Acanthogobius ommaturus is one of the suitable species to study the genetic mechanism of adaptive evolution, but there are few reports on its genetics. In the present study, the genomic survey sequencing method was used to analyze the genome characters of A. ommaturus. A total of 50.50 G high-quality sequence data were obtained in the present study. From the 19-mer distribution frequency, the estimated genome size was 928.01 Mb. The calculated sequence repeat rate was about 38.31%, the heterozygosity was approximately 0.17%, and the GC% content was approximately 40.88%. Moreover, 475,724 simple sequence repeats (SSRs) were identified. Among them, dinucleotide repeats were the most (53.70% of the total SSRs), followed by tri- (35.36%), hexa- (4.59%), tetra- (4.57%) and penta- (1.77%) nucleotide repeats type. This is the first genome-wide feature of this species to be reported.

Acanthogobius ommaturus belongs to the order Perciformes, Gobioidei, Gobiidae and Acanthogobius. It is a large-body goby fish distributed in the subsea or brackish waters of China, Korea, Japan and Indonesia [1]. The suborder Gobioidei is the largest group of fish species, with more than 2000 species reported [2]. Most goby fish have a wide distribution range, strong adaptability and high diversity, which are suitable for studying the genetic mechanism of adaptive evolution [3]. Acanthogobius ommaturus is an economically important species that has been reported to be the dominant species across its distribution area [4–6]. In recent years, many studies focusing on its ecological habits, morphological traits, biochemical determinations and feeding strategy in aquaculture industry have been conducted. For example, Feng et al. studied the morphological character, individual fecundity and ovary histology [1]. Huang et al. explored the effects of cryopreservation on several immune and antioxidative enzyme activities [7]. The characterization of Lactate Dehydrogenase in A. ommaturus was also analyzed [8]. Insoo and Hong investigated its feeding ecology in a tidal flat in Korea [9]. However, only few studies have involved its genetic diversity and phylogeographic patterns. Until now NCBI has only published the genome information of seven species of Gobioidei, and it cannot provide us valid information of A. ommaturus because of distinct genetic relationship between them [10]. In order to understand the genetic background of A. ommaturus and conduct further genetic studies, it is necessary to understand genome characteristics of A. ommaturus.

At present, high-throughput next-generation sequencing (NGS) is the main method of genome investigation, which is an important and effective strategy to generate genetic and genomic information [11]. Genome survey sequencing bases on the low-depth sequencing data of the small fragment library [12], and can effectively evaluate the genome size, GC content and heterozygosity through K-mer analysis [13,14]. Compared with traditional methods, it is also an effective method to characterize microsatellite markers of target species. As a codominant molecular marker, microsatellite marker has been widely used in studies related to genetic diversity and population structure [15,16]. Although there were some population genetic studies based on different molecular markers, such as mitochondrial DNA and AFLP [17,18], there were no studies on population genetics of A. ommaturus by using microsatellite markers. Therefore, in the present study, based on high-throughput sequencing, K-mer analysis was used to investigate the genomic characteristics of A. ommaturus and also give profile of its microsatellite primers. The results of the present study will provide theoretical basis for the subsequent development of the whole genome de novo sequencing assembly strategy [11,19], and can be used to describe the genetic background of many other marine animals [11,20–22]. In addition, the draft genome that is obtained by this sequencing is also conducive to the detecting of SNP loci.

We have read the policies relating to animal experiments and confirmed the present study complied (ARRIVE guidelines; EU Directive 2010/63/EU for animal experiments). All procedures performed in the present study were approved by the Institutional Animal Care and Use Committee of Ocean University of China.

Sample preparation and DNA extraction

The samples were obtained by trawl net by fishermen from coastal water of Qingdao, China in October 2019, and we bought the lively and undamaged individuals in Qingdao wharf. Then, these fish were quickly transported to our laboratory for recovery. Approximately 7 days later, an active, healthy and undamaged A. ommaturus (overall length: 15.80 cm; body length: 12.9 cm; body weight: 30.50 g) was selected as the sample for sequencing. Selected A. ommaturus were killed to death with an overdose of anesthetic (tricaine methanesulfonate, MS-222, 100 mg/l). Subsequently, the fish were rapidly dissected, and part of the muscle tissues on the caudal peduncle were collected and preserved in 95% alcohol in ice box. The tissue samples were then quickly transformed to our lab where the samples were stored in −80°C until analysis. The traditional phenol–chloroform method was used to extract genomic DNA. The total DNA was treated with RNase to obtain DNA with high purity and without RNA pollution.

Library construction and sequencing

After DNA extraction, library construction and sequencing were conducted. The qualified genomic DNA was broken to the target fragment (350 base pairs) by physical fragmentation (ultrasonic vibration), and then the small-fragment sequencing library was constructed through the steps of terminal repair, addition of A, addition of connector, selection of target fragment and PCR. Agilent 2100 and Q-PCR were used to detect the size of library fragments and library quantification to determine whether the library meets the sequencing standard. The library was fixed to the sequencing chip by bridge PCR. After library quality inspection and chip fixation, two paired-end DNA libraries were constructed with insert size of 350 base pairs, and then sequenced using the Illumina HiSeq X ten platform following the manufacturer’s protocol. The data generated by sequencing was used for the next step of bioinformation analysis after quality control. Entire read sets were deposited in the short read archive (SRA) databank (http://www.ncbi.nlm.nih.gov/sra/) and are available under accession number PRJNA658176.

Analysis method

After GC distribution statistics and evaluation of quality values Q20 (ratio of bases with sequencing quality value above 20) and Q30 (proportion of bases with sequencing quality values above 30), clean reads were obtained after filtering the data of paired-end sequencing, which were used for a K-mer analysis. The K-mer analysis can be used to obtain the genome length, heterozygosity and repeat content.

The information of peak depth and number of predicted best K-mer could be obtained according to the results of the K-mer analysis, and subsequently this information was used to estimate the size of the genome. The formula was as follows in eqn (1):
Genome size =K-mer numberK-mer peak depth
(1)

where K-mer number was the total number of predicted best K-mer, and peak depth was the expected value of the K-mer depth. The heterozygosity ratio and repeat sequence ratio were determined according to the methods described in Song et al. [23]. The software Jellyfish (v2.0, http://www.genome.umd.edu/jellyfish.html) and GenomeScope (v1.0, http://qb.cshl.edu/genomescope/analysis.php?code=example6) were used to conduct K-mer analyses [24,25]. The clean reads were assembled into contigs in software SOAPdenovo (v2.01, https://github.com/aquaskyline/SOAPdenovo2/) by applying the de Bruijn graph structure [26]. The paired-end information was then used to join the unique contigs into scaffolds [11]. To assess the completeness of the assembly, a Benchmarking Universal Single-Copy Orthologs (BUSCO, v.4.0, https://busco.ezlab.org/) evaluation was performed using the Eukaryota vertebrata_odb10 database [27]. The Perl script MISA was performed to identify microsatellite motifs in the de novo draft genome, and search parameters were set as minimum of 6, 5, 5, 5 and 5 repeats for detecting di-, tri-, tetra-, penta- and hexanucleotide microsatellite motifs, respectively. Primer3 (v2.3.7) software with standard parameters was used to identify the microsatellite loci [28,29].

Genome sequencing and sequence quality estimation

A 350-bp insert library was constructed using the genomic DNA of the A. ommaturus sample, which was sequenced and filtered on the Illumina sequencing platform. A total of high-quality data of 50.50 Gb was obtained, with a total sequencing depth of approximate 54×. The Q20 and Q30 values were over 96% and 90%, respectively.

Genome size and repeat ratio, and heterozygosity assessment

Genome survey sequencing yielded a total of 50.50 Gb high-quality data, which were used for subsequent K-mer analysis. Using a 350-bp library data to construct a K-mer distribution map of K = 19 (Figure 1), and the genome size, repeat sequence ratio and heterozygosity were evaluated. The average K-mer depth that corresponded to the main peak, was 44, and the total number of K-mers obtained from the sequencing data was 44,434,973,013. After removing the anomalies of depth, a total of 41,080,854,391 K-mers were used for genome length estimation.

K-mer (K = 19) analysis for estimation of the genome size of A. ommaturus

Figure 1
K-mer (K = 19) analysis for estimation of the genome size of A. ommaturus

The X-axis was depth and the Y-axis was the proportion that represented the frequency at that depth.

Figure 1
K-mer (K = 19) analysis for estimation of the genome size of A. ommaturus

The X-axis was depth and the Y-axis was the proportion that represented the frequency at that depth.

Close modal

The calculated genome length was about 928.01 Mb. According to the reported genome data, the genome size of most fish is generally small and less than 1 Gb, except for a few fish species such as Cyprinus carpio, Megalobrama amblycephala and Sinocyclocheilus graphami [30]. Moreover, very few fish, like Salmo salar [31] and Oncorhynchus mykiss [32], are around 1.9-3.0 Gb because of a round of full gene duplication and the expansion of repeat sequence. The genome sizes of four Mudskippers (Gobioidei) were 0.983, 0.806, 0.780 and 0.739 Gb for Periophthalmodon schlosseri, Periophthalmus magnuspinnatus, Boleophthalmus pectinirostris and Scartelaos histophorus, respectively [10]. In the present study, genome size of A. ommaturus was estimated to be 928.01 Mb (0.928 Gb), which was in accordance with most fish.

According to the K-mer distribution, it was estimated that the repeat sequence content was about 38.31%. Most teleost fish exhibit a low repeat content, with an average of less than 30% [33], such as Gadus morhua (25.4%) [34], Gasterosteus aculeatus santaeannae (25.2%) [35] and Larimichthys crocea (16.38%) [36]. A few fish, such as Danio rerio and Salmo salar, have higher repeat sequences, reaching 52.2% [33] and 60% [31], respectively.

In the present study, there was no obvious heterozygous peak. Overall, the heterozygosity was low, and the estimated heterozygosity was about 0.17%. According to the characteristics of fish genome, draft genome can be obtained by next-generation sequencing technology.

Genome assembly and GC content

With 45, 55, 65 and 75 bp K-mers, de novo assembly was performed using all of the clean reads (Table 1). When 55 bp K-mer value was used, the length of N50 could get the maximum value. The BUSCO evaluation results showed a low score caused by lacking of sufficient genomic data of related species The intrinsic defects of the second-generation sequencing technology may be also one of the reasons for the poor assembly effect (Table 2). Genomic GC content had a significant effect on the randomness of second-generation genome sequencing. Too high (>65%) or too low (<25%) GC content will lead to sequencing bias and seriously affect the results of genomic analysis. By analyzing the sequencing data of the survey library, the GC content of A. ommaturus genome was approximately 40.88%, which was suitable for further analysis and genome assembly.

Table 1
Statistics of A. ommaturus assembled genome sequences
K-mer depth (bp)ContigsScaffolds
Total length (bp)Total numberMax length (bp)N50 (bp)N90 (bp)Total length (bp)Total numberMax length (bp)N50 (bp)N90 (bp)
45 705,413,224 1,522,798 35,678 1120 203 720,996,115 598,674 51,912 2458 436 
55 765,231,759 1,421,279 43,980 1237 216 776,582,763 717,328 60,036 2528 343 
65 819,722,014 1,395,163 40,568 1267 227 827,753,551 848,081 66,118 2443 301 
75 861,431,452 1,387,872 54260 1282 236 867,370,160 944,567 67,355 2343 283 
K-mer depth (bp)ContigsScaffolds
Total length (bp)Total numberMax length (bp)N50 (bp)N90 (bp)Total length (bp)Total numberMax length (bp)N50 (bp)N90 (bp)
45 705,413,224 1,522,798 35,678 1120 203 720,996,115 598,674 51,912 2458 436 
55 765,231,759 1,421,279 43,980 1237 216 776,582,763 717,328 60,036 2528 343 
65 819,722,014 1,395,163 40,568 1267 227 827,753,551 848,081 66,118 2443 301 
75 861,431,452 1,387,872 54260 1282 236 867,370,160 944,567 67,355 2343 283 
Table 2
A assembly evaluation summary by BUSCO
TermResults
Complete (C) 1300 (38.76%) 
Complete (C) and single-copy (S) 1285 (38.31%) 
Complete (C) and duplicated (D) 15 (0.45%) 
Fragmented (F) 1150 (34.29%) 
Missing (M) 904 (26.95%) 
Total 3354 
TermResults
Complete (C) 1300 (38.76%) 
Complete (C) and single-copy (S) 1285 (38.31%) 
Complete (C) and duplicated (D) 15 (0.45%) 
Fragmented (F) 1150 (34.29%) 
Missing (M) 904 (26.95%) 
Total 3354 

Identification of SSR

Based on the genome survey sequence, the total number of identified microsatellite or simple sequence repeats (SSRs) was 349,138 (Table 3). Among them, 89,834 sequences contained more than one microsatellite site. The dinucleotide repeats were dominant (53.70%), followed by trinucleotide repeats (35.36%), hexanucleotide repeats (4.59%), tetranucleotide repeats (4.57%) and pentanucleotide repeats (1.77%). According to the distribution and frequency of microsatellite motifs, the dinucleotide repeats had the highest number and type of repeats, which was similar to the previous studies of other fish, such as Pseudosciaena crocea [37] and Megalobrama amblycephala [38]. The frequency of repetitions decreases exponentially with the length of the repetitions, because long mutations have high mutation rates [39]. This is consistent with the result that the number of repetitions is inversely proportional to the length of repetitions reported by Chen et al. (2010) [40].

Table 3
Simple sequence repeat (SSR) distribution statistics for A. ommaturus
StatisticsDi-Tri-Tetra-Penta-Hexa-
SSR number 187,501 123,452 15,969 6,182 16,034 
Percentage 53.70% 35.36% 4.57% 1.77% 4.59% 
StatisticsDi-Tri-Tetra-Penta-Hexa-
SSR number 187,501 123,452 15,969 6,182 16,034 
Percentage 53.70% 35.36% 4.57% 1.77% 4.59% 

The AC repeat motif was the most frequent among all twelve types of dinucleotide repeat, whereas GC was the least frequent (Figure 2A). The AAT repeat motif was the most frequent among all sixty types of trinucleotide repeat (Figure 2B). The AAAT repeat motif was the most frequent among all types of tetranucleotide repeat (Figure 2C). The AATTG repeat motif was the most frequent pentanucleotide repeat (Figure 2D), and the TTCTGA was the most frequent hexanucleotide repeat (Figure 2E). There were only eight types of repeats detected in hexanucleotide repeats, and the number of repeats detected in dinucleotide repeats, trinucleotide repeats, tetranucleotide repeats and pentanucleotide repeats were higher than 10, even up to 35.

The distribution and frequency of microsatellite motifs

Figure 2
The distribution and frequency of microsatellite motifs

(A) Frequency of different dinucleotide microsatellite motifs. (B) Frequency of different trinucleotide microsatellite motifs. (C) Frequency of different tetranucleotide microsatellite motifs. (D) Frequency of different pentabase microsatellite motifs. (E) Frequency of different hexanucleotide microsatellite motifs.

Figure 2
The distribution and frequency of microsatellite motifs

(A) Frequency of different dinucleotide microsatellite motifs. (B) Frequency of different trinucleotide microsatellite motifs. (C) Frequency of different tetranucleotide microsatellite motifs. (D) Frequency of different pentabase microsatellite motifs. (E) Frequency of different hexanucleotide microsatellite motifs.

Close modal

To provide more accurate information for SSR primer verification in future research, we selected 4996 pairs of sequences suitable for synthesizing microsatellite primers from the 89,834 sequences, with the following criterions: repetition times of SSR units are more than 6; length of SSR units is 3-10 bp; expected length of PCR products is 130–300 bp; primers that repeat the same four bases continuously are excluded. The selected SSR primers information were provided in the Supplementary Material.

In summary, for A. ommaturus, the genome size was approximately 928.01 Mb, the proportion of repeats was approximately 38.31%, the heterozygosity rate was approximately 0.17%, and the GC content of the genome was approximately 40.88%. From the perspective of the basic structural characteristics of the genome, it was a simple genome. This paper could be conductive to the construction of subsequent fine map of the genome.

All research data are included in the main text of the manuscript. Entire read sets were deposited in the short read archive (SRA) databank (http://www.ncbi.nlm.nih.gov/sra/) and are available under accession number PRJNA658176.

The authors declare that there are no competing interests associated with the manuscript.

This study was supported by the National Key R & D Program of China [grant number 2018YFD0900905] and National Natural Science Foundation of China [grant numbers 41506158 and 41776171].

Conceptualization: B.C. Software: B.C. and F.R. Validation: N.S. and T.G. Investigation: B.C. and Z.S. Resources: B.C. and N.S. Writing—original draft preparation: B.C. Writing—review and editing: B.C. and N.S. Project administration: Z.S. Funding acquisition: N.S.

We thank Dr Shengyong Xu and Mr. Jian Zheng for English editing.

NGS

next-generation sequencing

SRA

short read archive

SSR

simple sequence repeat

1.
Feng
G.
,
Zhang
L.
,
Zhuang
P.
,
Chen
J.
,
Zhang
T.
and
Hou
J.
(
2010
)
Morphological character, individual fecundity and ovary histology of Acanthogobius ommaturus in Yangtze River estuary
.
Mar. Fish.
3
,
257
263
2.
Wu
H.
and
Zhong
J.
(
2008
)
Osteichthys, Perciformes, Gobioidei
.
Fauna Sinica
1st ed
Beijing, China
Science Press
.
Osteichthys, Perciformes 5, 136-718
3.
Song
C.
,
Feng
Z.
,
Li
C.
,
Sun
Z.
,
Gao
T.
,
Song
N.
et al.
(
2019
)
Profile and development of microsatellite primers for Acanthogobius ommaturus based on high-throughput sequencing technology
.
J. Oceanol. Limnol.
4.
Tong
C.
(
2012
)
Characteristics of the fish assemblages in the intertidal salt marsh zone and adjacent mudflat in the Yangtze Estuary
.
Acta Ecol. Sin.
32
,
6501
6510
5.
Li
F.
,
Zhang
H.
,
Lv
Z.
,
Xu
B.
and
Zheng
L.
(
2013
)
Species composition and community diversity of nekton in Laizhou Bay, China
.
Biodivers. Sci.
21
,
537
546
6.
Liu
X.
,
Li
Y.
,
Wang
W.
,
Duan
Y.
,
Chai
Y.
and
Dong
J.
(
2015
)
The fish community structure and diversity in the north of Liaodong Bay
.
J. Fish. China
39
,
1155
1165
7.
Huang
X.
,
Zhang
L.
,
Zhuang
P.
,
Jiang
Q.
,
Yao
Z.
and
Liu
J.
(
2010
)
Effects of cryopreservation on enzyme activity in Acanthogobius ommaturus ovum
.
Mar. Fish.
32
,
48
53
8.
Yum
J.
(
2008
)
Characterization of Lactate Dehydrogenase in Acanthogobius hasta
.
J. Life Sci.
18
,
264
272
9.
Insoo
S.
and
Hong
J.
(
2006
)
Feeding Ecology of Gavelin Goby (Acanthogobius hasta) and Fine Spot Goby (Chaeturichtys stigmatias) in the Jangbong Tidal Flat, Incheon, Korea
.
Korean J. Fisheries Aquat. Sci.
39
,
165
179
10.
You
X.
,
Bian
C.
,
Zan
Q.
,
Xu
X.
,
Liu
X.
,
Chen
J.
et al.
(
2014
)
Mudskipper genomes provide insights into the terrestrial adaptation of amphibious fishes
.
Nat. Commun.
5
,
1
8
11.
Xu
S.
,
Song
N.
,
Xiao
S.
and
Gao
T.
(
2020
)
Whole genome survey analysis and microsatellite motif identification of Sebastiscus marmoratus
.
Biosci. Rep.
5
,
1
8
12.
Lu
X.
,
Luan
S.
,
Kong
J.
,
Hu
L.
,
Mao
Y.
and
Zhong
S.
(
2017
)
Genome-wide mining, characterization, and development of microsatellite markers in Marsupenaeus japonicus by genome survey sequencing
.
Chin. J. Oceanol. Limn.
35
,
203
214
13.
Capobianchi
M.R.
,
Giombini
E.
and
Rozera
G.
(
2013
)
Next-generation sequencing technology in clinical virology
.
Clin. Microbiol. Infec.
19
,
15
22
14.
Motalebipour
E.Z.
,
Kafkas
S.
,
Khodaeiaminjan
M.
,
Çoban
N.
and
Gözel
H.
(
2016
)
Genome survey of pistachio (Pistacia vera L.) by next generation sequencing: development of novel SSR markers and genetic diversity in Pistacia species
.
Genomics
17
,
998
[PubMed]
15.
Englbrecht
C.C.
,
Freyhof
J.
,
Nolte
A.
,
Rassmann
K.
,
Schliewen
U.
and
Tautz
D.
(
2000
)
Phylogeography of the bullhead Cottus gobio (Pisces: Teleostei: Cottidae) suggests a pre-Pleistocene origin of the major central European populations
.
Mol. Ecol.
9
,
709
722
[PubMed]
16.
Whitehead
A.
,
Anderson
S.L.
,
Kuivila
K.M.
,
Roach
J.L.
and
May
B.
(
2003
)
Genetic variation among interconnected populations of Catostomus occidentalis: implications for distinguishing impacts of contaminants from biogeographical structuring
.
Mol. Ecol.
12
,
2817
2833
[PubMed]
17.
Song
N.
,
Zhang
X.M.
,
Sun
X.F.
,
Yanagimoto
T.
and
Gao
T.X.
(
2010
)
Population genetic structure and larval dispersal potential of spottedtail goby Synechogobius ommaturus in the north-west Pacifific
.
J. Fish Biol.
77
,
388
402
[PubMed]
18.
Song
N.
,
Zhang
X.
and
Gao
T.
(
2010
)
Genetic diversity and population structure of spottedtail goby(Synechogobius ommaturus) based on AFLP analysis
.
Biochem. Syst. Ecol.
38
,
1089
1095
19.
Portis
E.
,
Portis
F.
,
Valente
L.
,
Moglia
A.
,
Barchi
L.
,
Lanteri
S.
et al.
(
2016
)
A genome-wide survey of the microsatellite content of the globe artichoke genome and the development of a web-based database
.
PLoS ONE
11
,
e0162841
[PubMed]
20.
Adelyna
M.N.
,
Jung
H.
,
Chand
V.
,
Mather
P.B.
and
Azizah
M.S.
(
2016
)
A genome survey sequence (GSS) analysis and microsatellite marker development for Indian mackerel, Rastrelliger kanagurta, using ion torrent technology
.
Meta Gene
10
,
67
72
21.
Castoe
T.A.
,
Poole
A.W.
,
de Koning
A.J.
,
Jones
K.L.
,
Tomback
D.F.
,
Oyler-McCance
S.J.
et al.
(
2012
)
Rapid microsatellite identification from Illumina paired-end genomic sequencing in two birds and a snake
.
PLoS ONE
7
,
e30953
[PubMed]
22.
Zhou
W.
,
Hu
Y.
,
Sui
Z.
,
Fu
F.
,
Wang
J.
,
Chang
L.
et al.
(
2013
)
Genome survey sequencing and genetic background characterization of Gracilariopsis lemaneiformis (Rhodophyta) based on next-generation sequencing
.
PLoS ONE
8
,
e69909
[PubMed]
23.
Li
G.
,
Song
L.
,
Jin
C.
,
Li
M.
,
Gong
S.
and
Wang
Y.
(
2019
)
Genome survey and SSR analysis of Apocynum venetum
.
Biosci. Rep.
39
,
BSR20190146
[PubMed]
24.
Guillaume
M.
and
Carl
K.
(
2011
)
A fast, lock-free approach for efficient parallel counting of occurrences of k-mers
.
Bioinformatics
27
,
764
770
[PubMed]
25.
Gregory
W.V.
,
Fritz
J.S.
,
Maria
N.
,
Charles
J.U.
,
Han
F.
,
James
G.
et al.
(
2017
)
GenomeScope: fast reference-free genome profiling from short reads
.
Bioinformatics
33
,
2202
2204
[PubMed]
26.
Luo
R.
,
Liu
B.
,
Xie
Y.
,
Li
Z.
,
Huang
W.
,
Yuan
J.
et al.
(
2012
)
SOAPdenovo2: an empirically improved memory-efficient short-read de novo assembler
.
Gigascience
1
,
2047
217X
27.
Simão
F.A.
,
Waterhouse
R.M.
,
Ioannidis
P.
,
Kriventseva
E.V.
and
Zdobnov
E.M.
(
2015
)
BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs
.
Bioinformatics
31
,
3210
3212
[PubMed]
28.
Untergasser
A.
,
Cutcutache
I.
,
Koressaar
T.
,
Ye
J.
,
Faircloth
B.C.
,
Remm
M.
et al.
(
2012
)
Primer3-new capabilities and interfaces
.
Nucleic Acids Res.
40
,
e115
[PubMed]
29.
Koressaar
T.
and
Remm
M.
(
2007
)
Enhancements and modifications of primer design program Primer3
.
Bioinformatics
23
,
1289
1291
[PubMed]
30.
Chen
S.
,
Xu
W.
and
Liu
Y.
(
2019
)
Fish genomic research: Decade review and prospect
.
J. Fish China
43
,
1
14
31.
Lien
S.
,
Koop
B.F.
,
Sandve
S.R.
,
Miller
J.R.
,
Kent
M.P.
,
Nome
T.
et al.
(
2016
)
The Atlantic salmon genome provides insights into rediploidization
.
Nature
533
,
200
212
[PubMed]
32.
Berthelot
C.
,
Brunet
F.
,
Chalopin
D.
,
Juanchich
A.
,
Bernard
M.
,
Noël
B.
et al.
(
2014
)
The rainbow trout genome provides novel insights into evolution after whole-genome duplication in vertebrates
.
Nat. Commun.
5
,
3657
[PubMed]
33.
Howe
K.
,
Clark
M.D.
,
Torroja
C.F.
,
Torrance
J.
,
Berthelot
C.
,
Muffato
M.
et al.
(
2013
)
The zebrafish reference genome sequence and its relationship to the human genome
.
Nature
496
,
498
503
[PubMed]
34.
Star
B.
,
Nederbragt
A.J.
,
Jentoft
S.
,
Grimholt
U.
,
Malmstrøm
M.
,
Gregers
T.F.
et al.
(
2011
)
The genome sequence of Atlantic cod reveals a unique immune system
.
Nature
477
,
207
210
[PubMed]
35.
Jones
F.C.
,
Grabherr
M.G.
,
Chan
Y.F.
,
Russell
P.
,
Mauceli
E.
,
Johnson
J.
et al.
(
2012
)
The genomic basis of adaptive evolution in threespine sticklebacks
.
Nature
484
,
55
61
[PubMed]
36.
Wu
C.
,
Zhang
D.
,
Kan
M.
,
Lv
Z.
,
Zhu
A.
,
Su
Y.
et al.
(
2014
)
The draft genome of the large yellow croaker reveals well-developed innate immunity
.
Nat. Commun.
5
,
5227
[PubMed]
37.
Li
Q.
,
Li
Z.
,
Dai
G.
,
Cao
Y.
,
Chen
X.
,
Chen
L.
et al.
(
2014
)
Isolation and characterization of eleven microsatellite loci in the marbled rockfish, Sebastiscus marmoratus (Scorpaenidae)
.
Conserv. Genet. Resour
6
,
53
55
38.
Zeng
C.
,
Gao
Z.
,
Luo
W.
,
Liu
X.
,
Wang
W.
,
Zhang
X.
et al.
(
2013
)
Characteristics of microsatellites in blunt snout bream (Megalobrama amblycephala) EST sequences using 454 FLX
.
Acta Hydrobiol. Sin.
37
,
982
988
39.
Katti
M.V.
,
Ranjekar
P.K.
and
Gupta
V.S.
(
2001
)
Differential distribution of simple sequence repeats in eukaryotic genome sequences
.
Mol. Biol. Evol.
18
,
1161
1167
[PubMed]
40.
Chen
M.
,
Tan
Z.
,
Zeng
G.
and
Peng
J.
(
2010
)
Comprehensive analysis of simple sequence repeats in pre-miRNAs
.
Mol. Biol. Evol.
27
,
2227
2232
[PubMed]
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).