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

Transcriptome Analysis of Pacific White Shrimp (Litopenaeus vannamei) Hepatopancreas in Response to Taura Syndrome Virus (TSV) Experimental Infection

  • Digang Zeng,

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

  • Xiuli Chen,

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

  • Daxiang Xie,

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

  • Yongzhen Zhao,

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

  • Chunling Yang,

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

  • Yongmei Li,

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

  • Ning Ma,

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

  • Min Peng,

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

  • Qiong Yang,

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

  • Zhenping Liao,

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

  • Hui Wang,

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

  • Xiaohan Chen

    gxnnscs@126.com

    Affiliation Guangxi Key Laboratory of Aquatic Genetic Breeding and Healthy Aquaculture, Guangxi Institute of Fisheries, Nanning, China

Abstract

Background

The Pacific white shrimp, Litopenaeus vannamei, is a worldwide cultured crustacean species with important commercial value. Over the last two decades, Taura syndrome virus (TSV) has seriously threatened the shrimp aquaculture industry in the Western Hemisphere. To better understand the interaction between shrimp immune and TSV, we performed a transcriptome analysis in the hepatopancreas of L. vannamei challenged with TSV, using the 454 pyrosequencing (Roche) technology.

Methodology/Principal Findings

We obtained 126919 and 102181 high-quality reads from TSV-infected and non-infected (control) L. vannamei cDNA libraries, respectively. The overall de novo assembly of cDNA sequence data generated 15004 unigenes, with an average length of 507 bp. Based on BLASTX search (E-value <10−5) against NR, Swissprot, GO, COG and KEGG databases, 10425 unigenes (69.50% of all unigenes) were annotated with gene descriptions, gene ontology terms, or metabolic pathways. In addition, we identified 770 microsatellites and designed 497 sets of primers. Comparative genomic analysis revealed that 1311 genes differentially expressed in the infected shrimp compared to the controls, including 559 up- and 752 down- regulated genes. Among the differentially expressed genes, several are involved in various animal immune functions, such as antiviral, antimicrobial, proteases, protease inhibitors, signal transduction, transcriptional control, cell death and cell adhesion.

Conclusions/Significance

This study provides valuable information on shrimp gene activities against TSV infection. Results can contribute to the in-depth study of candidate genes in shrimp immunity, and improves our current understanding of this host-virus interaction. In addition, the large amount of transcripts reported in this study provide a rich source for identification of novel genes in shrimp.

Introduction

Taura syndrome virus (TSV) is a contagious viral disease of penaeid shrimp [1]. Over the last two decades, TSV has seriously threatened the shrimp aquaculture industry and caused serious economic losses [2], [3]. In cultured Pacific white shrimp (Litopenaeus vannamei), which has become the major aquacultured crustacean species in the Western Hemisphere [4], TSV can cause a cumulative mortality ranged from 40 to >90% [5]. The survival shrimp of TSV infections may carry the virus for life [6], [7]. TSV was first discovered in South America, but has later spread to North America, Hawaii and Asia [8][11]. TSV is a small, simple RNA virus. The genome of TSV is a positive-sense singlestranded RNA of 10.2 kb containing two open reading frames [12]. Although considerable progress has been made in the molecular characterization of TSV [13], no effective cure for this disease has yet been found [14]. Shrimp lack an acquired immune system. Their defense is considered to depend entirely on an innate, non-adaptive mechanism to defense invasion by pathogens [15]. An understanding of the host-pathogen interaction will be helpful in controlling infectious diseases in shrimp[15]. Therefore, the molecular response of shrimp to viral infection is becoming an increasingly important subject for study [16]. Recently, the molecular mechanism of the shrimp-virus interaction has made notable progress [17], many genes involved in viral infection in shrimp were found, such as lectins, antimicrobial peptide, blue blood protein and superoxide dismutase [18][21]. Most of these genes were identificated using the suppression subtractive hybridization (SSH) technology. SSH is an effective approach for identifying differentially expressed genes among different biological processes[22]. However, subtractive hybridization does not provide a quantitative measure of expression differences, and its experimental results often contain a large number of false positives [23]. Recently, the high-throughput sequencing technologies, such as the Illumina Genome Analyzer, the Applied Biosystems Solid platform, and the 454 Life Sciences (Roche) pyrosequencing platform, provide a rapid and high-throughput method to identify differentially expressed genes and their expression profile [24][26].

Identification of host genetic factors in response to pathogen is of great significance for shrimp breeding and production. However, information on the host genes involved in TSV pathogenesis is still limited. To our knowledge, there is no previous report of isolating genes that are involved in TSV infection. In this study, we performed a transcriptome analysis of the hepatopancreas of L. vannamei challenged with TSV, using a high-throughput sequencing method (Roche 454 pyrosequencing). The aim of this study was to discover new genes involved in TSV infection, and better understand the virus-host interaction. Furthermore, the high-throughput sequencing will produce a large number of transcripts in this study, providing a strong basis for future genomic research on shrimp.

Materials and Methods

Shrimp, Virus and Challenge

L. vannamei from a specific pathogen-free (SPF) line (High Health Aquaculture, Kona, Hawaii, USA) were used in this study. The shrimp (11–12 g body weight) were provided from the National and Guangxi Vannamei Genetic Breeding Center, Guangxi Province, China, and held in the environmentally controlled 1000-liter glass saltwater tanks (32-ppt salinity, 25 to 26°C) and fed an artificial pellet feed. The shrimp were randomly sampled and tested by PCR for certify to be free of TSV by PCR[5]. In the challenge experiment, there were two shrimp groups: 1 TSV challenge group, and 1 negative control group (20 shrimp each group). The challenge group was fed once a day for 3 consecutive days with minced virus infected tail tissue at 10% of their body weight. In parallel, the negative control group was fed once a day throughout the test period of 3 days with minced PCR-confirmed [5] healthy tail tissue at 10% of their body weight. At 72 hours after infection, the hepatopancreas tissues of shrimp were collected in cryotubes and stored in liquid nitrogen for later RNA extraction.

RNA Extraction, cDNA Library Construction, and Deep Sequencing

Total RNAs were extracted from TSV-infected and non-infected shrimp hepatopancreas using TriReagent (Qiagen), and the mRNAs were purified from the total RNAs using the PolyATtract mRNA isolation systems (Promega) following the manufacturer’s instructions. Integrity and size distribution were checked with Bioanalyzer 2100 (Agilent technologies, USA). Equal amounts of the high-quality mRNA samples from each group were then pooled for cDNA synthesis and sequencing. The normalized cDNA library was prepared following the 454 mRNA pyrosequencing sample preparation procedure (Roche, IN, USA). Library construction and pyrosequencing was carried out by Beijing Autolab Biotechnology Co., Ltd. on a 454 GS FLX system (Roche).

De novo Assembly and Functional Annotation

Raw sequencing reads were quality trimmed, and adaptor sequences were removed before the assembly. After removal of low quality reads, processed reads were assembled using CAP3 software with default parameters[27]. The overall assembly was performed using the combined sequence data for both the TSV-infected sample and the non-infected sample. The contigs and singletons were generally referred to as unigenes. Subsequently, the unigenes were subjected to BLASTX similarity search against NCBI non-redundant protein database and the swissprot database using BLASTALL programs with an E-value threshold of 10−5[28]. All annotated unigenes were used to determine the COG term, GO term and KEGG pathway with a cut-off E-value of 10−5 using BLASTX[29], [30].

Identification of Differentially Expressed Genes

For differential gene expression analyzes, RPKM (reads per kilobase per million reads) were used as the value of normalized gene expression levels[31]. Statistical comparison of RPKM values between the TSV-infected sample and the non-infected sample was conducted using a web tool IDEG6 (http://telethon.bio.unipd.it/bioinfo/IDEG6_form/) [32]. FDR (false discovery rate) <0.001 was used as the threshold of P-value in multiple test to judge the significance of gene expression difference [33]. Genes were considered differentially expressed in a given library when the p-value ≤0.001 and a greater than two-fold change (absolute value of log2 ratio >1) in expression across libraries was observed.

Quantitative RT-PCR Analysis

To validate our 454 sequencing data, six differentially expressed L. vannamei genes (cathepsin-L, arginine kinase, fatty acids binding protein, alternative splicing factor, sorbitol dehydrogenase and hemocyanin) were selected for quantitative RT-PCR analysis, using the same RNA samples as for the transcriptome profiling. Primers were designed using the Primer5 software (Premier Biosoft International) (Table S1). First strand cDNA was synthesized from 1 μg of RNA using M-MuLV reverse transcriptase (Qiagen). The qPCR reaction mixture (20 μL) consisted of 2× Power SYBR Green PCR Master mix, 0.9M each of the forward and reverse primers, and 1 μL of template cDNA. PCR amplification was performed under the following conditions: 50°C for 2 min and 95°C for 30 s, followed by 40 cycles of 95°C for 15 s and 62°C for 1 min, and a final extension at 72°C for 5 min.

Identification of Microsatellites

All the assembled cDNA contigs from both the infected library and the control library were used for identification of microsatellites. All types of microsatellites from dinucleotides to hexanucleotides were detected using the MISA software [34] with default parameters (for all repeat types, minimum total length = 15 bp and minimum repeats = 3). Primers were designed using the primer3 software [35].

Results

Pyrosequencing and Assembly

To identify the genes involved in L. vannamei response to TSV infection, we created two cDNA libraries from pooled mRNAs extracted from the hepatopancreas of TSV-infected and non-infected (control) groups, respectively. The two libraries were subjected to a pyrosequencing run on the 454 GS FLX system, resulting in 131745 (TSV-infected sample) and 110721 (non-infected sample) raw reads, respectively. Files containing these data were deposited in the Short Read Archive of the National Center for Biotechnology Information (NCBI) with accession numbers of SRR554365 (TSV-infected) and SRR556131 (non-infected). After filtering for adaptors and low-quality sequences, the TSV-infected library generated 126919 cleaned reads, ranging from 41 bp to 620 bp, with the average length of 367 bp and N50 length of 454 bp (Table S2). In the non-infected library, a total of 102181 cleaned reads were obtained, ranging from 45 bp to 619 bp, with the average length of 364 bp and N50 length of 454 bp (Table S3). The overall assembly was performed using the combined cleaned reads from the two libraries. De novo assembly using the CAP3 software produced 15004 unigenes (including contigs and singletons) with an average length of 507 bp, ranging from 42 to 8750 bp (Table 1).

Functional Annotation

All unigenes were compared with the Swiss-Prot and the NCBI non-redundant (NR) protein databases for functional annotation by using BLASTX with an e-value threshold of 10−5. Among the 15004 unigenes from both the TSV-infected library and the non-infected library, 4400 (29.33%) showed significant matches (E-value ≤10−5) in the Swiss-Prot database. An additional 10412 (69.39%) unigenes showed significant matches (E-value ≤10−5) in the NR database. In total, 10425 (69.50%) unigenes were annotated in Swiss-Prot or NR database.

Gene ontology (GO) analysis was performed with the unigenes from both the infected library and the control library. A total of 6567 and 6604 unigenes map to biological processes, 2977 and 2828 unigenes map to molecular functions, and 5206 and 5222 unigenes map to cellular components in the TSV-infected library and non-infected control library, respectively (Table S4). The functional distribution of the genes of the two libraries was similar. In both libraries, most of the corresponding biological process genes were involved in metabolic processes and cellular processes. Most of the molecular function genes were associated with catalytic activity, binding, and molecular transducer activity; most of the cellular component genes encode proteins associated with cell, parts of cell and cell organelles (Figure 1).

thumbnail
Figure 1. GO annotations of unigenes in the TSV-infected and non-infected Litopenaeus vannamei cDNA libraries.

Most unigenes can be divided into three major categories, including biological process, cellular component, and molecular function.

https://doi.org/10.1371/journal.pone.0057515.g001

We also searched the annotated sequences for the genes involved in COG classification. Among the 25 COG categories, the cluster for ‘General function prediction only’ (15.388%) represented the largest group, followed by the ‘Posttranslational modification, protein turnover, chaperones’ (12.83%) and ‘Translation, ribosomal structure and biogenesis’ (7.227%) clusters (Figure 2).

thumbnail
Figure 2. COG function classification of unigenes in the TSV-infected and non-infected Litopenaeus vannamei cDNA libraries.

All putative proteins were aligned to the COG database and can be classified functionally into at least 25 molecular families.

https://doi.org/10.1371/journal.pone.0057515.g002

To identify the biological pathways that are active in L. vannamei, we mapped all the unigenes to the referential canonical pathways in the KEGG database. A total of 14496 unigenes were assigned to 174 KEGG pathways (Table S5). The top 10 pathways are metabolic pathways (1762 members), phagosome (1035 members), focal adhesion (815 members), tight junction (779 members), adherens junction (768 members), biosynthesis of secondary metabolites (624 members), lysosome (374 members), ribosome (320 members), oxidative phosphorylation (267 members), and tyrosine metabolism (217 members).

In summary, for functional annotation, all the unigenes were searched against NR, Swissprot, GO, COG, and KEGG databases by BLASTX with a cut-off E-value of 10−5. By this method, 10425 unigenes (69.50% of all unigenes) returned an above cut-off BLAST result (Table 2). Of them, 10412 unigenes were annotated by NR (69.39%), and 4400 (29.33%), 3397 (22.64%), 3979 (26.52%), and 3393 (22.61%) unigenes by SwissProt, GO, COG and KEGG, respectively.

Identification of Differentially Expressed Genes

To identify differentially expressed genes potentially involved in TSV infection, we constructed two normalized cDNA libraries from pooled mRNAs extracted from the hepatopancreas of TSV-infected and non-infected groups, respectively. Subsequently, these libraries were sequenced by 454 GS FLX technology. Comparison of gene expression revealed 1311 genes differentially expressed in TSV-infected shrimp compared to the control, including 559 up-regulated genes and 752 down-regulated genes. The number of down-regulated genes is much larger than that of up-regulated genes, which might be consistent with the observation that viral-infected individuals of shrimp are less active. Among the 1311 differentially expressed genes, 1061(80.93%) genes were well annotated, whereas the remaining 250(19.07%) genes had low sequence homology to known sequences in public databases, suggesting that they might be putative novel genes in L. vanname involved in the response to TSV infection. All differentially expressed unigenes with their Nr, Nt, Swissprot, GO, COG, KEGG and ORF analysis are presented in additional Table S6.

To validate our RNA-seq results, six differentially regulated genes with the different total transcript reads (range 44–1076) were selected for quantitative real time-PCR (qRT-PCR) analysis. The results indicate that the qRT-PCR analysis of the relatively high abundant genes (>500 reads) agrees well with the 454 sequencing analysis. For example, based on 454 sequencing analysis, cathepsin-L (CATL), arginine kinase (AK) and fatty acids binding protein (FABP) were differentially regulated 1.48, −1.26 and −2.53 log2-fold, respectively, and showed 1.25, −1.62 and −2.31 log2-fold changes, respectively in qRT-PCR analyses (Figure 3). However, the qRT-PCR analysis of the relatively low abundant genes (<500 reads), including alternative splicing factor (ASF), sorbitol dehydrogenase (SDH) and hemocyanin (HCS), do not match the 454 sequencing analysis perfectly, even if it shows similar trends in up- or down-regulation of genes analysised by 454 sequencing (Figure 3). Nevertheless, qRT-PCR analysis confirmed the change direction detected by the 454 sequencing analysis.

thumbnail
Figure 3. Comparison of the expression profiles of selected genes as determined by 454 sequencing (blue) and qRT-PCR (red).

Target gene abbreviations are as follows: CATL - cathepsin L, AK - arginine kinase, FABP - fatty acids binding protein, ASF - alternative splicing factor, SDH - sorbitol dehydrogenase, HCS - hemocyanin.

https://doi.org/10.1371/journal.pone.0057515.g003

Candidate Genes Involved in L. vanname Immune Response

Among the genes that were found to be differentially expressed in the TSV-infected shrimp compared to non-infected controls, several are involved in various processes of animal immune response (Table 3). These were classified under 8 functions, including antiviral proteins, antimicrobial proteins, proteases, protease inhibitors, signal transducers and transcription factors, cell death and cell adhesion. Antiviral proteins are proteins that are induced by interferon in virus-infected human or animal cells and mediates interferon inhibition of virus replication [36]. Among the differentially expressed genes homologous to antiviral proteins, we found that a homolog of Zinc finger CCCH-type antiviral protein was significantly up regulated in TSV-infected shrimp compared to non-infected controls. The up-regulation of this gene after viral infection suggests that it may be involved in shrimp immune response. Antimicrobial proteins are an important component of the natural defenses of most living organisms against invading pathogens [15]. They interfere with microbial integrity or metabolism by targeting structures or nutrients specific to microbes [37]. Of the antimicrobial proteins identified in this study, Lysozyme and Histone H2A were significantly up regulated after TSV infection, indicating they may play important roles in shrimp defense against virus. Also of interest for the study of virus-host interactions is the identification of genes involved in signal transduction, as signal transduction molecules have been suggested to play important roles in viral recognition and replication [38]. We identified 8 differentially expressed genes involved in signal transduction, including P38 mapk, Map kinase-interacting serine/threonine, Serine/threonine-protein phosphatase alpha-1 isoform, Senescence-associated protein, Transmembrane BAX inhibitor motif-containing protein 4, C-type lectin, Innexin, and Fatty-acid amide hydrolase 2. All of these genes were up regulated. The other category of genes that are involved in transcriptional control, cell adhesion and cell death processes, may also play important roles during the TSV infection, as processes regulated by these genes have been suggested to modulate phagocytic events, cellular remodeling, recruitment of immune cells to sites of insult, and extracellular immune cascades such as the melanization response [39].

thumbnail
Table 3. Candidate genes involved in L. vanname immune response.

https://doi.org/10.1371/journal.pone.0057515.t003

Identification of Microsatellites

Microsatellites (or simple sequence repeats, SSRs) are repetitive sequence motifs of 1–6 bp [40], [41]. Although they are widely used as molecular markers due to their variability and abundance in the genome, codominant expression and inheritance in a Mendelian fashion [42], only a limited number of microsatellite sequences have been reported for L.vannamei [43]. In this study, we obtained 770 microsatellites, of which 23.90% were di-nucleotide repeats (184), followed by 36.88% tri-nucleotide repeats (284) and 36.23% tetra-nucleotide repeats (279), as well as 2.99% penta-nucleotide repeats (23) (Table S7). We also designed 497 primer sets using the Primer3 program (Table S8). These identified microsatellites have potential utility to genetic mapping, population structure and gene flow studies of L. vannamei.

Discussion

Taura syndrome (TS) is a major cause of shrimp mortality in cultured L. vannamei in the Americas and Asia [1]. Although there are many published reports of characterization and detection of TSV, little is known about the interaction between this virus and shrimp. Understanding the interaction between host and its pathogen is useful, not only for studies on the molecular immune mechanisms, but also for agricultural practice that aims to provide a theoretical basis for developing effective strategies to prevent viral disease. Roche 454 RNA sequencing (RNA-Seq) is a powerful new method for discovering novel genes and investigating gene expression patterns, especially in non-model organisms that do not have sequenced genomes[44]. Like many other crustacean species with significant economic value, L. vannamei lacks a complete genome sequences and most other genetic tools and resources. In this study, we used the 454 RNA-Seq to investigate the gene expression changes associated with the TSV infection. We identified a total of 15004 unigenes in L. vannamei, 4579(30.52%) of which were new transcripts compared to known genes in public databases. Comparative analysis of transcriptome changes between TSV-infected and non-infected shrimp revealed 1311 differentially expressed genes, of which 559 genes were up-regulated and 752 down-regulated. Our sequencing data analyses indicate that TSV infection has a significant impact on the transcriptome profile of L. vannamei hepatopancreas.

Among the differentially expressed genes found in this study, several had been previously reported to be involved in the shrimp response against white spot syndrome virus (WSSV), such as C-type lectin and hemocyanin [15], [17], [45][48]. Animal C-type lectins play important roles in innate immunity to recognize and eliminate pathogens efficiently[49]. In invertebrates, C-type lectins are involved in non-self immune recognition and pathogen phagocytosis through opsonization[50]. Several studies reported the expression of C-type lectins in shrimp hepatopancreas was greatly affected after challenge by WSSV[20], [51][53]. In this study, we found 15 unigenes homologous to C-type lectins, and their expression exhibited to change significantly after TSV infection. Hemocyanin is another well-known immune-related gene previously reported to be involved in viral infection [54]. Hemocyanins are the oxygen-transporting proteins in arthropods and molluscs[55]. Hemocyanins have the defense-related functions that are mediated through phenoloxidase activity. Several previous studies reported that hemocyanins in shrimp were greatly over expressed during WSSV infection[56][58]. Similarly, several other previously reported differentially expressed genes, such as heat shock protein, lysozyme and fatty acid-binding protein, strongly up-regulated in shrimp challenged with WSSV[59][61]. In the present study, up-regulation of these genes have also been observed in shrimp challenged with TSV. It suggests that these genes might have the similar expression pattern in response to virus infection, regardless of the virus species.

Although some of the differentially expressed genes found in this study had not been previously reported to be involved in virus-host interaction, they were annotated in the pathway known to be involved in various processes of animal defense against pathogens, such as cell death/apoptosis and mitogen-activated protein kinase (MAPK). Cell death/apoptosis pathway is known to be related to the cell hypersensitivity response, blocking pathogen progression and systemic resistance[62], [63]. Among the genes involved in cell death/apoptosis, we found lysosomal aspartic protease, ATP binding cassette transmembrane transporter, caspase, beclin protein and apoptosis-inducing factor 1. Such genes may respond to the viral infection through controlling the extent of the cell death in the defense response. MAPK is another noteworthy pathway that was activated during virus infection and contributed to virus replication in animal or plant cells[64]. Among the differentially expressed genes, we found map kinase-interacting serine/threonine-protein kinase 2, heat shock protein 70, P38 mapk and polyubiquitin-C shared homology to signaling molecules of the MAPK pathway. These genes are likely to be involved in response to the TSV infection.

In summary, we employed the 454 pyrosequencing technique to investigate the transcriptome profile of L. vanname challenged with TSV. Comparative transcriptome analysis between TSV-infected and control groups revealed significant differences in gene expression. Although the molecular functions of some genes and their associated pathways remain largely unknown, this study provides valuable information on the antiviral mechanism in shrimp and the role of the differentially expressed genes in response to TSV infection. Furthermore, the large number of transcripts and molecular markers obtained in this study provides a strong basis for future genomic research on shrimp.

Supporting Information

Table S1.

Primers used in qRT-PCR for validation of differentially expressed genes.

https://doi.org/10.1371/journal.pone.0057515.s001

(XLS)

Table S2.

Length distribution of cleaned reads in the TSV infected library.

https://doi.org/10.1371/journal.pone.0057515.s002

(XLS)

Table S3.

Length distribution of cleaned reads in the non-infected library.

https://doi.org/10.1371/journal.pone.0057515.s003

(XLS)

Table S6.

List of differentially expressed genes with Nr, Nt, Swissprot, GO, COG, KEGG and ORF analysis.

https://doi.org/10.1371/journal.pone.0057515.s006

(XLS)

Table S8.

PCR primers used for microsatellites.

https://doi.org/10.1371/journal.pone.0057515.s008

(XLS)

Author Contributions

Conceived and designed the experiments: XC. Performed the experiments: DZ XC DX YZ CY YL NM MP. Analyzed the data: DZ QY ZL HW. Contributed reagents/materials/analysis tools: QY ZL HW. Wrote the paper: DZ.

References

  1. 1. Ongvarrasopone C, Saejia P, Chanasakulniyom M, Panyim S (2011) Inhibition of Taura syndrome virus replication in Litopenaeus vannamei through silencing the LvRab7 gene using double-stranded RNA. Arch Virol 156: 1117–1123.
  2. 2. Naylor RL, Goldburg RJ, Primavera JH, Kautsky N, Beveridge MC, et al. (2000) Effect of aquaculture on world fish supplies. Nature 405: 1017–1024.
  3. 3. Lightner DV, Redman RM, Poulos BT, Nunan LM, Mari JL, et al. (1997) Risk of spread of penaeid shrimp viruses in the Americas by the international movement of live and frozen shrimp. Rev Sci Tech 16: 146–160.
  4. 4. Valles-Jimenez R, Cruz P, Perez-Enriquez R (2004) Population genetic structure of Pacific white shrimp (Litopenaeus vannamei) from Mexico to Panama: microsatellite DNA variation. Mar Biotechnol (NY) 6: 475–484.
  5. 5. Nunan LM, Poulos BT, Lightner DV (1998) Reverse transcription polymerase chain reaction (RT-PCR) used for the detection of Taura syndrome virus (TSV) in experimentally infected shrimp. Dis Aquat Organ 34: 87–91.
  6. 6. Hasson KW, Lightner DV, Mohney LL, Redman RM, Poulos BT, et al. (1999) Taura syndrome virus lesion development and the disease cycle in the pacific white shrimp Penaeus vannamei. Dis Aquat Organ 36: 81–93.
  7. 7. Lotz JM (1997) Effect of host size on virulence of Taura virus to the marine shrimp Penaeus vannamei (Crustacea: Penaeidae). Dis Aquat Organ 30: 45–51.
  8. 8. Overstreet RM, Lightner DV, Hasson KW, McIlwain S, Lotz JM (1997) Susceptibility to Taura Syndrome Virus of Some Penaeid Shrimp Species Native to the Gulf of Mexico and the Southeastern United States. J Invertebr Pathol 69: 165–176.
  9. 9. Do JW, Cha SJ, Lee NS, Kim YC, Kim JW, et al. (2006) Taura syndrome virus from Penaeus vannamei shrimp cultured in Korea. Dis Aquat Organ 70: 171–174.
  10. 10. George SK, Kaizer KN, Betz YM, Dhar AK (2011) Multiplication of Taura syndrome virus in primary hemocyte culture of shrimp (Penaeus vannamei). J Virol Methods 172: 54–59.
  11. 11. Tu C, Huang HT, Chuang SH, Hsu JP, Kuo ST, et al. (1999) Taura syndrome in Pacific white shrimp Penaeus vannamei cultured in Taiwan. Dis Aquat Organ 38: 159–161.
  12. 12. Mari J, Poulos BT, Lightner DV, Bonami JR (2002) Shrimp Taura syndrome virus: genomic characterization and similarity with members of the genus Cricket paralysis-like viruses. J Gen Virol 83: 915–926.
  13. 13. Aranguren LF, Salazar M, Tang K, Caraballo X, Lightner D (2012) Characterization of a new strain of Taura syndrome virus (TSV) from Colombian shrimp farms and the implication in the selection of TSV resistant lines. J Invertebr Pathol.
  14. 14. Cote I, Poulos BT, Redman RM, Lightner DV (2009) Development and characterization of a monoclonal antibody against Taura syndrome virus. J Fish Dis 32: 989–996.
  15. 15. James R, Thampuran N, Lalitha KV, Rajan LA, Joseph TC (2010) Differential gene expression profile of the hepatopancreas of white spot syndrome virus infected Fenneropenaeus indicus by suppression subtractive hybridization. Fish Shellfish Immunol 29: 884–889.
  16. 16. Robalino J, Almeida JS, McKillen D, Colglazier J, Trent HF, et al. (2007) Insights into the immune transcriptome of the shrimp Litopenaeus vannamei: tissue-specific expression profiles and transcriptomic responses to immune challenge. Physiol Genomics 29: 44–56.
  17. 17. Leu JH, Chang CC, Wu JL, Hsu CW, Hirono I, et al. (2007) Comparative analysis of differentially expressed genes in normal and white spot syndrome virus infected Penaeus monodon. BMC Genomics 8: 120.
  18. 18. Gomez-Anduro GA, Ascencio-Valle F, Peregrino-Uriarte AB, Campa-Cordova A, Yepiz-Plascencia G (2012) Cytosolic manganese superoxide dismutase genes from the white shrimp Litopenaeus vannamei are differentially expressed in response to lipopolysaccharides, white spot virus and during ontogeny. Comp Biochem Physiol B Biochem Mol Biol 162: 120–125.
  19. 19. Scherbaum S, Ertas B, Gebauer W, Burmester T (2010) Characterization of hemocyanin from the peacock mantis shrimp Odontodactylus scyllarus (Malacostraca: Hoplocarida). J Comp Physiol B 180: 1235–1245.
  20. 20. Song KK, Li DF, Zhang MC, Yang HJ, Ruan LW, et al. (2010) Cloning and characterization of three novel WSSV recognizing lectins from shrimp Marsupenaeus japonicus. Fish Shellfish Immunol 28: 596–603.
  21. 21. Tassanakajon A, Amparyup P, Somboonwiwat K, Supungul P (2011) Cationic antimicrobial peptides in penaeid shrimp. Mar Biotechnol (NY) 13: 639–657.
  22. 22. Huang X, Li Y, Niu Q, Zhang K (2007) Suppression Subtractive Hybridization (SSH) and its modifications in microbiological research. Appl Microbiol Biotechnol 76: 753–760.
  23. 23. Moody DE (2001) Genomics techniques: An overview of methods for the study of gene expression. J Anim Sci 79: E128–E135.
  24. 24. Margulies M, Egholm M, Altman WE, Attiya S, Bader JS, et al. (2005) Genome sequencing in microfabricated high-density picolitre reactors. Nature 437: 376–380.
  25. 25. Liu L, Li Y, Li S, Hu N, He Y, et al. (2012) Comparison of next-generation sequencing systems. J Biomed Biotechnol 2012: 251364.
  26. 26. Guryev V, Cuppen E (2009) Next-generation sequencing approaches in genetic rodent model systems to study functional effects of human genetic variation. FEBS Lett 583: 1668–1673.
  27. 27. Huang X, Madan A (1999) CAP3: A DNA sequence assembly program. Genome Res 9: 868–877.
  28. 28. Mount DW (2007) Using the Basic Local Alignment Search Tool (BLAST). CSH Protoc 2007: pdb top17.
  29. 29. Tatusov RL, Galperin MY, Natale DA, Koonin EV (2000) The COG database: a tool for genome-scale analysis of protein functions and evolution. Nucleic Acids Res 28: 33–36.
  30. 30. Kanehisa M, Goto S (2000) KEGG: kyoto encyclopedia of genes and genomes. Nucleic Acids Res 28: 27–30.
  31. 31. Marioni JC, Mason CE, Mane SM, Stephens M, Gilad Y (2008) RNA-seq: an assessment of technical reproducibility and comparison with gene expression arrays. Genome Res 18: 1509–1517.
  32. 32. Romualdi C, Bortoluzzi S, D'Alessi F, Danieli GA (2003) IDEG6: a web tool for detection of differentially expressed genes in multiple tag sampling experiments. Physiol Genomics 12: 159–162.
  33. 33. Storey J (2002) A direct approach to false discovery rates. Journal of the Royal Statistical Society 64: 19.
  34. 34. Thiel T, Michalek W, Varshney RK, Graner A (2003) Exploiting EST databases for the development and characterization of gene-derived SSR-markers in barley (Hordeum vulgare L.). Theor Appl Genet 106: 411–422.
  35. 35. Rozen S, Skaletsky H (2000) Primer3 on the WWW for general users and for biologist programmers. Methods Mol Biol 132: 365–386.
  36. 36. Rubin BY, Gupta SL (1980) Interferon-induced proteins in human fibroblasts and development of the antiviral state. J Virol 34: 446–454.
  37. 37. Ganz T (2001) Antimicrobial proteins and peptides in host defense. Semin Respir Infect 16: 4–10.
  38. 38. Mogensen TH (2009) Pathogen recognition and inflammatory signaling in innate immune defenses. Clin Microbiol Rev 22: 240–273, Table of Contents.
  39. 39. Robalino J, Almeida JS, McKillen D, Colglazier J, Trent HF, 3rd, et al (2007) Insights into the immune transcriptome of the shrimp Litopenaeus vannamei: tissue-specific expression profiles and transcriptomic responses to immune challenge. Physiol Genomics 29: 44–56.
  40. 40. Litt M, Luty JA (1989) A hypervariable microsatellite revealed by in vitro amplification of a dinucleotide repeat within the cardiac muscle actin gene. Am J Hum Genet 44: 397–401.
  41. 41. Schlotterer C (2000) Evolutionary dynamics of microsatellite DNA. Chromosoma 109: 365–371.
  42. 42. Ozaki A, Khoo SK, Sakamoto T, Okamoto N (2003) Thirty-one polymorphic microsatellite markers for genetic mapping in rainbow trout (Oncorhynchus mykiss). Anim Genet 34: 315–317.
  43. 43. Meehan D, Xu Z, Zuniga G, Alcivar-Warren A (2003) High frequency and large number of polymorphic microsatellites in cultured shrimp, Penaeus (Litopenaeus) vannamei [Crustacea:Decapoda]. Mar Biotechnol (NY) 5: 311–330.
  44. 44. Yao B, Zhao Y, Zhang H, Zhang M, Liu M, et al. (2012) Sequencing and de novo analysis of the Chinese Sika deer antler-tip transcriptome during the ossification stage using Illumina RNA-Seq technology. Biotechnol Lett 34: 813–822.
  45. 45. Gross PS, Bartlett TC, Browdy CL, Chapman RW, Warr GW (2001) Immune gene discovery by expressed sequence tag analysis of hemocytes and hepatopancreas in the Pacific White Shrimp, Litopenaeus vannamei, and the Atlantic White Shrimp, L. setiferus. Dev Comp Immunol 25: 565–577.
  46. 46. Goncalves-Soares D, Seiffert WQ, Schlindwein AD, Toledo-Silva G, Zanette J, et al. (2012) Identification of differentially transcribed genes in shrimp Litopenaeus vannamei exposed to osmotic stress and challenged with WSSV virus. Comp Biochem Physiol Part D Genomics Proteomics 7: 73–81.
  47. 47. Wang B, Li F, Luan W, Xie Y, Zhang C, et al. (2008) Comparison of gene expression profiles of Fenneropenaeus chinensis challenged with WSSV and Vibrio. Mar Biotechnol (NY) 10: 664–675.
  48. 48. Dhar AK, Dettori A, Roux MM, Klimpel KR, Read B (2003) Identification of differentially expressed genes in shrimp (Penaeus stylirostris) infected with White spot syndrome virus by cDNA microarrays. Arch Virol 148: 2381–2396.
  49. 49. Adachi Y, Ishii T, Ikeda Y, Hoshino A, Tamura H, et al. (2004) Characterization of beta-glucan recognition site on C-type lectin, dectin 1. Infect Immun 72: 4159–4171.
  50. 50. Ma TH, Tiu SH, He JG, Chan SM (2007) Molecular cloning of a C-type lectin (LvLT) from the shrimp Litopenaeus vannamei: early gene down-regulation after WSSV infection. Fish Shellfish Immunol 23: 430–437.
  51. 51. Chen DD, Meng XL, Xu JP, Yu JY, Meng MX, et al. (2012) PcLT, a novel C-type lectin from Procambarus clarkii, is involved in the innate defense against Vibrio alginolyticus and WSSV. Dev Comp Immunol.
  52. 52. Wei X, Liu X, Yang J, Fang J, Qiao H, et al. (2012) Two C-type lectins from shrimp Litopenaeus vannamei that might be involved in immune response against bacteria and virus. Fish Shellfish Immunol 32: 132–140.
  53. 53. Zhao ZY, Yin ZX, Xu XP, Weng SP, Rao XY, et al. (2009) A novel C-type lectin from the shrimp Litopenaeus vannamei possesses anti-white spot syndrome virus activity. J Virol 83: 347–356.
  54. 54. Zhang X, Huang C, Qin Q (2004) Antiviral properties of hemocyanin isolated from shrimp Penaeus monodon. Antiviral Res 61: 93–99.
  55. 55. Linzen B (1989) [Blue blood: structure and evolution of hemocyanin]. Naturwissenschaften 76: 206–211.
  56. 56. Pan D, He N, Yang Z, Liu H, Xu X (2005) Differential gene expression profile in hepatopancreas of WSSV-resistant shrimp (Penaeus japonicus) by suppression subtractive hybridization. Dev Comp Immunol 29: 103–112.
  57. 57. Lei K, Li F, Zhang M, Yang H, Luo T, et al. (2008) Difference between hemocyanin subunits from shrimp Penaeus japonicus in anti-WSSV defense. Dev Comp Immunol 32: 808–813.
  58. 58. Havanapan PO, Kanlaya R, Bourchookarn A, Krittanai C, Thongboonkerd V (2009) C-terminal hemocyanin from hemocytes of Penaeus vannamei interacts with ERK1/2 and undergoes serine phosphorylation. J Proteome Res 8: 2476–2483.
  59. 59. Yan F, Xia D, Hu J, Yuan H, Zou T, et al. (2010) Heat shock cognate protein 70 gene is required for prevention of apoptosis induced by WSSV infection. Arch Virol 155: 1077–1083.
  60. 60. Rojtinnakorn J, Hirono I, Itami T, Takahashi Y, Aoki T (2002) Gene expression in haemocytes of kuruma prawn, Penaeus japonicus, in response to infection with WSSV by EST approach. Fish Shellfish Immunol 13: 69–83.
  61. 61. Yang L, Liu J, Liu M, Qian M, Zhang M, et al. (2011) Identification of fatty acid synthase from the Pacific white shrimp, Litopenaeus vannamei and its specific expression profiles during white spot syndrome virus infection. Fish Shellfish Immunol 30: 744–749.
  62. 62. Barakat A, DiLoreto DS, Zhang Y, Smith C, Baier K, et al. (2009) Comparison of the transcriptomes of American chestnut (Castanea dentata) and Chinese chestnut (Castanea mollissima) in response to the chestnut blight infection. BMC Plant Biol 9: 51.
  63. 63. Cohen JJ (1993) Apoptosis: the physiologic pathway of cell death. Hosp Pract (Off Ed) 28: 35–43.
  64. 64. Andrade AA, Silva PN, Pereira AC, De Sousa LP, Ferreira PC, et al. (2004) The vaccinia virus-stimulated mitogen-activated protein kinase (MAPK) pathway is required for virus multiplication. Biochem J 381: 437–446.