Explore
Communities in English
Advertise on Engormix

Teleosts Genomics: Progress and Prospects in Disease Prevention and Control

Published: August 12, 2020
By: Hetron Mweemba Munang’andu 1, Jorge Galindo-Villegas 2 and Lior David 3.
Summary

Genome wide studies based on conventional molecular tools and upcoming omics technologies are beginning to gain functional applications in the control and prevention of diseases in teleosts fish. Herein, we provide insights into current progress and prospects in the use genomics studies for the control and prevention of fish diseases. Metagenomics has emerged to be an important tool used to identify emerging infectious diseases for the timely design of rational disease control strategies, determining microbial compositions in different aquatic environments used for fish farming and the use of host microbiota to monitor the health status of fish. Expounding the use of antimicrobial peptides (AMPs) as therapeutic agents against different pathogens as well as elucidating their role in tissue regeneration is another vital aspect of genomics studies that had taken precedent in recent years. In vaccine development, prospects made include the identification of highly immunogenic proteins for use in recombinant vaccine designs as well as identifying gene signatures that correlate with protective immunity for use as benchmarks in optimizing vaccine efficacy. Progress in quantitative trait loci (QTL) mapping is beginning to yield considerable success in identifying resistant traits against some of the highly infectious diseases that have previously ravaged the aquaculture industry. Altogether, the synopsis put forth shows that genomics studies are beginning to yield positive contribution in the prevention and control of fish diseases in aquaculture.

Keywords: antimicrobial; disease; genomics; immunity; metagenomics; protection; vaccines.

1. Introduction
Teleosts fish species are one of the largest sources of protein for humans through their extensive production in aquaculture. There has been a growing trend in the last decades into the use of genomics to improve our disease control and prevention strategies against various diseases infecting different fish species in aquaculture. As such, genomics studies in aquaculture have evolved from reductionist single gene sequencing to high throughput next (HTS) generation sequencing, which has revolutionized our ability to read genomic data from entire metagenomes and transcriptomes [1,2,3]. These advances have transformed genomic studies into functional applications used for the discovery of genomic traits linked to disease resistance, discovery of novel pathogens, identification of novel probiotics and antimicrobial compounds [4,5,6,7,8,9,10,11]. In vaccinology, genomics studies are being used in search of immune genes that correlate with protective immunity to serve as benchmarks in optimizing vaccine efficacy [12,13,14,15]. In therapeutics, they are used to identify antimicrobial peptides (AMPs) needed to prevent pathogens invasion and remedial genes needed for tissue regeneration [16,17,18,19]. Altogether, these advances show that genomics studies are being transformed into functional applications contributing to reducing the disease burden in fish farming. In the current review, we provide an overview of the progress and prospects of functional genomics based on conventional single gene sequencing, cloning and characterization together with HTS in the prevention and control of fish diseases in aquaculture [20].
2. Metagenomics Analyses
Traditional diagnostic methods used for the identification of infectious agents depend on isolation, culture, characterization, immunoassays and PCR [21]. When these methods fail, enhanced molecular techniques such as nested, multiplex, differentially labelled probes or consensus PCRs using degenerate primers are used [22,23,24,25,26]. In some cases, enhanced molecular techniques also fail to identify novel pathogens resulting in emerging diseases to reach epidemic proportions before the etiological is identified. However, the upcoming of metagenomics, which is a proactive diagnostic approach that does not require prior knowledge of the infectious agent to be diagnosed has accelerated our ability to identify novel pathogens. Currently, there are two approaches used for metagenomics analysis namely (i) deep amplicon sequencing (DAS) that uses pre-sequencing PCR to amplify a selected taxonomic marker such as the 16S rRNA, and (ii) shotgun pyrosequencing in which all nucleic acids present in the sample are sequenced and de novo assembled [21]. Both approaches have been widely used in aquaculture.
 
2.1. Shotgun Pyrosequencing
Shotgun pyrosequencing has contributed to the discovery of more than 40 viral diseases in different aquatic organism in the last decade, which is by far more than the number of viral diseases discovered using the traditional methods in the last 50 years [27]. The duration at which novel viruses are being discovered using this approach is shorter than the period taken by traditional diagnostic methods in which etiological agents are often discovered after the disease has reached epidemic proportions [28]. By enhancing our ability to identify novel pathogens, metagenomics expedites our ability to develop timely disease control strategies. In terms of diagnosis, it has the advantage of being able to identify etiological agents of single and co-infections involving several etiological agents infecting the same fish enabling the selection of the appropriate treatment [29,30,31]. It is also used to identify epidemiological factors linked to the spread of different pathogens through ballast water, global aquaculture trade, sale of fish and fish-by-products and migratory aquatic organisms across the world thereby facilitating the design of effective transboundary disease control strategies and biosafety control measures [27,32,33,34,35]. It has also been used to evaluate the composition of pathogens that resist disinfection in closed recirculation systems indicating that it can be used for optimizing disinfection procedures for recirculation systems [36]. As the cost of shotgun pyrosequencing continues to decrease, we anticipate that metagenomics will become a part of routine diagnosis and epidemiological surveys in aquaculture.
2.2. Deep Amplicon Sequencing
Deep amplicon sequencing (DAS) using a guided taxonomic marker such as 16S rRNA has been used to study the microbial composition of different aquatic environments as an overture to designing to effective preventive measures against endemic fish diseases in each ecosystem. The bacterial composition on the skin surface has been shown to play a major role in protecting fish against pathogen invasion. Boutin et al. [37] used the metagenomics analysis to show changes on the skin surface microbiome. They noted that a shift from Methylobacteria dominated population, which is known to protect the skin surface, to pathogenic species such as Flavobacterium was indicative of susceptibility to infection clearly demonstrating the skin health status can be monitored based on changes in the surface microbial composition. Similarly, Gajardo et al. [38] used metagenomics to identify bacterial species that induce diet dysfunction to serve as biological markers for monitoring the gut health status of Atlantic salmon (Salmo salar L.). 16S rRNA metagenomics can be used to identify probiotics for use in fish as well being as a way of monitoring the gut microbiota changes induced by different feed formulations [39]. The ratio of Firmicutes to Bacteriodetes has been used as a measure of the health index in vertebrates [40,41] implying metagenomics analyses can be used to monitor the health indices of culture fish species based on the composition of gut microbiota. Metagenomics analysis can be used to monitor the safety of aquatic environments used in aquaculture by providing an overview of the prevailing microbial composition, predict the susceptibility of the skin to microbial invasion, determine the health status of the gut and provide the health index of fish based on the gut microbiota composition.
3. Antimicrobial Peptides (AMPs) and Antiviral Compounds
Masso-Silva and Diamon [42] have shown that fish are a great source of antimicrobial peptides (AMPs) that include defensins, cathelicidins, hepcidins, histone derived peptides and piscidins. In general, fish AMPs have broad spectrum antimicrobial properties able to kill both fish and mammalian pathogens and they also have immunodulatory properties [42]. Masso-Silva and Diamon [42] have provide a comprehensive list of fish species in which these AMPs have been detected [42]. However, in the current review, we highlight the use of fish AMPs in disease control and prevention using examples carried out in different studies. Although most AMPs discovered this far have not been commercialized, indications are that they have the potential to play an important role in disease control and prevention in aquaculture.
3.1. Hepcidins
Huang et al. [43] used phage hybridization to identify three hepcidin-like antimicrobial peptides from Oreochromis mossambicus (tilapia) namely Th1-5, TH2-2, and TH2-3. These peptides have antimicrobial properties against bacteria, fungi and viruses [43] and they have been identified, cloned and characterized in different fish species [44,45,46]. They have immunomodulatory functions in fish and they have been shown to produce antitumor effects in human cells [16]. Wang et al. [18] showed high protection against nervous necrosis virus (NNV) using hepcidin1-5 in medaka (Oryzias latipes). They also showed that the level of protection was dose dependent in which fish injected with 1.0 μg, 0.5 μg, 0.01 μg, 0.01 μg/fish had protection levels of RPS = 86, 68, 59 and 48%, respectively.
As for application against bacterial infections, hepcidins produce antimicrobial effects against a wide range of extracellular infections. For example, Liu et al. [44] characterized the functional analysis of the hepcidin gene in roughskin sculpin (Trachidermus fasciatus) and showed that it exhibited a wide spectrum of potent antimicrobial activity against Escherichia coli, Vibrio Anguillarum, Klebsiella pneumoniae, Pseudomonas aeruginosa, Staphylococcus aureus, Bacillus subtilis, B. thuringiensis, and B. megaterium with minimum inhibitory concentrations (MICs) of 5–80 μg/mL (0.825–13.2 μM). In taimen (Hucho taimen, Pallas) it showed antimicrobial activities against Micrococcus lysodeikticus and S. aureusand E. coli [45] while in spotted scat it showed antiviral activities against Siniperca chuatsi rhabdovirus (SCRV) and largemouth bass Micropterus salmoides reovirus (MsReV) as well as a wide range of Gram-negative and positive bacteria species [47].
Apart from showing antimicrobial properties against extracellular bacteria, hepcidin also shows antimicrobial properties against intracellular bacteria as shown by Chen et al. [48] that Hepcidins 1 and 2 had the capacity to hydrolyze purified genomic bacteria DNA for Edwardsiella tarda infection in Mudskipper (Boleophthalmus pectinirostris). Similarly, Alvarez et al. [49] showed that the oxidized hepcidin peptide was more effective than the reduced peptide against Piscirikettsia salmonis, which is an intracellular bacterium shown to cause high mortalities in salmonids. Therefore, these observations show that AMPs could serve as therapeutic agents against intracellular bacterial infections that have posed a significant challenge in vaccine development in aquaculture.
3.2. Epinecidin-1
Epinecidin-1 is an important AMP originally isolated from a complementary cDNA library of grouper (Epinephelus coioides) having a broad range of antimicrobial properties against bacteria, fungi and viruses [17,50]. It has also been characterized from other fish species such as Epinephelus akaara [51]. In human cells, it induces apoptosis that enhances antitumor effects as well as immunomodulatory effects by increasing the expression of different cytokines such as TNFα, IL-10, IL-6 and IFN [52]. Wang et al. [53] examined the efficacy of pre-, co- and post-treatment effects of epinecidin-1 against NNV and showed that co-treatment of epinecidin-1 with NNV decreased grouper mortality. They also observed that challenge with NNV in co-treated fish increased fish survival. In addition, they also showed that fish inoculated with NNV followed with epinecidin-1 treatment showed significantly higher protection suggesting that epinecidin-1 can be used as a drug to rescue infected grouper. Immunohistochemistry staining showed that epi-1 treatment during or after infection cleared off the virus. Apart from increasing survival in fish exposed to viral infections, it has also been shown to increase the survival of fish exposed to bacterial infections such as Vibrio vulnificus in grouper and zebrafish [54].
3.3. Other Antimicrobial Peptides
Other AMPs identified include piscidins, β-defensins and Cathelicidins [42,55,56,57]. Piscidins have antimicrobial activities against a wide range of Gram-positive bacteria inclusive of Streptococci species and Gram-negative bacteria species such as the Vibrio species. In addition, they possess antiviral properties against viruses such as channel catfish virus [58], anti-fungal properties against Psuedosciaena crocea [59] and Candida albican [60] and anti-parasitic properties against Cryptocaryon irritans and Ichthyophthirius multifiliis [59,61]. Similarly, β-defensins have antibacterial properties against A. hydrophila [62], and antiviral properties against a wide range of viruses such as viral hemorrhagic septicemia virus (VHSV), Singapore grouper iridovirus (SGIV) and NNV [63,64,65]. The cathelicidins have been shown to produce antibacterial properties against Yersinia ruckeri [66] and antifungal properties against Candida albicans [67].
3.4. Interferon
Interferons (IFNs) are antiviral compounds known to prevent virus infection whose protective role has been demonstrated in different fish species. Ooi et al. [68] have shown that IFNα-2 is protective against infectious hematopoietic necrosis virus (IHNV) in rainbow trout (Oncorhynchus mykiss) and that protection was dose-dependent with highest relative percent survival (RPS = 90%) being from a dosage of 1.0 g/fish while the 0.1 μg/g dose had a low RPS (39%). Kuo et al. [69] showed that increasing the oral dosage of IFN encapsulated in PLGA nanoparticles from 0.05 to 0.1 μg of PLGA-IFNγ/fish, increased the RPS from 12.4 to 46.7%. Similar results have been reported in stickleback (Gasterosteus aculeatus), medaka and European sea bass (Dicentrarchus labrax) treated with recombinant IFN [70]. However, as pointed out by Ooi et al. [68] the duration of protection induced by IFN treatment is shorter than that produced by humoral responses and, hence, this would account for reasons why IFNs are not widely used as therapeutic agents in aquaculture. Moreover, not all pathogens succumb to IFN treatment as shown by Gadan et al. [71] that infectious pancreatic necrosis virus (IPNV) was able to replicate at a higher rate in the presence of type I IFN in Atlantic salmon (Salmo salar L.) subjected to stress conditions. At cellular level, IFN produces immunomodulation effect as shown by Ooi et al. [68] that IFNa-2 treatment in rainbow trout increased systemic expression of several IFN-induced genes (ISGs) such as IFN1, IRF 1 and 2, MHC-I, STAT1 and vig-1. Similarly, Xu et al. [3,72,73] used transcriptome analyses to show that type I IFNs modulates the expression of several antiviral genes and activate several signaling pathways that lead to the production of several other antiviral genes.
3.5. Therapeutic Compounds
Apart from the discovery of AMPs targeted at protecting fish from invading pathogens, current advances in genomic studies are paving the way into the discovery of remedial peptides useful for tissue regeneration. For example, Hoppe et al. [19] used transcriptome analysis to identify the differentially expressed genes (DEGs) after renal damage caused by gentamicin in killifish (Nothobranchius furzeri). Among the DEGs was miR-21, which was upregulated and was shown to be essential for kidney regeneration in killifish. In fish treated with gentamicin, miR-21 upregulation was blocked by injecting anti-miR-21 prior to induction of renal injury. They noted that apoptosis in renal tubules was caused by miR-21 inhibition while its upregulation positively influenced the initiation of renal tubule regeneration. Hence, it is likely that as studies on transcriptome analyses in search of remedial peptides continues, more genes will be discovered linked to tissue regeneration in fish.
 
4. Application of Genomic Studies in Vaccine Development
In vaccine development, genomics studies have dual functions of identifying protective antigens and elucidating the underlying mechanisms of immune protection induced by vaccination. Several approaches have been used for the search of immunogenic proteins for viral and bacterial vaccines. In the case of viral vaccines, reverse genetics have been used for the identification of immunogenic proteins for IPNV [74], peptide scanning for NNV [75] and IPNV [76], phage display for infectious salmon anemia virus (ISAV) [77] and IHNV [78] while bioinformatics has been used to identify immunogenic proteins for SGIV [79]. The search of bacterial immunogenic proteins is mostly centered on identifying antigens that provide protective immunity across variant serotypes of the same bacterial species. Wang et al. [80] used the tandem mass tag (TMT) labelling-based quantitative proteomics to identify the most immunogenic outer membrane proteins (OMPs) of A. hydrophila. Another approach used for bacterial vaccines is the in vivo induced antigen technology (IVIAT), which has been widely used for the selection of antigens for different bacterial species such as Edwardsiella tarda [81], A. salmonicida [82] and Streptococcus iniae [83]. Taken together, these advances play an important role in the design of genetically engineered vaccines encoding immunogenic proteins in heterologous vectors used in the production of DNA, fusion protein and subunit recombinant vaccines. Apart from genetically engineered vaccines, genomic studies have also been used for the identification of avirulent strains for use as live attenuated vaccines as shown by Zhang et al. [84] who used Illumina sequencing for the characterization of attenuated strains of Streptococcus agalactiae with the potential to serve as a live vaccine.
In fish vaccinology, transcriptome studies are being used to elucidate the mechanisms underlying vaccine protection. For example, Zhang et al. [85] used transcriptome analyses to show that early immune responses in Large yellow croaker (Larimichthys crocea) vaccinated against Vibrio parahaemolyticus and A. hydrophila were dominated by innate defense molecules involved in antigen uptake and inflammasome formation. RNA-seq data generated by Yan et al. [86] showed an activated MHC-I and inhibited MCH-II pathway in zebrafish immunized with a live attenuated E. tarda vaccine during the early stages of immune response to vaccination. Yasuike et al. [87] identified five interferon-inducible genes (ISG) inclusive of Mx that correlated with high protection induced by the G-protein encoded in the DNA vaccine used to immunize Japanese flounder against Hirame rhabdovirus (HIRRV). Similarly, Liu et al. [88] used deep sequencing to identify a repertoire of seven genes comprising of IFN-I, TNFα, CRP, IL-8, IgM, MHC-I and CD8α that correlated with high protection in grass carp (Ctenopharyngodon idella) vaccinated against A. hydrophila using a live recombinant vaccine while Bridle et al. [15] identified a biosignature dominated by IgH and selenoprotein predictive of the level of protection in Atlantic salmon vaccinated against Yersinia ruckeri by immersion. Other genes that have been shown to correlate with vaccine protection include GATA-3 expressed in Atlantic salmon vaccinated against IPNV [12] and Mx in rainbow trout vaccinated against viral hemorrhagic septicemia virus (VHSV) [89]. Even though antibodies are the most widely used measure of immune response to vaccination in fish [13,90,91,92] they cannot be used as correlates of protection for cytotoxic T-lymphocyte (CTL) responses [93]. Hence, the use of transcriptome analyses does not only provide a global understanding of the host immune response to vaccination, but it unravels gene markers that correlate with cellular-mediated immune responses for use as benchmarks in vaccine development.
5. Application of Functional Genomics in Disease Control
5.1. Genetic Markers of Disease Resistance and Susceptibility
Different approaches have been used for the identification of disease resistance and susceptibility traits in fish. These include the random fragment length polymorphisms (RFLP), random amplified polymorphism DNA (RAPD), amplified fragment length polymorphism (AFLP) and quantitative trait loci (QTL) [94]. In Norway, identification of a QTL for resistance against IPNV in Atlantic salmon has led to the significant decrease in the number of IPNV cases reported annually from more than 200 cases in 2005 to less than 30 cases in 2016 [95] clearly demonstrating that QTL analysis can reliably be used to identify disease resistance traits for diseases having a devastating economic impact in aquaculture. Table 1 shows this approach is being explored for the search of markers against other several diseases in aquaculture.
Table 1. Traits for disease resistance determined by quantitative trail loci mapping.
Teleosts Genomics: Progress and Prospects in Disease Prevention and Control - Image 1
Apart from disease resistance traits, genetic markers for susceptibility have also made positive contributions to disease control in aquaculture. Fish that are highly susceptible to disease infection have found a positive contribution in vaccine production as one of the most important elements required for optimizing challenge models needed in vaccine efficacy trials. As pointed out previously [13,14,115], for a challenge model to be reproducible and reliable, it must have a wide discriminatory capacity between the vaccinated and control fish. Given that relative percent survival (RPS) is the most widely used measure of vaccine efficacy, fish used in challenge models must show high mortality after lethal challenge. Hence, genetic selection for susceptibility traits serves as an important tool for identifying fish strains required for developing optimal challenge models needed for vaccine efficacy trials. Recently, we used a highly susceptible strain of Atlantic salmon in optimization the cohabitation challenge model for IPNV in which we obtained high mortality in unvaccinated fish (>80%) in three independent studies leading to a wide discriminatory capacity (RPS > 50%) between the vaccinated and control fish [115]. These studies demonstrate that genomic studies are being used in vaccine production and disease control strategies in aquaculture while fish that are highly susceptible to disease infection are used in optimization of vaccine performance.
5.2. Immunological Markers of Disease Resistance
Different immune parameters such as the major histocompatibility (MHC) molecules can be used for the selection of disease susceptibility and resistance traits in fish. MHC variants influence several vital biological traits such as susceptibility to infection, autoimmunity and immune recognition governed by their ability to encode cell surface glycoproteins that bind peptides processed for presentation to T-lymphocytes [116]. Different studies have shown that the loci that code for peptide binding on MHC-molecules are among the most polymorphic loci in vertebrates [116]. As shown in Table 2, the association between MHC polymorphism and disease resistance has been reported in different fish species. Xu et al. [117] found a single allele linked to resistant to Vibrio anguillarum infection in 12 selective Japanese flounder families. In another study, Yang et al. [118] identified a single nucleotide polymorphism (SNP) associated with resistance against Singapore grouper iridovirus (SGIV) in Orange spotted grouper (Epinephelus coioides). To confirm this, a fish encoding the highly resistant (HR) and less resistant (LR) SNP were infected with SGIV resulting in low mortality in the HR compared to LR fish. Similarly, Palti et al. [119] found an association between SNPs tightly linked to MHC-II and IHNV resistance in backcrosses of rainbow trout and cutthroat trout (Oncorhynchus clarkii).
Table 2. Immune markers of disease resistance.
Teleosts Genomics: Progress and Prospects in Disease Prevention and Control - Image 2
Apart from MHC molecules, SNPs have been reported in other immune genes such as the toll like receptors (TLRs) and retinoic acid inducible gene I (RIG-) like receptors (RLRs), which are germline coded pathogen recognition receptors (PRRs) involved in the recognition of pattern associated molecular patterns (PAMPs) expressed by different pathogens. As shown in Table 2, SNPs on PRRs have been associated with disease resistance and susceptibility traits. For example, Heng et al. [128] selected eight SNPs in the non-coding region of TLR3 and analyzed the genotype allele distribution of susceptibility and resistance traits of carp to grass carp reovirus (GCRV) infection using PCR-RFLP. They found one allele, which was significantly associated with resistance to GCRV. To confirm the correlation, they carried out a challenge study, which showed significantly low mortality in fish encoding the disease resistance SNP. Similarly, Wang et al. [130] identified six SNPs on the MDA5 genome of grass carp using PCR-RFLP and examined their association with resistance to GCRV. They found two loci linked to resistance against GCRV and carried out verification challenge studies in which the HR group had significantly low mortality unlike the LR group that had high mortality. Other PRRs shown to have SNPs linked to disease resistance against GCRV are TLR22, RIG-I like and LGP2 [131,132,136]. Apart from PRRs, Kongchum et al. [126] found an association between IL-10a SNP and resistance to Cyprinid herpesvirus infection in common carp (Cyprinus carpio).
6. Conclusions
In general, genomics studies in aquaculture have progressively been transformed into functional applications aimed at improving the quality of fish products by contributing to reducing the disease burden in aquaculture. Metagenomics analyses focus on identifying novel pathogens, epidemiological monitoring of the microbiome composition of aquatic environments used for aquaculture and establishing biological markers for monitoring skin and gut health status based on microbiota composition. Discovery of a wide range of AMPs as therapeutic drugs has been attained for different fish species with prospective applications that would contribute to reduction of antibiotic use in aquaculture. Progress in vaccine development has not only led to discovery of various immunogenic proteins essential for optimizing vaccine performance, but is slowly being turned into vaccine efficacy optimization using genomics by trying to identify gene signatures that correlate with protective immunity for use as measures of efficacy in vaccine production. Finally, single nucleotide polymorphisms and QTL mapping are showing great prospects in disease prevention and could lead to disease eradication in some parts of the world as shown in the use of QTL resistance traits for IPNV in Atlantic salmon, which has significantly reduced the number of cases of this devastating disease in Norway. In summary, these observations show that current advances in genomics studies are positively contributing to prevention and control of fish diseases in aquaculture.
Acknowledgments
This study was funded by the Targeted disease prophylaxis in European Fish farming project, European Union grant agreement 311993 and the Nanoparticle Encapsulation of Plant based vaccines against Piscine Reovirus, Norwegian Research Council (NRC) Project number 239140.
Author Contributions
Hetron Mweemba Munang’andu, Jorge Galindo-Villegas and Lior David all conceived the idea, participated in literature search and preparation of the manuscript. All authors approved publication of the manuscript.
Conflicts of Interest
The authors declare no conflict of interest.
This article was originally published in International Journal of Molecular Sciences. 2018, 19, 1083; doi:10.3390/ijms19041083. This is an Open Access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).

1. Valenzuela-Miranda, D.; Boltana, S.; Cabrejos, M.E.; Yanez, J.M.; Gallardo-Escarate, C. High-throughput transcriptome analysis of ISAV-infected Atlantic salmon Salmo salar unravels divergent immune responses associated to head-kidney, liver and gills tissues. Fish Shellfish Immunol. 201545, 367–377.

2. Wynne, J.W.; O’Sullivan, M.G.; Cook, M.T.; Stone, G.; Nowak, B.F.; Lovell, D.R.; Elliott, N.G. Transcriptome analyses of amoebic gill disease-affected Atlantic salmon (Salmo salar) tissues reveal localized host gene suppression. Mar. Biotechnol. 200810, 388–403.

3. Xu, C.; Evensen, O.; Munang’andu, H.M. De novo assembly and transcriptome analysis of Atlantic salmon macrophage/dendritic-like TO cells following type I IFN treatment and Salmonid alphavirus subtype-3 infection. BMC Genom. 201516, 96.

4. Tsai, H.; Hamilton, A.; Guy, D.; Houston, R. Single nucleotide polymorphisms in the insulin-like growth factor 1 (IGF1) gene are associated with growth-related traits in farmed Atlantic salmon. Anim. Genet. 201445, 709–715.

5. Baranski, M.; Moen, T.; Våge, D.I. Mapping of quantitative trait loci for flesh colour and growth traits in Atlantic salmon (Salmo salar). Genet. Sel. Evolut. 201042, 17.

6. Paibomesai, M.I.; Moghadam, H.K.; Ferguson, M.M.; Danzmann, R.G. Clock genes and their genomic distributions in three species of salmonid fishes, Associations with genes regulating sexual maturation and cell cycling. BMC Res. Notes 20103, 215.

7. Moen, T.; Baranski, M.; Sonesson, A.K.; Kjoglum, S. Confirmation and fine-mapping of a major QTL for resistance to infectious pancreatic necrosis in Atlantic salmon (Salmo salar): Population-level associations between markers and trait. BMC Genom. 200910, 368.

8. Moen, T.; Sonesson, A.K.; Hayes, B.; Lien, S.; Munck, H.; Meuwissen, T.H. Mapping of a quantitative trait locus for resistance against infectious salmon anaemia in Atlantic salmon (Salmo salar): Comparing survival analysis with analysis on affected/resistant data. BMC Genet. 20078, 53.

9. Vallejo, R.L.; Palti, Y.; Liu, S.; Evenhuis, J.P.; Gao, G.; Rexroad, C.E., 3rd; Wiens, G.D. Detection of QTL in rainbow trout affecting survival when challenged with Flavobacterium psychrophilum. Mar. Biotechnol. 201416, 349–360.

10. Verrier, E.R.; Dorson, M.; Mauger, S.; Torhy, C.; Ciobotaru, C.; Hervet, C.; Dechamp, N.; Genet, C.; Boudinot, P.; Quillet, E. Resistance to a rhabdovirus (VHSV) in rainbow trout: Identification of a major QTL related to innate mechanisms. PLoS ONE 20138, e55302.

11. Tsai, H.Y.; Hamilton, A.; Guy, D.R.; Tinch, A.E.; Bishop, S.C.; Houston, R.D. The genetic architecture of growth and fillet traits in farmed Atlantic salmon (Salmo salar). BMC Genet. 201516, 1.

12. Munang’andu, H.M.; Fredriksen, B.N.; Mutoloki, S.; Dalmo, R.A.; Evensen, O. The kinetics of CD4+ and CD8+ T-cell gene expression correlate with protection in Atlantic salmon (Salmo salar L.) vaccinated against infectious pancreatic necrosis. Vaccine 201331, 1956–1963.

13. Munang’andu, H.M.; Fredriksen, B.N.; Mutoloki, S.; Dalmo, R.A.; Evensen, O. Antigen dose and humoral immune response correspond with protection for inactivated infectious pancreatic necrosis virus vaccines in Atlantic salmon (Salmo salar L.). Vet. Res. 201344, 7.

14. Munang’andu, H.M.; Mutoloki, S.; Evensen, O. Acquired immunity and vaccination against infectious pancreatic necrosis virus of salmon. Dev. Comp. Immunol. 201443, 184–196.

15. Bridle, A.R.; Koop, B.F.; Nowak, B.F. Identification of surrogates of protection against yersiniosis in immersion vaccinated Atlantic salmon. PLoS ONE 20127, e40841.

16. Chen, J.-Y.; Lin, W.-J.; Lin, T.-L. A fish antimicrobial peptide, tilapia hepcidin TH2-3, shows potent antitumor activity against human fibrosarcoma cells. Peptides 200930, 1636–1642.

17. Pan, C.-Y.; Chen, J.-Y.; Cheng, Y.-S.E.; Chen, C.-Y.; Ni, I.-H.; Sheen, J.-F.; Pan, Y.-L.; Kuo, C.-M. Gene expression and localization of the epinecidin-1 antimicrobial peptide in the grouper (Epinephelus coioides), and its role in protecting fish against pathogenic infection. DNA Cell Biol. 200726, 403–413.

18. Wang, Y.-D.; Kung, C.-W.; Chen, J.-Y. Antiviral activity by fish antimicrobial peptides of epinecidin-1 and hepcidin 1–5 against nervous necrosis virus in medaka. Peptides 201031, 1026–1033.

19. Hoppe, B.; Pietsch, S.; Franke, M.; Engel, S.; Groth, M.; Platzer, M.; Englert, C. MiR-21 is required for efficient kidney regeneration in fish. BMC Dev. Biol. 201515, 43.

20. Tan, B.; Ng, C.; Nshimyimana, J.P.; Loh, L.L.; Gin, K.Y.-H.; Thompson, J.R. Next-generation sequencing (NGS) for assessment of microbial water quality: Current progress, challenges, and future opportunities. Front. Microbiol.20156, 1027.

21. Miller, R.R.; Montoya, V.; Gardy, J.L.; Patrick, D.M.; Tang, P. Metagenomics for pathogen detection in public health. Genome Med. 20135, 81.

22. Mahony, J.B.; Hatchette, T.; Ojkic, D.; Drews, S.J.; Gubbay, J.; Low, D.E.; Petric, M.; Tang, P.; Chong, S.; Luinstra, K. Multiplex PCR tests sentinel the appearance of pandemic influenza viruses including H1N1 swine influenza. J. Clin. Virol. 200945, 200–202.

23. Rohayem, J.; Berger, S.; Juretzek, T.; Herchenröder, O.; Mogel, M.; Poppe, M.; Henker, J.; Rethwilm, A. A simple and rapid single-step multiplex RT-PCR to detect Norovirus, Astrovirus and Adenovirus in clinical stool samples. J. Virol. Methods 2004118, 49–59.

24. Molenkamp, R.; van der Ham, A.; Schinkel, J.; Beld, M. Simultaneous detection of five different DNA targets by real-time Taqman PCR using the Roche LightCycler480: Application in viral molecular diagnostics. J. Virol. Methods 2007141, 205–211.

25. Roh, K.H.; Kim, J.; Nam, M.-H.; Yoon, S.; Lee, C.K.; Lee, K.; Yoo, Y.; Kim, M.J.; Cho, Y. Comparison of the Seeplex reverse transcription PCR assay with the R-mix viral culture and immunofluorescence techniques for detection of eight respiratory viruses. Ann. Clin. Lab. Sci. 200838, 41–46. 

26. Yamaguchi, S.; Kaji, N.; Munang’andu, H.; Kojima, C.; Mase, M.; Tsukamoto, K. Quantification of chicken anaemia virus by competitive polymerase chain reaction. Avian Pathol. 200029, 305–310.

27. Munang’andu, H.M.; Mugimba, K.K.; Byarugaba, D.K.; Mutoloki, S.; Evensen, Ø. Current advances on virus discovery and diagnostic role of viral metagenomics in aquatic organisms. Front. Microbiol. 20178, 406.

28. Munang’andu, H.M. Environmental viral metagenomics analyses in aquaculture: Applications in epidemiology and disease control. Front. Microbiol. 20167, 1986.

29. Yang, J.; Yang, F.; Ren, L.; Xiong, Z.; Wu, Z.; Dong, J.; Sun, L.; Zhang, T.; Hu, Y.; Du, J. Unbiased parallel detection of viral pathogens in clinical samples by use of a metagenomic approach. J. Clin. Microbiol. 201149, 3463–3469.

30. Ng, T.F.F.; Manire, C.; Borrowman, K.; Langer, T.; Ehrhart, L.; Breitbart, M. Discovery of a novel single-stranded DNA virus from a sea turtle fibropapilloma by using viral metagenomics. J. Virol. 200983, 2500–2509.

31. Hewson, I.; Button, J.B.; Gudenkauf, B.M.; Miner, B.; Newton, A.L.; Gaydos, J.K.; Wynne, J.; Groves, C.L.; Hendler, G.; Murray, M. Densovirus associated with sea-star wasting disease and mass mortality. Proc. Natl. Acad. Sci. USA 2014111, 17278–17283.

32. Haenen, O.; Way, K.; Bergmann, S.; Ariel, E. The emergence of koi herpesvirus and its significance to European aquaculture. Bull. Eur. Assoc. Fish Pathol. 200424, 293–307.

33. Lightner, D.; Redman, R.; Poulos, B.; Nunan, L.; Mari, J.; Hasson, K. Risk of spread of penaeid shrimp viruses in the Americas by the international movement of live and frozen shrimp. Rev. Sci. Tech. 199716, 146–160.

34. Walker, P.J.; Winton, J.R. Emerging viral diseases of fish and shrimp. Vet. Res. 201041, 51.

35. Kim, Y.; Aw, T.G.; Rose, J.B. Transporting ocean viromes: Invasion of the aquatic biosphere. PLoS ONE 201611, e0152671.

36. Rosario, K.; Nilsson, C.; Lim, Y.W.; Ruan, Y.; Breitbart, M. Metagenomic analysis of viruses in reclaimed water. Environ. Microbiol. 200911, 2806–2820.

37. Boutin, S.; Sauvage, C.; Bernatchez, L.; Audet, C.; Derome, N. Inter individual variations of the fish skin microbiota: Host genetics basis of mutualism? PLoS ONE 20149, e102649.

38. Gajardo, K.; Jaramillo-Torres, A.; Kortner, T.M.; Merrifield, D.L.; Tinsley, J.; Bakke, A.M.; Krogdahl, Å. Alternative protein sources in the diet modulate microbiota and functionality in the distal intestine of Atlantic salmon (Salmo salar). Appl. Environ. Microbiol. 201783, e02615–e02616.

39. Kim, D.-H.; Austin, B. Innate immune responses in rainbow trout (Oncorhynchus mykiss, Walbaum) induced by probiotics. Fish Shellfish Immunol. 200621, 513–524.

40. Nicholson, J.K.; Holmes, E.; Kinross, J.; Burcelin, R.; Gibson, G.; Jia, W.; Pettersson, S. Host-gut microbiota metabolic interactions. Science 2012336, 1262–1267.

41. Kelder, T.; Stroeve, J.H.; Bijlsma, S.; Radonjic, M.; Roeselers, G. Correlation network analysis reveals relationships between diet-induced changes in human gut microbiota and metabolic health. Nutr. Diabetes 20144, e122.

42. Masso-Silva, J.A.; Diamond, G. Antimicrobial peptides from fish. Pharmaceuticals 20147, 265–310.

43. Huang, P.-H.; Chen, J.-Y.; Kuo, C.-M. Three different hepcidins from tilapia, Oreochromis mossambicus: Analysis of their expressions and biological functions. Mol. Immunol. 200744, 1922–1934.

44. Liu, Y.; Han, X.; Chen, X.; Yu, S.; Chai, Y.; Zhai, T.; Zhu, Q. Molecular characterization and functional analysis of the hepcidin gene from roughskin sculpin (Trachidermus fasciatus). Fish Shellfish Immunol. 201768, 349–358.

45. Wang, D.; Li, S.; Zhao, J.; Liu, H.; Lu, T.; Yin, J. Genomic organization, expression and antimicrobial activity of a hepcidin from Taimen (Hucho taimen, Pallas). Fish Shellfish Immunol. 201656, 303–309.

46. Li, Z.; Hong, W.S.; Qiu, H.T.; Zhang, Y.T.; Yang, M.S.; You, X.X.; Chen, S.X. Cloning and expression of two hepcidin genes in the mudskipper (Boleophthalmus pectinirostris) provides insights into their roles in male reproductive immunity. Fish Shellfish Immunol. 201656, 239–247.

47. Gui, L.; Zhang, P.; Zhang, Q.; Zhang, J. Two hepcidins from spotted scat (Scatophagus argus) possess antibacterial and antiviral functions in vitro. Fish Shellfish Immunol. 201650, 191–199.

48. Chen, J.; Nie, L.; Chen, J. Mudskipper (Boleophthalmus pectinirostris) Hepcidin-1 and Hepcidin-2 Present Different Gene Expression Profile and Antibacterial Activity and Possess Distinct Protective Effect against Edwardsiella tarda Infection. Probiotics Antimicrob. Proteins 2017, 1–10.

49. Álvarez, C.A.; Guzmán, F.; Cárdenas, C.; Marshall, S.H.; Mercado, L. Antimicrobial activity of trout hepcidin. Fish Shellfish Immunol. 201441, 93–101.

50. Pan, C.-Y.; Chen, J.-Y.; Lin, T.-L.; Lin, C.-H. In vitro activities of three synthetic peptides derived from epinecidin-1 and an anti-lipopolysaccharide factor against Propionibacterium acnesCandida albicans, and Trichomonas vaginalisPeptides 200930, 1058–1068.

51. Mao, M.-G.; Jiang, J.-L.; Perálvarez-Marín, A.; Wang, K.-J.; Lei, J.-L. Characterization of the Mx and hepcidin genes in Epinephelus akaara asymptomatic carriers of the nervous necrosis virus. Aquaculture 2013408, 175–183.

52. Chen, J.-Y.; Lin, W.-J.; Wu, J.-L.; Her, G.M.; Hui, C.-F. Epinecidin-1 peptide induces apoptosis which enhances antitumor effects in human leukemia U937 cells. Peptides 200930, 2365–2373.

53. Wang, Y.-D.; Kung, C.-W.; Chi, S.-C.; Chen, J.-Y. Inactivation of nervous necrosis virus infecting grouper (Epinephelus coioides) by epinecidin-1 and hepcidin 1–5 antimicrobial peptides, and downregulation of Mx2 and Mx3 gene expressions. Fish Shellfish Immunol. 201028, 113–120.

54. Pan, C.-Y.; Huang, T.-C.; Wang, Y.-D.; Yeh, Y.-C.; Hui, C.-F.; Chen, J.-Y. Oral administration of recombinant epinecidin-1 protected grouper (Epinephelus coioides) and zebrafish (Danio rerio) from Vibrio vulnificus infection and enhanced immune-related gene expressions. Fish Shellfish Immunol. 201232, 947–957.

55. Desriac, F.; Jégou, C.; Brillet, B.; Le Chevalier, P.; Fleury, Y. Antimicrobial Peptides from Fish. In Utilization of Fish Waste; CRC Press: Boca Raton, FL, USA, 2013.

56. Rajanbabu, V.; Chen, J.-Y. Applications of antimicrobial peptides from fish and perspectives for the future. Peptides201132, 415–420.

57. Najafian, L.; Babji, A.S. A review of fish-derived antioxidant and antimicrobial peptides: Their production, assessment, and applications. Peptides 201233, 178–185.

58. Chinchar, V.; Bryan, L.; Silphadaung, U.; Noga, E.; Wade, D.; Rollins-Smith, L. Inactivation of viruses infecting ectothermic animals by amphibian and piscine antimicrobial peptides. Virology 2004323, 268–275.

59. Niu, S.-F.; Jin, Y.; Xu, X.; Qiao, Y.; Wu, Y.; Mao, Y.; Su, Y.-Q.; Wang, J. Characterization of a novel piscidin-like antimicrobial peptide from Pseudosciaena crocea and its immune response to Cryptocaryon irritans. Fish Shellfish Immunol. 201335, 513–524.

60. Cho, J.; Lee, D.G. Oxidative stress by antimicrobial peptide pleurocidin triggers apoptosis in Candida albicans. Biochimie 201193, 1873–1879.

61. Colorni, A.; Ullal, A.; Heinisch, G.; Noga, E. Activity of the antimicrobial polypeptide piscidin 2 against fish ectoparasites. J. Fish Dis. 200831, 423–432.

62. Zhao, J.-G.; Zhou, L.; Jin, J.-Y.; Zhao, Z.; Lan, J.; Zhang, Y.-B.; Zhang, Q.-Y.; Gui, J.-F. Antimicrobial activity-specific to Gram-negative bacteria and immune modulation-mediated NF-κB and Sp1 of a medaka β-defensin. Dev. Comp. Immunol. 200933, 624–637.

63. Falco, A.; Chico, V.; Marroqui, L.; Perez, L.; Coll, J.; Estepa, A. Expression and antiviral activity of a β-defensin-like peptide identified in the rainbow trout (Oncorhynchus mykiss) EST sequences. Mol. Immunol. 200845, 757–765.

64. Guo, M.; Wei, J.; Huang, X.; Huang, Y.; Qin, Q. Antiviral effects of β-defensin derived from orange-spotted grouper (Epinephelus coioides). Fish Shellfish Immunol. 201232, 828–838.

65. Jin, J.-Y.; Zhou, L.; Wang, Y.; Li, Z.; Zhao, J.-G.; Zhang, Q.-Y.; Gui, J.-F. Antibacterial and antiviral roles of a fish β-defensin expressed both in pituitary and testis. PLoS ONE 20105, e12883.

66. Bridle, A.; Nosworthy, E.; Polinski, M.; Nowak, B. Evidence of an antimicrobial-immunomodulatory role of Atlantic salmon cathelicidins during infection with Yersinia ruckeri. PLoS ONE 20116, e23417.

67. Broekman, D.C.; Zenz, A.; Gudmundsdottir, B.K.; Lohner, K.; Maier, V.H.; Gudmundsson, G.H. Functional characterization of codCath, the mature cathelicidin antimicrobial peptide from Atlantic cod (Gadus morhua). Peptides 201132, 2044–2051.

68. Ooi, E.L.; Verjan, N.; Haraguchi, I.; Oshima, T.; Kondo, H.; Hirono, I.; Aoki, T.; Kiyono, H.; Yuki, Y. Innate immunomodulation with recombinant interferon-α enhances resistance of rainbow trout (Oncorhynchus mykiss) to infectious hematopoietic necrosis virus. Dev. Comp. Immunol. 200832, 1211–1220.

69. Kuo, H.P.; Goh, L.L.; Lu, M.W.; Kong, Z.L. Delivery of Grouper Interferon by Chitosan-Modified Poly (lactic-co-glycolic acid) Nanoparticles to Protect Nervous Necrosis Virus Infection. J. Nanosci. Nanotechnol. 201616, 7521–7529.

70. Ohtani, M.; Hikima, J.; Kondo, H.; Hirono, I.; Jung, T.S.; Aoki, T. Characterization and antiviral function of a cytosolic sensor gene, MDA5, in Japanese flounder, Paralichthys olivaceusDev. Comp. Immunol. 201135, 554–562.

71. Gadan, K.; Sandtrø, A.; Marjara, I.S.; Santi, N.; Munang’andu, H.M.; Evensen, Ø. Stress-induced reversion to virulence of infectious pancreatic necrosis virus in naive fry of Atlantic salmon (Salmo salar L.). PLoS ONE 20138, e54656.

72. Xu, C.; Evensen, O.; Munang’andu, H.M. De Novo Transcriptome Analysis Shows That SAV-3 Infection Upregulates Pattern Recognition Receptors of the Endosomal Toll-Like and RIG-I-Like Receptor Signaling Pathways in Macrophage/Dendritic Like TO-Cells. Viruses 20168, 114.

73. Xu, C.; Evensen, O.; Munang’andu, H.M. A de novo transcriptome analysis shows that modulation of the JAK-STAT signaling pathway by salmonid alphavirus subtype 3 favors virus replication in macrophage/dendritic-like TO-cells. BMC Genom. 201617, 390.

74. Munang’andu, H.M.; Sandtrø, A.; Mutoloki, S.; Brudeseth, B.E.; Santi, N.; Evensen, Ø. Immunogenicity and cross protective ability of the central VP2 amino acids of infectious pancreatic necrosis virus in Atlantic salmon (Salmo salar L.). PLoS ONE 20138, e54263.

75. Coeurdacier, J.-L.; Laporte, F.; Pepin, J.-F. Preliminary approach to find synthetic peptides from nodavirus capsid potentially protective against sea bass viral encephalopathy and retinopathy. Fish Shellfish Immunol. 200314, 435–447.

76. Jofré, C.; Guzmán, F.; Cárdenas, C.; Albericio, F.; Marshall, S.H. A natural peptide and its variants derived from the processing of infectious pancreatic necrosis virus (IPNV) displaying enhanced antimicrobial activity: A novel alternative for the control of bacterial diseases. Peptides 201132, 852–858.

77. Ojeda, N.; Cárdenas, C.; Guzmán, F.; Marshall, S.H. Chemical Synthesis and In Vitro Evaluation of a Phage Display-Derived Peptide Active against Infectious Salmon Anemia Virus. Appl. Environ. Microbiol. 201682, 2563–2571.

78. Xu, L.-M.; Liu, M.; Zhao, J.-Z.; Yin, J.-S.; Liu, H.-B.; Lu, T. Epitope mapping of the infectious hematopoietic necrosis virus glycoprotein by flow cytometry. Biotechnol. Lett. 201436, 2109–2116.

79. Ou-yang, Z.; Wang, P.; Huang, Y.; Huang, X.; Wan, Q.; Zhou, S.; Wei, J.; Zhou, Y.; Qin, Q. Selection and identification of Singapore grouper iridovirus vaccine candidate antigens using bioinformatics and DNA vaccination. Vet. Immunol. Immunopathol. 2012149, 38–45.

80. Wang, Y.; Chen, H.; Guo, Z.; Sun, L.; Fu, Y.; Li, T.; Lin, W.; Lin, X. Quantitative proteomic analysis of iron-regulated outer membrane proteins in Aeromonas hydrophila as potential vaccine candidates. Fish Shellfish Immunol. 201768, 1–9.

81. Jiao, X.-D.; Dang, W.; Hu, Y.-H.; Sun, L. Identification and immunoprotective analysis of an in vivo-induced Edwardsiella tarda antigen. Fish Shellfish Immunol. 200927, 633–638.

82. Menanteau-Ledouble, S.; Soliman, H.; Kumar, G.; El-Matbouli, M. Use of in vivo induced antigen technology to identify genes from Aeromonas salmonicida subsp. salmonicida that are specifically expressed during infection of the rainbow trout Oncorhynchus mykissBMC Vet. Res. 201410, 298.

83. Sun, Y.; Hu, Y.-H.; Liu, C.-S.; Sun, L. Construction and analysis of an experimental Streptococcus iniae DNA vaccine. Vaccine 201028, 3905–3912.

84. Zhang, Z.; Yu, A.; Lan, J.; Zhang, Y.; Zhang, H.; Li, Y.; Hu, M.; Cheng, J.; Wei, S.; Lin, L. Draft Genome Sequence of an Attenuated Streptococcus agalactiae Strain Isolated from the Gut of a Nile Tilapia (Oreochromis niloticus). Genome Announc. 20175, e01627-16.

85. Zhang, X.; Mu, Y.; Mu, P.; Ao, J.; Chen, X. Transcriptome Analysis Reveals Comprehensive Insights into the Early Immune Response of Large Yellow Croaker (Larimichthys crocea) Induced by Trivalent Bacterial Vaccine. PLoS ONE 201712, e0170958.

86. Yang, D.; Liu, Q.; Yang, M.; Wu, H.; Wang, Q.; Xiao, J.; Zhang, Y. RNA-seq liver transcriptome analysis reveals an activated MHC-I pathway and an inhibited MHC-II pathway at the early stage of vaccine immunization in zebrafish. BMC Genom. 201213, 319.

87. Yasuike, M.; Kondo, H.; Hirono, I.; Aoki, T. Difference in Japanese flounder, Paralichthys olivaceus gene expression profile following hirame rhabdovirus (HIRRV) G and N protein DNA vaccination. Fish Shellfish Immunol.200723, 531–541.

88. Liu, L.; Gong, Y.-X.; Zhu, B.; Liu, G.-L.; Wang, G.-X.; Ling, F. Effect of a new recombinant Aeromonas hydrophila vaccine on the grass carp intestinal microbiota and correlations with immunological responses. Fish Shellfish Immunol. 201545, 175–183.

89. McLauchlan, P.; Collet, B.; Ingerslev, E.; Secombes, C.; Lorenzen, N.; Ellis, A. DNA vaccination against viral haemorrhagic septicaemia (VHS) in rainbow trout, size, dose, route of injection and duration of protection—early protection correlates with Mx expression. Fish Shellfish Immunol. 200315, 39–50.

90. Dubey, S.; Avadhani, K.; Mutalik, S.; Sivadasan, S.M.; Maiti, B.; Girisha, S.K.; Venugopal, M.N.; Mutoloki, S.; Evensen, O.; Karunasagar, I.; et al. Edwardsiella tarda OmpA Encapsulated in Chitosan Nanoparticles Shows Superior Protection over Inactivated Whole Cell Vaccine in Orally Vaccinated Fringed-Lipped Peninsula Carp (Labeo fimbriatus). Vaccines 20164, 40.

91. Dubey, S.; Avadhani, K.; Mutalik, S.; Sivadasan, S.M.; Maiti, B.; Paul, J.; Girisha, S.K.; Venugopal, M.N.; Mutoloki, S.; Evensen, O.; et al. Aeromonas hydrophila OmpW PLGA Nanoparticle Oral Vaccine Shows a Dose-Dependent Protective Immunity in Rohu (Labeo rohita). Vaccines 20164.

92. Munang’andu, H.M.; Paul, J.; Evensen, O. An Overview of Vaccination Strategies and Antigen Delivery Systems for Streptococcus agalactiae Vaccines in Nile Tilapia (Oreochromis niloticus). Vaccines 20164, 48.

93. Munang’andu, H.M.; Evensen, O. A Review of Intra- and Extracellular Antigen Delivery Systems for Virus Vaccines of Finfish. J. Immunol. Res. 20152015, 960859.

94. Yáñez, J.M.; Houston, R.D.; Newman, S. Genetics and genomics of disease resistance in salmonid species. Front. Genet. 20145, 415.

95. Hjeltnes, B.; Bornø, G.; Jansen, M.D.; Haukaas, A.; Walde, C.S. Fish Health Report 3B—2016; The Norwegian Veterinary Insititute: Oslo, Norway, 2016.

96. Huang, R.; Sun, J.; Luo, Q.; He, L.; Liao, L.; Li, Y.; Guo, F.; Zhu, Z.; Wang, Y. Genetic variations of body weight and GCRV resistance in a random mating population of grass carp. Oncotarget 20156, 35433–35442.

97. Geng, X.; Sha, J.; Liu, S.; Bao, L.; Zhang, J.; Wang, R.; Yao, J.; Li, C.; Feng, J.; Sun, F.; et al. A genome-wide association study in catfish reveals the presence of functional hubs of related genes within QTLs for columnaris disease resistance. BMC Genom. 201516, 196.

98. Shao, C.; Niu, Y.; Rastas, P.; Liu, Y.; Xie, Z.; Li, H.; Wang, L.; Jiang, Y.; Tai, S.; Tian, Y.; et al. Genome-wide SNP identification for the construction of a high-resolution genetic map of Japanese flounder (Paralichthys olivaceus), applications to QTL mapping of Vibrio anguillarum disease resistance and comparative genomic analysis. DNA Res. 201522, 161–170.

99. Robinson, N.; Baranski, M.; Mahapatra, K.D.; Saha, J.N.; Das, S.; Mishra, J.; Das, P.; Kent, M.; Arnyasi, M.; Sahoo, P.K. A linkage map of transcribed single nucleotide polymorphisms in rohu (Labeo rohita) and QTL associated with resistance to Aeromonas hydrophila. BMC Genom. 201415, 541.

100. Wiens, G.D.; Vallejo, R.L.; Leeds, T.D.; Palti, Y.; Hadidi, S.; Liu, S.; Evenhuis, J.P.; Welch, T.J.; Rexroad, C.E., 3rd. Assessment of genetic correlation between bacterial cold water disease resistance and spleen index in a domesticated population of rainbow trout: Identification of QTL on chromosome Omy19. PLoS ONE 20138, e75749.

101. Rodriguez-Ramilo, S.T.; De La Herran, R.; Ruiz-Rejon, C.; Hermida, M.; Fernandez, C.; Pereiro, P.; Figueras, A.; Bouza, C.; Toro, M.A.; Martinez, P.; et al. Identification of quantitative trait loci associated with resistance to viral haemorrhagic septicaemia (VHS) in turbot (Scophthalmus maximus), a comparison between bacterium, parasite and virus diseases. Mar. Biotechnol. 201416, 265–276.

102. Ozaki, A.; Yoshida, K.; Fuji, K.; Kubota, S.; Kai, W.; Aoki, J.Y.; Kawabata, Y.; Suzuki, J.; Akita, K.; Koyama, T.; et al. Quantitative trait loci (QTL) associated with resistance to a monogenean parasite (Benedenia seriolae) in yellowtail (Seriola quinqueradiata) through genome wide analysis. PLoS ONE 20138, e64987.

103. Rodriguez-Ramilo, S.T.; Fernandez, J.; Toro, M.A.; Bouza, C.; Hermida, M.; Fernandez, C.; Pardo, B.G.; Cabaleiro, S.; Martinez, P. Uncovering QTL for resistance and survival time to Philasterides dicentrarchi in turbot (Scophthalmus maximus). Anim. Genet. 201344, 149–157.

104. Baerwald, M.R.; Petersen, J.L.; Hedrick, R.P.; Schisler, G.J.; May, B. A major effect quantitative trait locus for whirling disease resistance identified in rainbow trout (Oncorhynchus mykiss). Heredity 2011106, 920–926.

105. Massault, C.; Franch, R.; Haley, C.; de Koning, D.J.; Bovenhuis, H.; Pellizzari, C.; Patarnello, T.; Bargelloni, L. Quantitative trait loci for resistance to fish pasteurellosis in gilthead sea bream (Sparus aurata). Anim. Genet.201142, 191–203.

106. Gheyas, A.A.; Haley, C.S.; Guy, D.R.; Hamilton, A.; Tinch, A.E.; Mota-Velasco, J.C.; Woolliams, J.A. Effect of a major QTL affecting IPN resistance on production traits in Atlantic salmon. Anim. Genet. 201041, 666–668.

107. Gheyas, A.A.; Houston, R.D.; Mota-Velasco, J.C.; Guy, D.R.; Tinch, A.E.; Haley, C.S.; Woolliams, J.A. Segregation of infectious pancreatic necrosis resistance QTL in the early life cycle of Atlantic Salmon (Salmo salar). Anim. Genet. 201041, 531–536.

108. Houston, R.D.; Gheyas, A.; Hamilton, A.; Guy, D.R.; Tinch, A.E.; Taggart, J.B.; McAndrew, B.J.; Haley, C.S.; Bishop, S.C. Detection and confirmation of a major QTL affecting resistance to infectious pancreatic necrosis (IPN) in Atlantic salmon (Salmo salar). Dev. Biol. 2008132, 199–204.

109. Houston, R.D.; Haley, C.S.; Hamilton, A.; Guy, D.R.; Tinch, A.E.; Taggart, J.B.; McAndrew, B.J.; Bishop, S.C. Major quantitative trait loci affect resistance to infectious pancreatic necrosis in Atlantic salmon (Salmo salar). Genetics2008178, 1109–1115.

110. Li, J.; Boroevich, K.A.; Koop, B.F.; Davidson, W.S. Comparative genomics identifies candidate genes for infectious salmon anemia (ISA) resistance in Atlantic salmon (Salmo salar). Mar. Biotechnol. 201113, 232–241.

111. Gilbey, J.; Verspoor, E.; Mo, T.A.; Sterud, E.; Olstad, K.; Hytterod, S.; Jones, C.; Noble, L. Identification of genetic markers associated with Gyrodactylus salaris resistance in Atlantic salmon Salmo salar. Dis. Aquat. Organ. 200671, 119–129.

112. Sakamoto, T.; Fuji, K.; Kobayashi, K.; Hasegawa, O.; Ozaki, A.; Okamoto, N. Marker-assisted breeding for viral disease resistance in Japanese flounder (Paralichthys olivaceus). Aquaculture 2007272, 4.

113. Fuji, K.; Kobayashi, K.; Hasegawa, O.; Coimbra, M.R.M.; Sakamoto, T.; Okamoto, N. Identification of a single major genetic locus controlling the resistance to lymphocystis disease in Japanese flounder (Paralichthys olivaceus). Aquaculture 2006254, 203–210.

114. Gonen, S.; Baranski, M.; Thorland, I.; Norris, A.; Grove, H.; Arnesen, P.; Bakke, H.; Lien, S.; Bishop, S.C.; Houston, R. Mapping and validation of a major QTL affecting resistance to pancreas disease (salmonid alphavirus) in Atlantic salmon (Salmo salar). Heredity 2015115, 405–414.

115. Munang’andu, H.M.; Santi, N.; Fredriksen, B.N.; Lokling, K.E.; Evensen, O. A Systematic Approach towards Optimizing a Cohabitation Challenge Model for Infectious Pancreatic Necrosis Virus in Atlantic Salmon (Salmo salar L.). PLoS ONE 201611, e0148467.

116. Sommer, S. The importance of immune gene variability (MHC) in evolutionary ecology and conservation. Front. Zool. 20052, 1.

117. Xu, T.-J.; Chen, S.-L.; Ji, X.-S.; Tian, Y.-S. MHC polymorphism and disease resistance to Vibrio anguillarum in 12 selective Japanese flounder (Paralichthys olivaceus) families. Fish Shellfish Immunol. 200825, 213–221.

118. Yang, M.; Wei, J.; Li, P.; Wei, S.; Huang, Y.; Qin, Q. MHC polymorphism and disease resistance to Singapore grouper iridovirus (SGIV) in the orange-spotted grouper, Epinephelus coioidesSci. Bull. 201661, 693–699.

119. Palti, Y.; Nichols, K.M.; Waller, K.I.; Parsons, J.E.; Thorgaard, G.H. Association between DNA polymorphisms tightly linked to MHC class II genes and IHN virus resistance in backcrosses of rainbow and cutthroat trout. Aquaculture 2001194, 283–289.

120. Johnson, N.A.; Vallejo, R.L.; Silverstein, J.T.; Welch, T.J.; Wiens, G.D.; Hallerman, E.M.; Palti, Y. Suggestive association of major histocompatibility IB genetic markers with resistance to bacterial cold water disease in rainbow trout (Oncorhynchus mykiss). Mar. Biotechnol. 200810, 429–437.

121. Du, M.; Chen, S.; Liu, Y.; Niu, B.; Yang, J.; Zhang, B. MHC polymorphism and disease-resistance to Edwardsiella tarda in six turbot (Scophthalmus maximus) families. Chin. Sci. Bull. 201257, 3262–3269.

122. Du, M.; Chen, S.-L.; Liu, Y.-H.; Liu, Y.; Yang, J.-F. MHC polymorphism and disease resistance to vibrio anguillarum in 8 families of half-smooth tongue sole (Cynoglossus semilaevis). BMC Genet. 201112, 1.

123. Gómez, D.; Conejeros, P.; Consuegra, S.; Marshall, S.H. MHC mediated resistance to Piscirickettsia salmonis in salmonids farmed in Chile. Aquaculture 2011318, 15–19.

124. Grimholt, U.; Larsen, S.; Nordmo, R.; Midtlyng, P.; Kjoeglum, S.; Storset, A.; Saebø, S.; Stet, R.J. MHC polymorphism and disease resistance in Atlantic salmon (Salmo salar); facing pathogens with single expressed major histocompatibility class I and class II loci. Immunogenetics 200355, 210–219.

125. Langefors, Å.; Lohm, J.; Grahn, M.; Andersen, Ø.; von Schantz, T. Association between major histocompatibility complex class IIB alleles and resistance to Aeromonas salmonicida in Atlantic salmon. Proc. R. Soc. Lond. B 2001268, 479–485.

126. Kongchum, P.; Sandel, E.; Lutzky, S.; Hallerman, E.M.; Hulata, G.; David, L.; Palti, Y. Association between IL-10a single nucleotide polymorphisms and resistance to cyprinid herpesvirus-3 infection in common carp (Cyprinus carpio). Aquaculture 2011315, 417–421.

127. Kongchum, P.; Palti, Y.; Hallerman, E.M.; Hulata, G.; David, L. SNP discovery and development of genetic markers for mapping innate immune response genes in common carp (Cyprinus carpio). Fish Shellfish Immunol. 201029, 356–361.

128. Heng, J.; Su, J.; Huang, T.; Dong, J.; Chen, L. The polymorphism and haplotype of TLR3 gene in grass carp (Ctenopharyngodon idella) and their associations with susceptibility/resistance to grass carp reovirus. Fish Shellfish Immunol. 201130, 45–50.

129. Su, J.; Heng, J.; Huang, T.; Peng, L.; Yang, C.; Li, Q. Identification, mRNA expression and genomic structure of TLR22 and its association with GCRV susceptibility/resistance in grass carp (Ctenopharyngodon idella). Dev. Comp. Immunol. 201236, 450–462.

130. Wang, L.; Su, J.; Yang, C.; Wan, Q.; Peng, L. Genomic organization, promoter activity of grass carp MDA5 and the association of its polymorphisms with susceptibility/resistance to grass carp reovirus. Mol. Immunol. 201250, 236–243.

131. Wan, Q.; Su, J.; Chen, X.; Yang, C.; Chen, L.; Yan, N.; Zhang, Y. Genomic sequence comparison, promoter activity, SNP detection of RIG-I gene and association with resistance/susceptibility to grass carp reovirus in grass carp (Ctenopharyngodon idella). Dev. Comp. Immunol. 201339, 333–342.

132. Wan, Q.; Wang, L.; Su, J.; Yang, C.; Peng, L.; Chen, L. Genetic structure, polymorphism identification of LGP2 gene and their relationship with the resistance/susceptibility to GCRV in grass carp, Ctenopharyngodon idellaGene 2013521, 166–175.

133. Fu, G.H.; Bai, Z.Y.; Xia, J.H.; Liu, F.; Liu, P.; Yue, G.H. Analysis of two lysozyme genes and antimicrobial functions of their recombinant proteins in Asian seabass. PLoS ONE 20138, e79743.

134. Fu, G.H.; Bai, Z.Y.; Xia, J.H.; Liu, X.J.; Liu, F.; Wan, Z.Y.; Yue, G.H. Characterization of the LECT2 gene and its associations with resistance to the big belly disease in Asian seabass. Fish Shellfish Immunol. 201437, 131–138.

135. Sahoo, P.; Das, S.; Mahapatra, K.D.; Saha, J.N.; Baranski, M.; Ødegård, J.; Robinson, N. Characterization of the ceruloplasmin gene and its potential role as an indirect marker for selection to Aeromonas hydrophila resistance in rohu, Labeo rohita. Fish Shellfish Immunol. 201334, 1325–1334.

136. Sun, B.; Robertsen, B.; Wang, Z.; Liu, B. Identification of an Atlantic salmon IFN multigene cluster encoding three IFN subtypes with very different expression properties. Dev. Comp. Immunol. 200933, 547–558.

Related topics
Authors:
Jorge Galindo-Villegas
Recommend
Comment
Share
Profile picture
Would you like to discuss another topic? Create a new post to engage with experts in the community.
Join Engormix and be part of the largest agribusiness social network in the world.