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

Phylogenetic analysis and ontogenetic changes in the cone opsins of the western mosquitofish (Gambusia affinis)

  • Chia-Hao Chang,

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Resources, Software, Writing – original draft, Writing – review & editing

    Affiliations TIGP Biodiversity Program, Tunghai University, Taichung City, Taiwan, Center for Ecology and Environment, Tunghai University, Taichung City, Taiwan

  • Yu-Chun Wang,

    Roles Formal analysis, Investigation, Resources

    Affiliation Planning and Information Division, Fisheries Research Institute, Keelung City, Taiwan

  • Yi Ta Shao,

    Roles Investigation, Methodology, Resources, Software

    Affiliation Institute of Marine Biology, National Taiwan Ocean University, Keelung, Taiwan

  • Shih-Hui Liu

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Software, Writing – review & editing

    liushihhui@gmail.com, liushihhui@mail.nsysu.edu.tw

    Affiliation Department of Biological Sciences, National Sun Yat-sen University, Kaohsiung City, Taiwan

Abstract

To convert external light into internal neural signal, vertebrates rely on a special group of proteins, the visual opsins. Four of the five types of visual opsins—short-wavelength sensitive 1 (Sws1), short-wavelength sensitive 2 (Sws2), medium-wavelength sensitive (Rh2), and long-wavelength sensitive (Lws)—are expressed in cone cells for scotopic vision, with the fifth, rhodopsin (Rh1), being expressed in rod cells for photopic vision. Fish often display differing ontogenetic cone opsin expression profiles, which may be related to dietary and/or habitat ontogenetic shift. The western mosquitofish (Gambusia affinis) is an aggressive invader that has successfully colonized every continent except Antarctica. The strong invasiveness of this species may be linked to its visual acuity since it can inhabit turbid waters better than other fishes. By genome screening and transcriptome analysis, we identify seven cone opsin genes in the western mosquitofish, including one sws1, two sws2, one rh2, and three lws. The predicted maximal absorbance wavelength (λmax) values of the respective proteins are 353 nm for Sws1, 449 nm for Sws2a, 408 nm for Sws2b, 516 nm for Rh2-1, 571 nm for Lws-1, and 519 nm for Lws-3. Retention of an intron in the lws-r transcript likely renders this visual opsin gene non-functional. Our real-time quantitative PCR demonstrates that adult male and female western mosquitofish do not differ in their cone opsin expression profiles, but we do reveal an ontogenetic shift in cone opsin expression. Compared to adults, larvae express proportionally more sws1 and less lws-1, suggesting that the western mosquitofish is more sensitive to shorter wavelengths in the larval stage, but becomes more sensitive to longer wavelengths in adulthood.

Introduction

Vertebrates rely on a group of specialized photoreceptor cells in the retina of eyes to convert external light into neural signals. The visual pigment located in the photoreceptor outer segment is responsible for deciphering light information and it comprises a light-absorbing protein (visual opsin) and a vitamin A-derived chromophore. Visual opsins are membrane-bound G-protein coupled receptors (GPCRs), and vitamin A-derived chromophores include vitamin A1 (retinal) or A2 (3,4-dehydroretinal). Differences in phylogenetic classes of visual opsin genes or in the types of chromophore can alter the maximal absorbance wavelength (λmax) of visual pigments [1]. Photoreceptor cells can be categorized as rod cells (scotopic or low-light vision) or cone cells (photopic or well-lit vision), with cone cells being further divided into single-, double-, twin-, triple-, or quadruple-cone cells [2]. Based on their expression of visual opsin genes, photoreceptor cells can also be categorized into five main classes: Rh1, Sws1, Sws2, Rh2, and Lws [3]. Most vertebrates only express one kind of rhodopsin rh1 gene in rods [4], and the other four classes of visual opsins are all expressed in cones with different spectral sensitivities: Sws1 is sensitive to UV and violet light, Sws2 to blue, Rh2 to green, and Lws to red [1].

Habitat has a profound effect on the visual system of the animals living in it [5, 6]. For example, the cichlids in Lake Victoria have distinct lws alleles with different λmax values that correspond to varied lighting environments with depth [7], and the opsin expression profiles of bluefin killifish (Lucania goodei) differ between clear springs and tannin-stained swamps [8, 9]. Apart from their lighting environments, habitats can also exert a selection pressure on visual systems in other ways. For example, predation risk variation in guppy has been demonstrated to alter lws allele frequencies and opsin expression profiles among populations [10].

Aquatic habitats provide more diverse lighting environments than terrestrial habitats since water itself selectively absorbs light and suspended or dissolved substances can absorb and/or scatter light [11, 12]. Multiple independent opsin gene duplication events have occurred over the evolutionary history of teleosts [13]. For example, the haplochromine cichlids have up to six cone opsins with distinct λmax values, including one sws1, two sws2, two rh2, and one lws [14]. Notably even for animals with multiple cone opsins, these opsins are not expressed equally or constantly [1517]. Lighting environments, diurnal periodicity, and diet all contribute to changes in cone opsin expression profiles [6, 15, 16]. Many fishes exhibit the ability to adjust their cone opsin expression profiles, termed “phenotypic plasticity”. One type of phenotypic plasticity is changes in expression profile as an individual grows. Such ontogenetic changes in opsin profiles have been reported for giant mottled eel (Anguilla marmorata), milkfish (Chanos chanos), barfin flounder (Verasper moseri), and spotted unicornfish (Naso brevirostris) [1720]. Two other types of phenotypic plasticity are developmental and adult plasticity; with respect to opsins, the former pertains to how different rearing conditions can result in distinct cone opsin expression profiles [21, 22], and the latter explains how adult fish can alter their expression profiles to adapt to local lighting environments [2325]. Both developmental and adult plasticity may enable fishes to flourish in turbid waters [22, 25].

Currently, the genomic arrangement of cone opsin genes has been fully investigated for two poeciliid fishes, i.e., guppy and green swordtail (Xiphophorous helleri) [26, 27]. Both of these poeciliids possess nine cone opsin genes, including one sws1, two sws2, two rh2, and four lws. The duplicated opsin paralogs are tandemly arranged, except for lws-r, with the two sws2, three lws, and two rh2 opsins forming a linkage group [28]. The evolutionary trajectories of lws genes in poeciliid fishes reflect interactions between genomic organization and sexual selection [29].

Not all poeciliid fishes have had their cone opsin genes investigated. Although the western mosquitofish (Gambusia affinis) (Poeciliidae, Cyprinodotiformes) has been introduced in many countries to control mosquito populations since the early 1900s [30], becoming one of the worst invasive species in the process [31], our understanding of its cone opsins remains limited [32]. To date, research has only demonstrated that the western mosquitofish possesses all four phylogenetic classes of cone opsin genes, including two sws2 genes (sws2a and sws2b), but exactly how many cone opsins it has remains uncertain [32]. Moreover, previous studies indicate that the western mosquitofish may display ontogenetic changes in cone opsin expression profiles because it exhibits an ontogenetic dietary shift, colors are critical to it in the mating choice of both sexes, and sex steroid estrogen influences its cone opsin expression profiles [3236].

Sensory-based approaches can be applied to protecting endangered species and controlling alien ones [37]. A comprehensive understanding of the visual spectral characteristics of western mosquitofish could reveal how this species is so aggressively invasive and provide insights into the sensory biology of other invasive species. Here, we sought to determine how many cone opsin genes are present in the western mosquitofish and to predict the λmax values of the respective opsins. Moreover, since longer-wavelength signals, i.e., yellow, are more attractive to adult individuals [34, 35], we also hypothesized that the western mosquitofish would display an ontogenetic change in cone opsin expression profiles, with the visual spectrum shifting to red during development.

Materials and methods

Overall, a total of 84 western mosquitofish were used in this study. One adult specimen was used for transcriptome analysis, ten adult specimens were used for cloning cone opsin genes, 30 adult specimens were used for reproductive experiments, and the resulting 43 laboratory-raised specimens (comprising 30 larva, seven adult males, and six adult females) were used to quantify ontogenetic cone expression.

Fish collection and experimental setup

Western mosquitofish samples were caught using a hand-net at two different localities, Tunghai Lake at Tunghai University (Taichung City, Taiwan) and Linyuan Ocean Wetland Park (Kaohsiung City, Taiwan). Adult fish were caught from 9:00 to 11:00 AM and then euthanized using 0.025% MS-222 (Ethyl 3-aminobenzoate, methanesulphonic acid salt). After each specimen was rendered comatose, its two eyes were removed and stored in RNAlater solution (AM7020, Invitrogen), with the rest of the body being preserved in 100% ethanol at -20˚C until DNA extraction. Eye samples for cone opsin gene cloning and RNA-seq were first kept at 4˚C for 24 hours and then stored at -80 ˚C until RNA extraction. Overall, three males and three females from Tunghai University and two males and two females from Linyuan Ocean Wetland Park were prepared for DNA and RNA extraction.

A further 30 live adult specimens from Tunghai University were transported to the Marine Research Station of Academia Sinica (Yilan, Taiwan). These specimens were housed indoors in a glass aquarium (90 x 40 x 40 cm length:width:height) with continuous dripping influx of dechlorinated tap water to maintain water conditions as clean as possible. A full-spectrum LED grow light (Hydroponics Lin Agritech) was used as light source and placed approximately 28 cm above the water’s surface. The photoperiod of the climate-controlled room was set to 14:10 h light: dark, and the spectral irradiance of the light source was measured using a spectrometer (MK350N, Gamma Scientific) (S1 Fig). Fish were fed ad libitum with artificial fish feed (Otohime B2, Marubeni Nisshin Feed) twice a day. After one month, we prepared three floating fish-breeding cages in the aquarium to host pregnant females and to collect F1 offspring. When we had collected more than 50 F1 individuals, all parental individuals were removed from the aquarium and all F1 offspring were released from the cages into the main aquarium. These F1 offspring then mated to produce F2 individuals under the same lighting conditions as their parents. The floating fish-breeding cages were again employed for pregnant F1 females. The F1 adults and newborn F2 larvae (less than 10 days old) were used for ontogenetic study. In order to limit the effect of diurnal variation in opsin expression, fish were euthanized (0.025% MS-222) between 10:00 and 11:00 AM. After each specimen was rendered comatose, larval specimens were directly preserved in RNAlater solution (AM7020, Invitrogen) whereas, for adult specimens, body weight and standard length were first determined before removing both eyes and storing them in RNAlater solution (AM7020, Invitrogen). These samples were also kept at 4˚C for 24 hours before being transferred to -80˚C until RNA was extracted. All experiments were performed and specimens handled with approval (107–28) from the Institutional Animal Care and Use Committee (IACUC) of Tunghai University.

DNA and RNA extraction, and cDNA preparation

DNA samples were extracted from fin tissues using a DNA extraction kit (Cat No./ID: GS100, Geneaid). An RNeasy Plus Universal Kit (Cat No./ID: 73404, QIAGEN) was used to isolate total RNA according to the manufacturer’s protocol. One eyeball from each adult specimen or five larval heads as a pooled sample was placed in a 2 ml microcentrifuge tube with stainless steel beads for tissue homogenization. Cone opsin genes are exclusively expressed in the retina, so RNA extraction from whole larval heads does not bias quantification of cone opsins [38]. The tissue was homogenized using a Roche MagNA Lyser system (Roche). Total RNA content and quality were measured using a NanoDrop 1000 (Thermo Scientific). Two micrograms (μg) of total RNA were reverse-transcribed using a Verso cDNA Synthesis Kit (Cat No. 00764129, ThermoFisher Scientific) with Random Hexamer and Anchored Oligo-dT 3:1 (v/v) in a final volume of 20 μl. The remaining RNA was preserved at -80˚C after adding RNase inhibitor (RG90925, Lucigen).

Transcriptomics

A female specimen collected from Tunghai University was subjected to RNA sequencing (RNA-seq) by Genomics Tech. (Taipei, Taiwan). The mRNA was purified from total RNA using poly-T oligo-attached magnetic beads and then fragmented by heating. The RNA library was prepared with the TruSeq® Stranded mRNA Library Prep system (Illumina, San Diego, CA, USA). An Agilent Bioanalyzer 2100 system and a DNA High Sensitivity Chip was employed for quality control. Then, the RNA library was sequenced on an Illumina NovaSeq 6000 platform using a 150-bp paired-end strategy with an output of 6 Gb.

The raw reads were trimmed using Trimmomatic v0.36 [39] and then de-novo assembled in Trinity v2.8.4 [40]. Identical transcripts were eliminated using CD‐HIT‐EST v4.8 [41, 42], with a 95% clustering cut-off. Then, the trimmed reads were aligned onto the transcripts to estimate the sequencing depth of each transcript using Bowtie2 v2.3.4.3 [43] and RSEM v1.2.28 [44]. The average Fragments Per Kilobase Million (FPKM) was 52.67. Open reading frames (ORFs) were predicted in Transdecoder v5.3.0 (https://github.com/TransDecoder/TransDecoder) and annotated in BLASTX v2.5.0 [45]. Finally, the transcripts were translated into amino acid sequences and annotated using BLASTP v2.5.0 [45].

Identification of cone opsin genes

Known guppy cone opsin genes––sws1, sws2a, sws2b, rh2, and lws (Table 1)––were used as references to query and annotate the cone opsin genes from the western mosquitofish draft genome [46] and our draft transcriptome using BLASTN 2.8.1+ [45]. We designed specific pairs of primers for each of the identified cone opsin genes. Furthermore, we aligned the complete β-actin gene sequence of guppy (EU143771) and the partial β-actin gene sequence (KP284099) of mosquitofish to design a pair of primers spanning the β-actin exon and intron regions. All primer sequences are presented in Table 2.

thumbnail
Table 1. List of cone opsin genes from cyprinodontiform fishes, medaka (Oryzias latipes), and zebrafish (Danio rerio).

https://doi.org/10.1371/journal.pone.0240313.t001

thumbnail
Table 2. Sequences of the primers used for amplifying β-actin and cone opsin genes and for quantifying expression levels of cone opsin genes in the western mosquitofish.

https://doi.org/10.1371/journal.pone.0240313.t002

PCR amplifications of opsin genes were performed in a final reaction volume of 25 μL, containing 2 ng cDNA, 1 μl each of the forward and reverse primers (10 μM), 12.5 μl of Fast-RunTM Advanced Taq Master Mix (ProTech, Taipei, Taiwan), and distilled water. The thermal cycling protocol was as follows: one cycle at 94°C for 4 min; 35 cycles of denaturation at 94°C for 30 sec, 55–60°C for 30 sec, and 72°C for 1 min; and a final single extension step at 72°C for 5 min. Sequencing was performed by Mission Biotech Inc., Taipei, Taiwan. The contig sequences were constructed using the program CodonCode Aligner 6.0.2, and the results were identified by BLAST analysis against the database of the National Center for Biotechnology Information (NCBI).

Phylogenetic analyses

The cone opsin gene sequences of other poeciliid fishes, medaka (Oryzias latipes), and zebrafish (Danio rerio) were downloaded from NCBI, and two non-visual opsin genes of the zebrafish—val-opsin and tmt-opsin—were selected as outgroups [38]. All opsin gene sequences included in our phylogenetic analysis are shown in Table 1. Gene sequences were aligned using MACSE v2, which is designed to align protein-coding nucleotide sequences based on their corresponding amino acid translations [47]. Phylogenetic analyses were performed using a partitioned Maximum Likelihood (ML) approach. RAxML 8 [48] was used for ML analyses (MLA). Partitions were set with respect to codon position, and the GTR+G+I model (with four discrete rate categories) was adopted for each partition [49]. The ML tree was obtained by performing 100 different runs using the default algorithm. The best ML tree was chosen from likelihood scores among suboptimal trees from each run. Nodal support for MLA was determined from bootstrap analysis using RAxML [50], employing 10,000 non-parametric bootstrap replications with the ML criterion.

Apart from targeting the coding region of opsins, we also conducted phylogenetic analysis on the 5’ untranslated region (UTR) of lws from many poeciliid fishes given that Watson et al. (2011) [26] and Sandkam et al. (2017) [29] have demonstrated that gene conversion between poeciliid lws paralogs is more frequent in coding regions than in non-coding regions. Upon adding our 5’ UTR data from western mosquitofish to the poeciliid data of Sandkam et al. (2017) [29], the sequences were aligned using MUSCEL in Seqotron [51]. MLA was performed with the GTR+G+I model (with four discrete rate categories) [49] in RAxML 8 [48]. The ML tree was obtained by performing 100 different runs using the default algorithm. The best ML tree was chosen from likelihood scores among suboptimal trees from each run. Nodal support for MLA was determined from bootstrap analysis with RAxML [50], employing 10,000 non-parametric bootstrap replications with the ML criterion.

Prediction of the λmax values of the cone opsins

To predict the λmax values of western mosquitofish cone opsins, we relied on inferences based on its closest evolutionary relative, the guppy, for which λmax values were determined previously from in vitro expression data [52]. Moreover, given that some amino acid residues of a visual opsin have been demonstrated to have more pronounced effects on the λmax value than other residues, for this study, not only did we target the five residues (the 180th, 197th, 277th, 285th, and 308th amino acid residues corresponding to the human M/LWS pigment) based on the “5-sites rule” [1, 53], but also an additional five residues (the 110th, 112th, 138th, 223th, and 227th amino acid residues corresponding to the human M/LWS pigment), which also influence the λmax value of Sws2a in barfin flounder (Verasper moseri) and Rh-2 opsins in Oryzia [20, 54]. We assumed that the λmax values of two opsins were identical when a western mosquitofish cone opsin presented the same ‘10 key-sites’ composition as the orthologous opsin in guppy.

Real-time quantitative PCR (qPCR)

The specific primers designed for qPCR were based on the annotated sequences from our whole-genome BLAST and our draft transcriptome. The amplification efficiency and melting curve of each qPCR primer pair was tested by 5-fold serial dilutions of the templates, with three replicates for each gene and sample. Each qPCR primer pair was adopted only when its amplification efficiency fell between 90% and 110% and the melting curve analysis revealed each pair of specific primers only generated a single product. Expression of opsin genes was determined by qPCR in a Roche LightCycler480 system (Roche). Each reaction contained 10 μl of Roche LightCycler480 SYBR Green I Master (Roche Applied Science, Indianapolis, IN), 50 ng of cDNA, and 1 μl of each primer (10 μM) (Table 2) in a final volume of 20 μl. The qPCR reactions were performed in a LightCycler 480 Multiwell Plate system (Roche 04729692001) with Optical Adhesive Film (Applied Biosystems Ref 4360954). The following thermal cycles were performed: one cycle of 50°C for 2 min and 95°C for 10 min; followed by 45 cycles of 95°C for 10 sec, 60°C for 10 sec, and 72°C for 10 sec; and then one cycle of 95°C for 5 sec and 65°C for 1 min. qPCR products were qualified according to a melting-curve analysis. Additionally, representative samples were electrophoresed to verify that only a single product (band) was present. RNA-free water was used as a template in the control reactions to determine non-specific primer amplification background levels. Three replicates were performed for each cone opsin gene for each specimen.

Proportional expression of each cone opsin was calculated using the following equation: where Ti/Tall is the proportional expression of a given opsin gene i, and Ei is the amplification efficiency for each pair of opsin primers.

Expression of each cone opsin was first evaluated using Levene’s test and then compared using one-way ANOVA with a post-hoc Tukey HSD when Levene’s test was not significant or using Welch’s ANOVA with a Games-Howell post-hoc test when Levene’s test was significant. Statistical tests were performed in R version 3.6.0 (R Foundation for Statistical Computing, Vienna, Austria).

Results

Identification of cone opsins

The naming of poeciliid cone opsins has varied through time, but was recently reviewed by Sandkam et al. (2018) [28]. Based on that review, we have labelled the cone opsins we identified in western mosquitofish according to the orthologous opsins in guppy. We identified five cone opsin genes from the published draft genome of western mosquitofish [46]: sws1, sws2a, sws2b, rh2-1, and lws-r. We annotated six cone opsins from our draft transcriptome, including sws1, sws2a, sws2b, rh2-1, lws-1, and lws-3. These newly identified cone opsin genes (apart from lws-r, see below) were amplified from cDNA using specific primers (Table 2), with accession numbers MN817658 to MN817663 (Table 1).

The size of β-actin fragments amplified using DNA or cDNA as template differed (Fig 1B), with DNA templates producing larger fragments. However, amplified lws-r fragments were all identical in size no matter which template was employed (Fig 1A). Compared to the cDNA sequence of lws-r in guppy (KX768568), the western mosquitofish lws-r transcript contains an intron and also has a 46-basepair deletion in the coding region (equivalent to positions 1166 to 1211 of the guppy complete lws-r sequence). This intronic retention in the western mosquitofish lws-r transcript would result in an early stop codon during translation (S2 Fig), so it was excluded from further analyses apart from our phylogenetic assessment of the 5’ UTR region of lws.

thumbnail
Fig 1. PCR-amplified fragment sizes of lws-r and β-actin.

(A) lws-r opsin, and (B) β-actin. The initial character of each lane label represents DNA (D) or cDNA (R), and the number represents a specific specimen: 1 to 3 are female specimens collected at Tunghai University; 4 to 6 are male specimens collected from Tunghai University; 7 and 8 are female specimens collected at Linyuan Ocean Wetland Park; and 9 and 10 are male specimens collected at Linyuan Ocean Wetland Park.

https://doi.org/10.1371/journal.pone.0240313.g001

Phylogenies of western mosquitofish cone opsins

The cone opsin gene dataset comprised a total of 85 taxa and 1344 aligned nucleotide positions, with 983 variable sites and 848 parsimony-informative sites (S1 Appendix). In our cone opsin ML tree, the sws1, sws2, rh2, and lws families are all respectively clustered with high statistical support (bootstrap values ≥ 70), with the lws family lying basal to these other three cone opsin families (Fig 2). This cone opsin ML tree also demonstrates that all six cone opsins from the western mosquitofish cluster with their corresponding cone opsin genes from other poeciliid fishes with high statistical support. The two western mosquitofish lws genes (lws-1 and lws-3) are clustered together and are paralogous to the corresponding genes of other poeciliid fishes (Fig 2E), and the other four cone opsin genes of the western mosquitofish (sws1, sws2a, sws2b, and rh2-1) are all orthologous to their corresponding genes in other poeciliids (Fig 2B to 2D).

thumbnail
Fig 2. Phylogenetic tree of the coding regions from cone opsin genes based on partitioned maximum likelihood analysis (three partitions).

Branch colors represent different cone opsin families. Phylogenies for each opsin family are shown in panels B to E. Solid circles on branch nodes indicate statistically robust nodes with bootstrap values ≥ 70.

https://doi.org/10.1371/journal.pone.0240313.g002

The 5’ UTR dataset of lws genes comprised a total of 64 taxa and 140 aligned nucleotide positions (S2 Appendix), with 111 variable sites and 84 parsimony-informative sites. From our lws 5’ UTR ML tree (Fig 3), it is clear that lws genes from poeciliid fishes can be divided into four strongly supported groups: lws-1 group, lws-2 group, lws-3 group, and lws-r group. The three lws genes possessed by western mosquitofish (lws-1, lws-3, and lws-r) are all orthologous to the corresponding lws genes of other poeciliid fishes. Notably, our genomic (BLAST) and transcriptomic analyses reveal lws-2 to be absent from western mosquitofish.

thumbnail
Fig 3. Phylogenetic tree of the 5’ untranslated region (UTR) of poeciliid lws opsin genes based on maximum likelihood analysis.

Oreochromis niloticus was used as outgroup. Solid circles on branch nodes indicate statistically robust nodes with bootstrap values ≥ 70.

https://doi.org/10.1371/journal.pone.0240313.g003

Predicted λmax values of western mosquitofish cone opsins

The λmax values of guppy cone opsins are 353 nm (Sws1), 438 nm (Sws2a), 408 nm (Sws2b), 516 nm (Rh2-1), 571 nm (Lws-1), and 519 nm (Lws-3) [52]. Apart from Sws2a, all western mosquitofish cone opsins have the same ‘10-sites’ composition as their orthologs in guppy, so we infer the λmax values of western mosquitofish cone opsins to be as follows: Sws1 = 353 nm, Sws2b = 408 nm, Rh2-1 = 516 nm, Lws-1 = 571 nm, and Lws-3 = 519 nm. Compared to western mosquitofish Sws2a, guppy Sws2a has an Ala-to-Thr mutation at residue 110 (Ala110Thr) and a Thr-to-Ala mutation at residue 285 (Thr285Ala). Comparing the two orthologous Sws2a opsins from guppy and medaka, we assume that the effect of Ala110Thr on the λmax value is -1 nm. Moreover, Cowing et al. (2002) [55] inserted the Thr285Ala mutation into the Sws2 opsin of Cottus gobio and observed an approximatley -10 nm shift in the λmax value. Accordingly, we infer a λmax value for western mosquitofish Sws2a of 449 (= 438+1+10) nm (Table 3).

thumbnail
Table 3. Ten-site amino acid haplotypes for cone opsin genes in guppy and western mosquitofish.

https://doi.org/10.1371/journal.pone.0240313.t003

Cone opsin gene expression

The rh2-1 cone opsin gene exhibited the greatest expression in all developmental stages we tested and for both sexes, whereas proportional expression levels of both the sws2a and lws-3 opsin genes were always lower than 1% (Fig 4). Apart from lws-3 (ANOVA F2, 16 = 1.3571, p > 0.05), proportional expression of all of the western mosquitofish cone opsin genes was significantly different among larval, adult male, and adult female specimens (Fig 4). One-way ANOVA and post-hoc Tukey HSD revealed that larvae expressed more Sws1 and Sws2a opsins than adults of both sexes (Sws1 F2, 16 = 4.9853, p < 0.05; Sws2a F2, 16 = 13.529, p < 0.05). A Welch’s ANOVA test also supported that expression levels of the sws2b (F2, 9.108 = 6.0534, p < 0.05), rh2-1 (F2, 10.672 = 13.231, p < 0.05), and lws-1 (F2, 9.81 = 16.72, p < 0.05) genes were all significantly different among larval, adult male, and adult female specimens. A post-hoc Games-Howell test revealed that larvae expressed less Sws2b opsin than adult males, less Lws-1 than adults of both sexes, but more Rh2-1 opsin than adults of both sexes.

thumbnail
Fig 4. Box-plot showing the median (—), 25th and 75th percentiles (box), 95% range (|), and outliers (•) of cone opsin proportional expression values for adult and larval specimens.

An asterisk (*) indicates p < 0.05.

https://doi.org/10.1371/journal.pone.0240313.g004

Discussion

We identified a total of seven cone opsin genes, including one sws1, two sws2, one rh2, and potentially three lws, in this study. In teleosts, a visual opsin genic tandem array is commonly observed for sws2, rh2, and lws paralogs [13], and the sws2-lws synteny is a synapomorphy from the common ancestor of mammals, birds, reptiles, and fish [57]. However, the western mosquitofish draft genome [46] only revealed that two sws2 genes are tandemly arrayed, indicating that the draft genome may not be complete. Lin et al. (2017) [13] considered one of the lws genes, lws-r, to be apomorphic in the Cyprinodontoidei since it lacks intron II-V but has a highly conserved exon/intron I. The lws-r gene is also thought to have arisen by retrotransposition and is located in an intron of gephyrin (gphn) [26, 58]. Here, we confirm by whole-genome BLAST analysis that the western mosquitofish has the lws-r gene and, as for guppy and green swordtail, its lws-r also contains only one intron and is located in a gphn gene intron.

We identified a further two western mosquitofish lws opsin transcripts (lws-1 and lws-3) by RNA-seq. In our phylogenetic tree of lws 5’ UTRs, these two lws genes cluster separately and with high statistical support with other poeciliid lws-1 and lws-3 genes, respectively (Fig 3). However, the phylogenetic analysis based on the coding regions revealed a very different topology in that these two lws genes grouped together with high statistical support rather than separately clustering with their own orthologous lws genes (Fig 2E). Thus, the lws-1 and lws-3 genes appear to group by species as if independent gene duplication events have occurred in poeciliid fishes. Sandkam et al. (2017) [29] stated that the genomic environment in poeciliid fishes tends to homogenize their lws-1 and lws-3 genes through gene conversion because of their genomic proximity, gene structure, and orientation, but that this gene conversion may be curbed in some lineages where bright red/orange male coloration is critical to sexual selection. Moreover, the distinct topologies of the coding region-based tree and the 5’ UTR-based tree is consistent with gene conversion being more frequent in coding regions than in non-coding ones [62]. Although the draft western mosquitofish genome does not reveal if lws-1 and lws-3 are in close proximity, have a similar gene structure, or are in inverted orientations, our coding region-based phylogenetic tree may indicate that gene conversion has homogenized the western mosquitofish lws-1 and lws-3 genes.

Compared to the guppy (nine cone opsins), western mosquitofish appear to lack two cone opsin genes, i.e., rh2-2 and lws-2. It is possible that these two opsins have been lost from the western mosquitofish genome, which is not unusual among the Cyprinodontoidei [13]. Alternatively, these two opsins do exist in the western mosquitofish genome but they are not (or barely) expressed in the larval/adult stages we sampled.

That the lws-r gene is located in the gphn intronic region raises two questions. First, does lws-r interfere with the expression of gphn and, second, is a hybrid transcript generated [27]? Based on our draft transcriptome, we can confirm that the western mosquitofish expresses gphn as well as the green swordtail, and can also rule out gene fusion between these two genes. However, although western mosquitofish lws-r is present in the draft genome and we successfully amplified it from cDNA, it is absent from our transcriptome assembly. Interestingly, the lws-r sequences from DNA and cDNA template (Fig 1A) were of the same length and both contained the intron region (S2 Fig). This outcome is not likely due to DNA contamination because β-actin fragments amplified from cDNA samples exhibited a single size and were shorter than those of DNA samples (Fig 1B).

In the process of library preparation for RNA-seq, we used poly-T oligo-attached magnetic beads to purify mRNA. Thus, the absence of lws-r in our RNA-seq dataset may be attributable to lws-r transcript not undergoing full post-transcriptional modification so that the poly(A) tail had not been added. Alternatively, transcription of the gphn gene may enable lws-r in guppy and green swordtail to manifest [27, 58]. This “hijack” hypothesis may explain why the premature lws-r, containing the intronic sequence and lacking the poly(A) tail, could be amplified from cDNA template. In our study, cDNA samples were reverse-transcribed using both Hexamer and Anchored Oligo-dT, so the spliced intronic fragments from post-transcriptional gphn modification could be converted into cDNA. It should be noted that we did not detect any “mature” mRNA of lws-r from either larval or adult samples, so it is doubtful that Lws-r exerts a light-sensing function in western mosquitofish. Yokoyama et al. (2014) [59] reported that the sws1 opsin gene of the pearleye (Benthalbella dentata) became a pseudogene about 130 million years ago, but it still produces transcripts and has a low nucleotide substitution rate, perhaps indicating that it retains some unknown functions. It would be interesting to determine if lws-r in the western mosquitofish is a functional visual opsin, has other unknown functions, or its transcript is just a side product of gphn expression.

Apart from Sws2a, we found that the other five identified cone opsins in western mosquitofish possess the same “10-sites” composition as their guppy orthologues. Accordingly, we inferred that these five cone opsins have the same λmax values as defined for guppy. A shift in the λmax value of a visual opsin can alter the evolutionary trajectory of another one. For instance, in many species of flounder, a shift to green of Sws2a has been suggested as having compensated for the functional loss of one of their Rh2 paralogs [20, 60]. This green-shift phenomenon might also occur in western mosquitofish. Compared to guppy Sws2a, the Ala110Thr and Thr285Ala mutations in western mosquitofish Sws2a may induce an 11-nm shift in long-wavelength [55, 56]. By incorporating more closely-related poeciliids in further genomic studies, it should be possible to determine if the greenward shift of Sws2a is a response to loss of the rh2-2 gene from the western mosquitofish genome, as has been reported for Verasper fishes [60].

Our qPCR results reveal that apart from lws-3 with less than 0.05% proportional expression for all developmental stages and sexes, the remaining five cone opsins all display significant differences in proportional expression between larval and adult stages. The significant ontogenetic decrease in sws1 expression and increase in lws-1 suggest that the visual spectrum of adult western mosquitofish is more sensitive to longer wavelengths than that of larvae. UV sensitivity is critical for detecting zooplankton [6166]. Many fishes display an ontogenetic long-wavelength shift in spectral sensitivity when they no longer feed on zooplankton [6770]. Since Gambusia fishes exhibit an ontogenetic dietary shift from micro to large invertebrates [33, 71], the decline in ultraviolet sensitivity of western mosquitofish is consistent with its ontogenetic dietary shift.

The lws-1 allele of western mosquitofish is 180S, which is a more red-sensitive allele (180S, λmax = 571 nm, compared to 180A λmax = 562 nm; Table 3). Sandkam et al. (2015) [10] demonstrated that female guppy carrying the 180S allele prefer males with more abundant red/orange coloration than those carrying the 180A allele. Therefore, allele type and expression levels of lws-1 could not only explain why adult male western mosquitofish are strongly attracted to the yellow spot around the urogenital opening of mature females [34], but also why female western mosquitofish are more inclined to breed with male guppies than conspecific males [72].

In conclusion, we have verified that the western mosquitofish has seven cone opsin genes, including one sws1, two sws2, one rh2, and three lws, but whether the lws-r gene is functional or not remains unclear. The predicted λmax value for Sws1 is 353 nm, it is 449 nm for Sws2a, 408 nm for Sws2b, 526 nm for Rh2-1, 571 nm for Lws-1, and 519 nm for Lws-3. The mutations at residues 110 and 285 of Sws2a have likely induced a green-shift in the λmax value of that opsin, which may be associated with loss of the rh2-2 gene. We have also revealed an ontogenetic red shift in the visual spectrum of western mosquitofish. The ontogenetic decline in UV-sensitivity could be correlated with a dietary shift. Furthermore, lws-1 allele type and expression level appears to reflect the asymmetric negative reproductive interference reported between sympatric guppy and western mosquitofish. The comprehensive understanding of cone opsins in western mosquitofish from this study will facilitate further explorations of the mechanisms, such as developmental and/or adult plasticity of opsin expression, that enable western mosquitofish to flourish better in turbid waters compared to other fishes [73].

Supporting information

S1 Fig. The lighting spectrum of the full-spectrum LED light deployed above the aquarium in this study.

https://doi.org/10.1371/journal.pone.0240313.s001

(JPG)

S2 Fig. Nucleotide sequence alignment of the western mosquitofish lws-r fragment from the published draft genome [51], the sequences we generated from DNA and cDNA templates in this study, and the published lws-r complete gene and cDNA sequences of guppy (KX768568).

https://doi.org/10.1371/journal.pone.0240313.s002

(PDF)

Acknowledgments

The first author is grateful to Dr. Feng-Yu Wang at National Applied Research Laboratories, Taiwan, and Dr. Yung-Che Tseng at the Institute of Cellular and Organismic Biology, Academia Sinica for offering technical assistance. The authors are also grateful to Dr. John O’Brien for editing assistance.

References

  1. 1. Yokoyama S. Molecular evolution of vertebrate visual pigments. Prog Retin Eye Res. 2000;19(4):385–491. pmid:10785616
  2. 2. Locket NA. Vertebrate photoreceptors. Archer SN, Djamgoz MBA, Loew ER, Partridge JC, Vallerga S, editors. Dordrecht: Springer; 1999.
  3. 3. Chang BSW, Crandall KA, Carulli JP, Hartl DL. Opsin phylogeny and evolution: a model for blue shifts in wavelength regulation. Mol Phylogenet Evol. 1995;4(1):31–43. pmid:7620634
  4. 4. Musilova Z, Cortesi F, Matschiner M, Davies WIL, Patel JS, Stieb SM, et al. Vision using multiple distinct rod opsins in deep-sea fishes. Science. 2019;364(6440):588–92. pmid:31073066
  5. 5. Lythgoe JN. Ecology of Vision. New York: Oxford University Press; 1979.
  6. 6. Hauser FE, Chang BS. Insights into visual pigment adaptation and diversity from model ecological and evolutionary systems. Curr Opin Genet Dev. 2017;47:110–20. pmid:29102895
  7. 7. Seehausen O, Terai Y, Magalhaes IS, Carleton KL, Mrosso HDJ, Miyagi R, et al. Speciation through sensory drive in cichlid fish. Nature. 2008;455:620–6. pmid:18833272
  8. 8. Fuller RC, Fleishman LJ, Leal M, Travis J, Loew E. Intraspecific variation in retinal cone distribution in the bluefin killifish, Lucania goodei. J Comp Physiol A. 2003;189(8):609–16. pmid:12879350
  9. 9. Fuller RC, Carleton KL, Fadool JM, Spady TC, Travis J. Population variation in opsin expression in the bluefin killifish, Lucania goodei: a real-time PCR study. J Comp Physiol A. 2004;190(2):147–54. pmid:14685760
  10. 10. Sandkam B, Young CM, Breden F. Beauty in the eyes of the beholders: colour vision is tuned to mate preference in the Trinidadian guppy (Poecilia reticulata). Mol Ecol. 2015;24(3):596–609. pmid:25556876
  11. 11. Kirk JTO. Use of a quanta meter to measure attenuation and underwater reflectance of photosynthetically active radiation in some inland and coastal south-eastern Australian waters. Australian J Mar Freshwater Res. 1977;28(1):9–21.
  12. 12. Kirk JTO. Attenuation of light in natural waters. Australian J Mar Freshwater Res. 1977;28(4):497–508.
  13. 13. Lin J-J, Wang F-Y, Li W-H, Wang T-Y. The rises and falls of opsin genes in 59 ray-finned fish genomes and their implications for environmental adaptation. Sci Rep. 2017;7:15568. pmid:29138475
  14. 14. Carleton K. Cichlid fish visual systems: mechanisms of spectral tuning. Integr Zool. 2009;4(1):75–86. pmid:21392278
  15. 15. Yourick MR, Sandkam BA, Gammerdinger WJ, Escobar‐Camacho D, Nandamuri SP, Clark FE, et al. Diurnal variation in opsin expression and common housekeeping genes necessitates comprehensive normalization methods for quantitative real‐time PCR analyses. Mol Ecol Resour. 2019;19(6):1447–60. pmid:31325910
  16. 16. Sandkam BA, Deere-Machemer KA, Johnson AM, Grether GF, Rodd FH, Fuller RC. Exploring visual plasticity: dietary carotenoids can change color vision in guppies (Poecilia reticulata). J Comp Physiol A. 2016;202:527–34. pmid:27283858
  17. 17. Tettamanti V, de Busserolles F, Lecchini D, Marshall NJ, Cortesi F. Visual system development of the spotted unicornfish, Naso brevirostris (Acanthuridae). J Exp Biol. 2019;222:jeb209916. pmid:31776185
  18. 18. Wang F-Y, Fu W-C, Wang I-L, Yan HY, Wang T-Y. The giant mottled eel, Anguilla marmorata, uses blue-shifted rod photoreceptors during upstream migration. PLoS One. 2014;9(8):e103953. pmid:25101636
  19. 19. Chang C-H, Chiao C-C, Yan HY. Ontogenetic changes in color vision in the milkfish (Chanos chanos Forsskål, 1775). Zool Sci. 2009;26(5):349–55. pmid:19715505
  20. 20. Kasagi S, Mizusawa K, Murakami N, Andoh T, Furufuji S, Kawamura S, et al. Molecular and functional characterization of opsins in barfin flounder (Verasper moseri). Gene. 2015;556(2):182–91. pmid:25433330
  21. 21. Kranz AM, Forgan LG, Cole GL, Endler JA. Light environment change induces differential expression of guppy opsins in a multi‐generational evolution experiment. Evolution. 2018. pmid:29920667
  22. 22. Ehlman SM, Sandkam BA, Breden F, Sih A. Developmental plasticity in vision and behavior may help guppies overcome increased turbidity. J Comp Physiol A. 2015;201(12):1125–35. pmid:26427995
  23. 23. Nandamuri SP, Yourick MR, Carleton KL. Adult plasticity in African cichlids: Rapid changes in opsin expression in response to environmental light differences. Mol Ecol. 2017;26(21):6036–52. pmid:28926160
  24. 24. Fuller RC, Claricoates KM. Rapid light-induced shifts in opsin expression: finding new opsins, discerning mechanisms of change, and implications for visual sensitivity. Mol Ecol. 2011;20(16):3321–35. pmid:21749514
  25. 25. Chang C-H, Yan HY. Plasticity of opsin gene expression in the adult red shiner (Cyprinella lutrensis) in response to turbid habitats. PLoS One. 2019;14(4):e0215376. pmid:30978235
  26. 26. Watson CT, Gray SM, Hoffmann M, Lubieniecki KP, Joy JB, Sandkam BA, et al. Gene duplication and divergence of long wavelength-sensitive opsin genes in the guppy, Poecilia reticulata. J Mol Evol. 2011;72(2):240–22. pmid:21170644
  27. 27. Watson CT, Lubieniecki KP, Loew E, Davidson WS, Breden F. Genomic organization of duplicated short wave-sensitive and long wave-sensitive opsin genes in the green swordtail, Xiphophorus helleri. BMC Evol Biol. 2010;10:87. pmid:20353595
  28. 28. Sandkam B, Dalton B, Breden F, Carleton K. Reviewing guppy color vision: integrating the molecular and physiological variation in visual tuning of a classic system for sensory drive. Curr Zool. 2018;64(4):535–45. pmid:30108634
  29. 29. Sandkam BA, Joy JB, Watson CT, Breden F. Genomic environment impacts color vision evolution in a family with visually based sexual selection. Genome Biol Evol. 2017;9(11):3100–7. pmid:29121209
  30. 30. Francis RA. A Handbook of Global Freshwater Invasive Species. New York: Earthscan; 2012.
  31. 31. Pyke GH. Plague minnow or mosquito fish? A review of the biology and impacts of introduced Gambusia species. Annu Rev Ecol Evol Syst. 2008;39:171–91.
  32. 32. Friesen CN, Ramsey ME, Cummings ME. Differential sensitivity to estrogen-induced opsin expression in two poeciliid freshwater fish species. Gen Comp Endocrinol. 2017;246:200–10. pmid:28013033
  33. 33. Lee F, Simon KS, Perry GLW. Prey selectivity and ontogenetic diet shift of the globally invasive western mosquitofish (Gambusia affinis) in agriculturally impacted streams. Ecol Freshw Fish. 2018;27(3):822–33.
  34. 34. Kodama I, Yamanaka A, Endo K, Koya Y. Role of the yellow spot around the urogenital opening of female mosquitofish (Gambusia affinis) as a cue for copulation. Zool Sci. 2008;25(12):1199–204. pmid:19267646
  35. 35. Polverino G, Liao JC, Porfiri M. Mosquitofish (Gambusia affinis) preference and behavioral response to animated images of conspecifics altered in their color, aspect ratio, and swimming depth. PLoS ONE. 2013;8(1):e54315. pmid:23342131
  36. 36. Casner AM, Fackelman HC, Degtyareva O, Kight SL. Do female western mosquitofish, Gambusia affinis, prefer ornaments that males lack? Ethology. 2016;122(7):561–70.
  37. 37. Madliger CL. Toward improved conservation management: a consideration of sensory ecology. Biodivers Conserv. 2012;21:3277–86.
  38. 38. Terakita A. The opsins. Genome Biol. 2005;6(3):213–22. pmid:15774036
  39. 39. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30(15):2114–20. pmid:24695404
  40. 40. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, et al. Trinity: reconstructing a full-length transcriptome without a genome from RNA-Seq data. Nat Biotechnol. 2011;29(7):644–52. pmid:21572440
  41. 41. Fu L, Niu B, Zhu Z, Wu S, Li W. CD-HIT: accelerated for clustering the next-generation sequencing data. Bioinformatics. 2012;28(23):3150–2. pmid:23060610
  42. 42. Li W, Jaroszewski L, Godzik A. Clustering of highly homologous sequences to reduce the size of large protein databases. Bioinformatics. 2001;17(3):282–3. pmid:11294794
  43. 43. Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nat Methods. 2012;9:357–9. pmid:22388286
  44. 44. Li B, Dewey CN. RSEM: accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinformatics. 2011;12:323. pmid:21816040
  45. 45. Camacho C, Coulouris G, Avagyan V, Ma N, Papadopoulos J, Bealer K, et al. BLAST+: architecture and applications. BMC Bioinformatics. 2009;10:421. pmid:20003500
  46. 46. Hoffberg SL, Troendle NJ, Glenn TC, Mahmud O, Louha S, Chalopin D, et al. A high-quality reference genome for the invasive mosquitofish Gambusia affinis using a Chicago library. G3-Genes Genomes Genet. 2018;8(6):1855–61. pmid:29703783
  47. 47. Ranwez V, Douzery EJP, Cambon C, Chantret N, Delsuc F. MACSE v2: toolkit for the alignment of coding sequences accounting for frameshifts and stop codons. Mol Biol Evol. 2018;35(10):2582–4. pmid:30165589
  48. 48. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30(9):1312–3. pmid:24451623
  49. 49. Abadi S, Azouri D, Pupko T, Mayrose I. Model selection may not be a mandatory step for phylogeny reconstruction. Nat Commun. 2019;10:934. pmid:30804347
  50. 50. Felsenstein J. Confidence limits on phylogenies: an approach using the bootstrap. Evolution. 1985;39(4):783–91. pmid:28561359
  51. 51. Fourment M, Holmes EC. Seqotron: a user-friendly sequence editor for Mac OS X. BMC Res Notes. 2016;9:106. pmid:26887850
  52. 52. Kawamura S, Kasagi S, Kasai D, Tezuka A, Shoji A, Takahashi A, et al. Spectral sensitivity of guppy visual pigments reconstituted in vitro to resolve association of opsins with cone cell types. Vision Res. 2016;127:67–73. pmid:27476645
  53. 53. Yokoyama S, Radlwimmer FB. The "five-sites" rule and the evolution of red and green color vision in mammals. Mol Biol Evol. 1998;15(5):560–7. pmid:9580985
  54. 54. Matsumoto Y, Oda S, Mitani H, Kawamura S. Orthologous divergence and paralogous anticonvergence in molecular evolution of triplicated green opsin genes in medaka fish, genus Oryzias. Genome Biol Evol. 2020;12(6):911–23. pmid:32467976
  55. 55. Cowing JA, Poopalasundaram S, Wilkie SE, Bowmaker JK, Hunt DM. Spectral tuning and evolution of short wave-sensitive cone pigments in cottoid fish from Lake Baikal. Biochemistry. 2002;41(19):6019–25. pmid:11993996
  56. 56. Matsumoto Y, Fukamachi S, Mitani H, Kawamura S. Functional characterization of visual opsin repertoire in Medaka (Oryzias latipes). Gene. 2006;371(2):268–78. pmid:16460888
  57. 57. Wakefield MJ, Anderson M, Chang E, Wei K-J, Kaul R, Graves JAM, et al. Cone visual pigments of monotremes: Filling the phylogenetic gap. Visual Neurosci. 2008;25(3):257–64. pmid:18598396
  58. 58. Laver CRJ, Taylor JS. RT-qPCR reveals opsin gene upregulation associated with age and sex in guppies (Poecilia reticulata)—a species with color-based sexual selection and 11 visual-opsin genes. BMC Evol Biol. 2011;11:81. pmid:21447186
  59. 59. Yokoyama S, Starmer WT, Liu Y, Tada T, Britt L. Extraordinarily low evolutionary rates of short wavelength-sensitive opsin pseudogenes. Gene. 2014;534(1):93–9. pmid:24125953
  60. 60. Kasagi S, Mizusawa K, Takahashi A. Green‐shifting of SWS2A opsin sensitivity and loss of function of RH2‐A opsin in flounders, genus Verasper. Ecol Evol. 2018;8(2):1399–410. pmid:29375806
  61. 61. Leech DM, Boeing WJ, Cooke SL, Williamson CE, Torres L. UV‐enhanced fish predation and the differential migration of zooplankton to UV radiation and fish. Limnol Oceanogr. 2009;54(4):1152–61.
  62. 62. Loew ER, McFarland WN, Mills EL, Hunter D. A chromatic action spectrum for planktonic predation by juvenile yellow perch, Perca flavescens. Can J Zool. 1993;71(2):384–6.
  63. 63. Flamarique IN. Diminished foraging performance of a mutant zebrafish with reduced population of ultraviolet cones. Proc R Soc B. 2016;283(1826):20160058. pmid:26936243
  64. 64. Flamarique IN. Opsin switch reveals function of the ultraviolet cone in fish foraging. Proc R Soc B. 2013;280(1752):20122490. pmid:23222448
  65. 65. Britt LL, Loew ER, McFarland WN. Visual pigments in the early life stages of Pacific northwest marine fishes. J Exp Biol. 2001;204:2581–7. pmid:11511674
  66. 66. McFarland WN, Loew ER. Ultraviolet visual pigments in marine fishes of the family pomacentridae. Vision Res. 1994;34(11):1393–6. pmid:8023448
  67. 67. Shand J, Hart NS, Thomas N, Partridge JC. Developmental changes in the cone visual pigments of black bream Acanthopagrus butcheri. J Exp Biol. 2002;205:3661–7. pmid:12409492
  68. 68. Cheng CL, Flamarique IN. Chromatic organization of cone photoreceptors in the retina of rainbow trout: single cones irreversibly switch from UV (SWS1) to blue (SWS2) light sensitive opsin during natural development. J Exp Biol. 2007;210:4123–35. pmid:18025012
  69. 69. Shand J, Partridge JC, Archer SN, Potts GW, Lythgoe JN. Spectral absorbance changes in the violet/blue sensitive cones of the juvenile pollack, Pollachius pollachius. J Comp Physiol A. 1988;163:699–703.
  70. 70. Matsumoto T, Ishibashi Y. Sequence analysis and expression patterns of opsin genes in the longtooth grouper Epinephelus bruneus. Fish Sci. 2016;82:17–27.
  71. 71. García‐Berthou E. Food of introduced mosquitofish: ontogenetic diet shift and prey selection. J Fish Biol. 1999;55(1):135–47.
  72. 72. Tsurui-Sato K, Fujimoto S, Deki O, Suzuki T, Tatsuta H, Tsuji K. Reproductive interference in live-bearing fish: the male guppy is a potential biological agent for eradicating invasive mosquitofish. Sci Rep. 2019;9:5439. pmid:30931976
  73. 73. Abrahams MV, Bassett DK, Montgomery JC. Sensory biology as a risk factor for invasion success and native fish decline. Trans Am Fish Soc. 2017;146(6):1238–44.