codon numbers and relative synonymous codon usage (RSCU) of 13 PCGs in the P. wilsoni mitogenome. table S11. Summary statistics for multiple alignment of various datasets.

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (134.97 kb)
Supplementary material 4 

The Python script for categorizing data

Deyuan Yang, Sheng Zeng, Zhi Wang, Yanjie Zhang, Dazuo Yang, Christopher J. Glasby, Jiang-Shiou Hwang, Lizhe Cai

Data type: docx

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (17.96 kb)
Supplementary material 5 

Heterogeneity of sequence composition of mitochondrial genomes for 5 different data sets

Deyuan Yang, Sheng Zeng, Zhi Wang, Yanjie Zhang, Dazuo Yang, Christopher J. Glasby, Jiang-Shiou Hwang, Lizhe Cai

Data type: pdf

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (1.73 MB)
Supplementary material 6 

10 trees from different datasets and tree-building methods

Deyuan Yang, Sheng Zeng, Zhi Wang, Yanjie Zhang, Dazuo Yang, Christopher J. Glasby, Jiang-Shiou Hwang, Lizhe Cai

Data type: pdf

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (852.81 kb)
Supplementary material 7 

Phylogenetic trees of Perinereis based on the COX1 dataset

Deyuan Yang, Sheng Zeng, Zhi Wang, Yanjie Zhang, Dazuo Yang, Christopher J. Glasby, Jiang-Shiou Hwang, Lizhe Cai

Data type: pdf

Explanation note: (a) Bayesian inference (BI, left) and (b) maximum likelihood (ML, right) method

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (503.68 kb)
Supplementary material 8 

Phylogenetic trees of Perinereis based on the 16S dataset

Deyuan Yang, Sheng Zeng, Zhi Wang, Yanjie Zhang, Dazuo Yang, Christopher J. Glasby, Jiang-Shiou Hwang, Lizhe Cai

Data type: pdf

Explanation note: (a) Bayesian inference (BI, left) and (b) maximum likelihood (ML, right) methods.

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (325.75 kb)
Supplementary material 9 

Phylogenetic trees of Perinereis based on the ITS dataset

Deyuan Yang, Sheng Zeng, Zhi Wang, Yanjie Zhang, Dazuo Yang, Christopher J. Glasby, Jiang-Shiou Hwang, Lizhe Cai

Data type: pdf

Explanation note: (a) maximum likelihood (ML, left) and (b) Bayesian inference (BI, right) methods.

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (238.84 kb)
login to comment