Next Article in Journal
The Effect of Name and Narrative Voice in Online Adoption Profiles on the Length of Stay of Sheltered Cats in the UK
Next Article in Special Issue
Combination of Freezing, Low Sodium Brine, and Cold Smoking on the Quality and Shelf-Life of Sea Bass (Dicentrarchus labrax L.) Fillets as a Strategy to Innovate the Market of Aquaculture Products
Previous Article in Journal
Wild-Type Zebrafish (Danio rerio) Larvae as a Vertebrate Model for Diabetes and Comorbidities: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Mechanisms Regulating Muscle Plasticity in Fish

1
Division of Animal and Nutritional Sciences, West Virginia University, Morgantown, WV 26506-6108, USA
2
USDA ARS National Center for Cool and Cold Water Aquaculture, Kearneysville, WV 25430, USA
*
Author to whom correspondence should be addressed.
Submission received: 2 December 2020 / Revised: 24 December 2020 / Accepted: 25 December 2020 / Published: 30 December 2020
(This article belongs to the Special Issue The Future of Aquaculture Research)

Abstract

:

Simple Summary

Muscle plasticity is defined as the ability of the muscle to respond to changes in environmental conditions. Muscle plasticity is exceptionally dynamic in fish; this is attributed in part to their ectothermic (cold-blooded) nature and ability of indeterminate or continual growth, throughout their lifespans. The molecular mechanisms regulating muscle growth in fish are not completely characterized; however, recent advancements have established that microRNAs and DNA methylation are important mechanisms regulating muscle plasticity. This review examines these mechanisms and describes how they are regulated by genetic and environmental (i.e., nutrition, temperature) factors and they in turn affect muscle growth and plasticity in fish.

Abstract

Growth rates in fish are largely dependent on genetic and environmental factors, of which the latter can be highly variable throughout development. For this reason, muscle growth in fish is particularly dynamic as muscle structure and function can be altered by environmental conditions, a concept referred to as muscle plasticity. Myogenic regulatory factors (MRFs) like Myogenin, MyoD, and Pax7 control the myogenic mechanisms regulating quiescent muscle cell maintenance, proliferation, and differentiation, critical processes central for muscle plasticity. This review focuses on recent advancements in molecular mechanisms involving microRNAs (miRNAs) and DNA methylation that regulate the expression and activity of MRFs in fish. Findings provide overwhelming support that these mechanisms are significant regulators of muscle plasticity, particularly in response to environmental factors like temperature and nutritional challenges. Genetic variation in DNA methylation and miRNA expression also correlate with variation in body weight and growth, suggesting that genetic markers related to these mechanisms may be useful for genomic selection strategies. Collectively, this knowledge improves the understanding of mechanisms regulating muscle plasticity and can contribute to the development of husbandry and breeding strategies that improve growth performance and the ability of the fish to respond to environmental challenges.

1. Introduction

Fish species that exhibit indeterminate growth display an exceptional ability for continual muscle growth throughout their life by an increase in muscle fiber number (hyperplasty) and in muscle fiber size (hypertrophy) [1,2,3,4,5,6,7]. In fish, myogenesis starts during embryonic development by somites from the mesodermal layer. Somites differentiate to myogenic precursor cells (MPCs) that are stored as a reservoir between the basal lamina and sarcolemma of mature muscle bundles and are sequestered during adult myogenesis [8]. Myogenesis is a complex mechanism initiated by activation, proliferation, differentiation, and maturation of MPCs; these processes are orchestrated by various myogenic regulatory factors (MRFs), signaling pathways, non-coding RNAs, and epigenetic mechanisms. The functional aspects of MRFs and signaling pathways in mammalian species are well documented in previous reviews [9,10,11,12], hence their role in skeletal muscle development is briefly discussed in the current review. Our focus is to discuss the role of epigenetic and microRNA mechanisms in myogenesis and their regulation by biological factors thus affecting muscle plasticity in fish.

2. Myogenic Regulatory Factors

Myogenic regulatory factors are basic helix-loop-helix transcription factors that drive the process of myogenesis, including MPC proliferation and differentiation (Table 1). Preferential expression of MRFs determines progression of MPCs [13]. It is accepted that MRF function is generally conserved in animals; this concept is supported by MRF protein sequences with high similarity and consistent expression patterns throughout MPC progression. Myf5 is an important MRF that functions to promote the commitment of satellite cells to myogenic lineage. The committed MPCs are cells that express MRF specific to satellite cells, including Pax7 [14,15], HGF receptor c-met [15], and Syndecan-4 [16]. Pax7 is necessary for the maintenance of satellite cells state [8,17,18,19,20]. Absence of Pax7 expression leads to defective muscle differentiation through cell cycle arrest and precocious differentiation [21,22]. MPCs enter a proliferative phase and are capable of asymmetric and symmetric cell division to either self-renew or differentiate into myoblasts. The self-renewed cells express high levels of Pax7 [18] and those that differentiate express high MyoD with a decrease in Pax7 expression [23]. Thus, MyoD is one of the major factors determining the myogenic cell fate of MPCs [24,25]. In addition to the MRFs necessary for satellite cell maintenance and differentiation, the expression of Mrf4 and Myogenin after proliferation and during differentiation marks the differentiation process to form a myofiber [26,27].
Signaling molecules play a significant role in regulating the expression of these MRFs, thus determining the fate of myogenesis. TGF-β family members are responsible for satellite cell maintenance while hedgehog signaling is responsible for transition from proliferative to quantile mitosis. Satellite cell markers Pax7 and Pax3 are downregulated and differentiation factor MyoD is upregulated by hedgehog in teleosts, thus determining myogenic commitment. Additionally, insulin-like growth factor-I (IGF-I) and insulin regulate myogenesis by binding to receptor tyrosine kinase and initiating signaling cascades. Growth factor signaling enhances myogenic proliferation and differentiation via mitogen-activated protein kinase (MAPK), Ras/Raf, and PI3K pathways thus controlling muscle regeneration, MPCs proliferation, and differentiation, respectively [34]. The signaling pathways initiate a cascade of reactions by phosphorylating different intermediates, including target of rapamycin (TOR), thereby activating MRFs and ultimately triggering gene expression responsible for muscle synthesis [35].

3. Epigenetics and DNA Methylation

Chemical modification of DNA bases was first reported in Tubercle bacillus and later in calf thymus DNA [36,37]. Further establishing of the importance of DNA methylation in transcriptional regulation of eukaryotic gene expression was reported by two simultaneous studies [38,39]. Since then, studies have reported the presence of such modifications in various organisms including prokaryotes, fungi, plants, and animals, including fish. These modifications are post-synthetic, as the methyl group is added after the nucleotides are incorporated into the DNA. Methylation of bases is carried out by DNA methyltransferases (Dnmt) and the methyl group is donated by S-adenosyl methionine. DNA methylation in mammals is carried out by three Dnmts: Dnmt1, Dnmt3a, and Dnmt3b. However, eight Dnmt genes were identified in zebrafish and included Dnmt1 and Dnmt2, while the rest were similar to the mammalian Dnmt3 [40,41,42]. Dnmt1 is a maintenance methyltransferase mainly functioning in methylating hemi-methylated DNA strands during replication. Dnmt3a and Dnmt3b are de novo methyltransferases methylating nascent DNA or hemi-methylated DNA. De novo methyltransferases are essential for laying methylation marks during embryo implantation and early development [43]. The majority of base modifications in eukaryotes are on cytosines, while some of the unicellular organisms present methylation on adenosine [44,45]. A methyl group is added to the carbon at fifth position of cytosine resulting in 5mC. Until recently it was believed that methylation of cytosine is restricted to CpG dinucleotide. Advanced sequencing techniques as well as the ability to map sequences at base pair resolution disclosed non-CpG methylations. Non-CpG methylations are less frequent than CpG methylations and are seen in mammalian oocytes [46], adult brain [47], and embryonic stem cells [48]. In general, DNA methylation is associated with transcriptional silencing, although there are exceptions [49]. Introduction of the methyl group on cytosine affects the binding of proteins such as repressors, histones, and hormone receptors to the DNA indicating regulation of gene expression [50,51,52] by sterically preventing binding. The methylated regions act as binding sites for certain transcriptional repressors including MeCP2, Mbd1, and Mbd2 leading to closed chromatin [53].

3.1. DNA Methylation during Skeletal Myogenesis

Studies in mammals confirmed the regulatory role of DNA methylation during myogenesis, including activation of MPC, proliferation, and differentiation. Comparative studies to understand the methylation landscape among myogenic differentiating cells and mature skeletal muscle suggested loss of methylation in mature fibers. In mice differentiating muscle cells exhibit approximately 90% higher methylation rates when compared to mature myofibers [29]. Differential expression of enzymes involved in DNA methylation was observed in a stage specific manner; Dnmt1 was upregulated during activation and downregulated during differentiation. Reduced expression of de novo methyltransferases Dnmt3a and demethylating enzymes Tet1, Tet2, and Tet3 during muscle precursor cell activation was reported while Dnmt3b remained unchanged [54,55,56,57]. The role of DNA methylation in cell fate commitment and in progression of myogenesis was supported by several mammalian studies using 5-azacytidine and antisense RNA to inhibit Dnmt1 and DNA methylation [58,59]. Proliferating myoblasts exposed to 5-azacytidine exhibited increased expression of myogenic genes including Myogenin. This indicated a functional role of DNA methylation in permitting binding of myogenic transcription factors to their target genes promoting differentiation [60]. Hyper and hypomethylation of various MRFs and genes involved in the process of myogenesis are listed in Table 1. Collectively these studies support the important role of DNA methylation in various stages on myogenesis. Although most of the studies represent CpG methylations, the importance of non-CpG methylations is yet to be identified. Additionally, technical methods used could not differentiate between methylated and hemi-methylated cytosines, hence studies understanding such differences are necessary to know the functional impact of DNA methylation in myogenesis.

3.2. Epigenetic Regulation of Muscle Plasticity in Fish

Most advances in the knowledge of epigenetic regulation of muscle plasticity in fish have occurred only within the last five years; findings have provided support for DNA methylation as a mechanism affecting muscle growth (Table 2). Comparing the methylome of fast and comparatively slow growing tilapia indicated approximately 1000 differentially methylated CpGs in both males and females, although there was very little overlap between the sexes [61]. These findings indicate that variations in DNA methylations are associated with faster growth, providing strong support for epigenetic mechanisms as a significant regulator of muscle plasticity. In particular, the autophagy-related gene Atg14 displayed a high association of methylation with growth in male tilapia, suggesting that suppression of muscle protein degradation contributes to muscle growth. Research in zebrafish reports similar findings; in muscle the autophagy-related genes Atg4b and Lc3b are both hypomethylated and up-regulated during nutrient deprivation [62], supporting that DNA methylation is responsible for starvation-induced loss of muscle protein. Nutritional regulation of muscle growth via epigenetic modifications is supported by additional studies in rainbow trout indicating hypermethylation in muscle of fish consuming higher dietary carbohydrates [63] and regulation of DNA methyltransferase expression by dietary protein in the Senegalese sole [64]. This concept is further supported by a nutritional programming study in tilapia in which injections of glucose into the yolk of alevins were associated with DNA hypomethylation in muscle 20 weeks post-injection [65], suggesting the existence of epigenetic mechanisms at the origin of programming that affect muscle growth.
Several studies have characterized the role of DNA methylation in response to temperature variation during early rearing stages. In Atlantic salmon, higher embryonic incubation temperature (4 °C vs. 8 °C) increases post-embryonic growth rates; in first-feeding larvae this response was associated with both reduced DNA methylation of the Myogenin promoter and higher Myogenin expression [66]. Effects of incubation temperature on muscle Dnmt expression extended to the parr stage in which expression of Dnmt1 and Dnmt3a increased with higher larval rearing temperature. Similarly, European sea bass larvae incubated at higher temperatures exhibited growth benefits, differential genomic methylation patterns, and increased expression of Myogenin, Dnmt1, and Dnmt3 [67]. Comparable findings were observed in the Senegalese sole; higher early rearing temperature enhanced hyperplastic muscle growth during metamorphosis [68], a response that was correlated with reduced methylation of the Myogenin promoter, increased Myogenin expression, and regulation of Dnmt1 and Dnmt3a expression [69]. In Atlantic cod, exposure to continuous illumination during early juvenile development improved growth potential and was associated with higher Dnmt1 and Dnmt3a expression in fast muscle. [70]. Although the directional regulation of Dnmt expression is not consistently associated with the faster growth phenotype, these studies provide evidence that temporal effects of early rearing conditions on muscle growth are regulated in part through modification of DNA methylation.
Although most studies have focused on changes in muscle Dnmt expression as an indicator of regulation of DNA methylation capacity, fewer have investigated changes in methylation of MRFs that affect myogenesis and muscle plasticity. As previously mentioned, in two studies improved growth rates induced by higher embryonic incubation temperatures corresponded to hypomethylation of the Myogenin promotor and increased expression in muscle [66,67]. In rainbow trout 17β-estradiol increased methylation within exon 1 of Myod and decreased expression of the Myod gene, providing evidence for down-regulation of a MRF by a steroid treatment that also promotes muscle atrophy [71]. Also in rainbow trout, unique histone methylation patterns of the three paralogous Pax7 genes were detected during in vitro myogenesis [74]. Expression of Pax7a2 exhibited decreased expression with decreased H3K27 trimethylation. In contrast, Pax7b expression increased and was correlated with decreased H3K9me3 and H3K27me3.

4. MicroRNA Regulation of Myogenesis

Small RNAs, particularly microRNAs (miRNAs), are extensively studied as post-transcriptional microregulators of gene expression governing either mRNA stability, rates of translation, or both [75]. Studies in mammals provide evidence for their transport [76] and nuclear existence [77], suggesting functions other than translational inhibition. Similarly, the presence of miRNA in exosomes [78] advocates their secretion and circulation through body fluids [79,80,81] to control gene expression in recipient cells [82]. Besides continuous efforts to understand various functional roles of miRNA, their biogenesis is well established [83]. Primary transcripts of miRNA are either generated by transcription of individual miRNA genes or by processing of intronic regions. MicroRNAs bind to target mRNA by either perfect or imperfect base-pairing, leading to translational repression or mRNA degradation. In addition, nuclear miRNAs bind to promoters and either activate or inhibit gene expression. The distinguishable characteristics or functional properties of activating and inhibiting miRNAs are not well understood [83].

4.1. Functional Regulation of miRNA in Muscle

Myogenesis involves orchestrated interaction of various mechanisms involving signaling pathways and coordinated gene expression of various myogenic regulatory factors (MRFs). The role of miRNA in myogenesis was first established in Dicer mutant mice exhibiting reduced muscle miRNA, perinatal death, and decreased skeletal muscle with abnormal myofiber morphology and apoptosis of myoblasts [84]. This study established the important role of miRNAs in vertebrate muscle development. MicroRNAs with muscle specific expression and function are designated as myomiRNAs. The involvement of myomiRNAs in MPC proliferation and differentiation is established [85,86,87]. For example, miR-1 and miR-206 promote myoblast differentiation through anti-proliferative effects by repressing expression of the MRFs Pax3, Pax7, and Notch 1. The myomiRNA miR-133 also suppresses proliferation and promotes differentiation through various mechanisms, including repression of the mitogen-activated protein kinase pathway, cyclin D2, and uncoupling protein-2. Nachtigall et al. [88] compared the evolution and organization of myomiRNAs in cartilaginous and bony-fish genomes through genome-wide comparative analysis and concluded synteny in myomiRNA distribution.

4.2. MicroRNAs Targeting Genes Involved in Muscle Development

In several fish species, including rainbow trout [89,90], zebrafish [91], Japanese flounder [92], sea bass [93], Nile tilapia [94], Chinese perch [95], and common carp [96], the most abundant miRNAs expressed in skeletal muscle are reported as miR-1, miR-206, and miR-133a. In vivo and in vitro studies in mammalian systems have established that interaction of miR-1, miR-206, and miR-133a with their target genes regulates skeletal muscle cell proliferation and differentiation [97,98,99]. These studies shed light on the regulatory network among the miRNA and MRFs. Mature miR-1 and miR-133 are derived from one polycistronic pre-miRNA targeting two different genes. miR-1 promotes myogenesis by targeting histone deacetylase 4, a transcriptional repressor of muscle gene expression, while miR-133 targets serum response factor which enhances muscle cell proliferation [97]. In zebrafish, miR-1 and miR-133 explain more than 54% of miRNA-induced regulation of gene expression, that when down-regulated cause disruption of actin organization in sarcomere assembly [91]. Additionally, muscle specific expression of miR-206 is controlled by Myod [99] which along with miR-1 targets and represses pax7 that is a marker for proliferation [98]. Regulatory expression of miR-206 by Myod is also associated with the negative regulation of Follistatin-like 1 and Utrophin genes which are necessary for muscle cell differentiation [99]. MicroRNA-206 is also involved in negative regulation of Igf-1 expression in tilapia skeletal muscle [100]. In addition, findings in tilapia and Chinese perch report negative regulation of Myod expression by miR-203b and miR-143, respectively; both miR-203b and miR-143 bind to the 3′-untranslated region (UTR) of MyoD, suppressing its expression [101,102]. Further promoting growth is miR-181a-5p that represses expression of Myostatin-b in tilapia [103]. Repression of Myostatin expression by miR-2014 and miR-1231-5p is also described in yellow croaker (Larimichthys crocea) [104].
Unique miRNA expression profiles are reported in muscle during different developmental stages in common carp [96], the ray-finned Schizothorax prenanti [105], and pacu (Piaractus mesopotamicus) [106], supporting a significant role in maintenance and growth of this tissue. Genetic and environmental factors are also established as significant regulators of miRNA in fish muscle (Table 3). Differences in expression of miRNAs between fish with divergent growth phenotypes suggest that genetic variation in miRNA expression is a significant mechanism affecting the capacity for muscle growth. These findings were reported in rainbow trout [107], tilapia [94], Chinese perch [108], and blunt snout sea bream [109]. Notable was an up-regulation of miR-133 in fast-growing tilapia [94]. Expression of miR-133, miR-181a-5p, and miR-206 is also correlated with body weight in rainbow trout, although indirectly [107]. However, miR-1, miR-133, and miR-206 were down-regulated in fast-growing, compared to slow-growing, Chinese perch [108]. Variation in miRNA expression that correlates with body weight warrants additional research that investigates whether genetic variation in miRNA expression has value as genomic markers for growth phenotypes.

4.3. MicroRNA Regulation of Muscle Cell Fate

MicroRNAs can determine muscle cell fate very early during somatogenesis. Fish exhibit two distinct muscle types, slow and fast muscles, and the role of miRNAs in muscle cell type determination was reported by studies using fish as model species. Ubiquitously expressing miR-214 synchronizes Hedgehog signaling in zebrafish by translational repression of its negative regulator Su(Fu), thus coordinating a balance among slow and fast muscle cell types. Knockdown of miR-214 in zebrafish embryos led to the development of fewer to a complete absence of slow muscle cells [120], and hence established miRNA as necessary for development of slow muscle cells. Similarly, knockdown and overexpression of miR-3906 in zebrafish embryos affected the fast muscle phenotype [121] through its function in maintenance of calcium ion concentration homeostasis specifically in fast muscles. Knockdown of miR-3906 increases the expression of its target gene Homer-1b, which in turn up-regulates fast muscle specific gene Fmhc4 and calcium sensitive gene Atp2a1, thus causing a surge in calcium ions concentration and disorganized sarcomeric actin in fast muscle resulting in swimming abnormality. However, over expression of miR-3906 decreases calcium ion concentration, resulting in bent bodies and shortened tails in zebrafish [121]. Another important miRNA extensively studied for its involvement in muscle cell fate determination is miR-499; it is highly expressed in slow skeletal muscle in Nile tilapia, pacu, and rainbow trout compared to fast skeletal muscle [88,106,122]. A regulatory network involving functional repression of Sox6 by miR-499 through Prdm1 for the maintenance of slow-twitch muscle was also reported in zebrafish [123], allowing restricted expression of Sox6 in fast-twitch muscle. In the Chinese perch, miRNA profiling in fast and slow muscle suggests that miR-181a-5p, miR-143, and miR-103 have central roles in regulating the performance of the muscle types [95,102,124]. All together these studies emphasize the regulatory mechanisms of miRNA in balanced determination of muscle cell fate and muscle-specific function.

4.4. Biological Factors Affecting miRNA Expression in Muscle

Understanding the role of miRNAs as a mechanism regulating muscle function during physiological perturbation has been the focus of numerous studies in fish. Perhaps one of the most significant environmental factors affecting muscle growth in fish is the availability of nutrients; several studies support that both nutrient intake and diet composition affect miRNA expression in muscle. Experimental designs that involve feed deprivation and refeeding have been valuable for identifying miRNA biomarkers for anabolic and catabolic responses. When Chinese perch [112] and grass carp [113] are subjected to feed deprivation and refeeding, miR-206, miR-133a-3p, and miR-181a-5p are up-regulated within just 1–3 h of refeeding, supporting that miRNA-induced regulation of MRFs like Myod and Myostatin are significant for rapid recovery growth. Additional in vivo and in vitro studies have investigated effects of specific nutrients on miRNA expression using rainbow trout and MPCs derived from its muscle. Methionine deficiency arrested cell differentiation and reduced expression of miRNAs, including miR-133a and the MRFs Myod and Myogenin, which were rescued by the introduction of methionine [114]. Analogous to in vitro experiments, decreased expression of miR-133a was also observed in muscle of rainbow trout consuming a methionine-deficient diet [115]. In tilapia muscle it was determined that regulation of Kruppel-like factor-15 (Klf15), a key regulator of branched-chain amino acid metabolism, during feed deprivation is mediated by changes in miR-125a-3p expression [116]. This mechanism is likely significant for the metabolic and physiological adaptations to nutrient deprivation in fish. Furthermore, miRNAs with established roles as regulators of myogenesis (i.e., miR-1, miR-133, miR-206) are affected by consumption of first feeds in the rainbow trout alevin [117] and Atlantic cod [118].
Temperature is an additional factor that affects miRNA expression. In zebrafish, changes in muscle growth trajectories (hyperplasia vs. hypertrophy) induced by manipulation of embryonic incubation temperature were associated with differential expression of miRNAs, with increased Myogenin expression in the higher temperatures that improved growth performance [110]. Similar findings are also reported in Senegalese sole, in which thermal plasticity of muscle growth was correlated with increased Myogenin expression and regulation of miRNAs with putative roles in muscle development and nutrient metabolism, such as miR-181-5p/3p and miR-206-3p [111,125]. These studies point towards miRNA-related mechanisms that regulate muscle plasticity during temperature manipulation and fluctuation; this information can contribute to the optimization of husbandry strategies to enhance muscle growth or improve fish health and performance. The need for such studies to understand the consequences of climate change on fish species is increasing.
In addition to biological factors of exogenous origin, endogenous factors such as hormones are established as regulators of muscle physiology through mechanisms involving miRNAs. Physiological changes in salmonid skeletal muscle during sexual maturation are characterized by protein degradation and lipid loss [126,127]. By utilizing the spawning salmonid as a model for muscle atrophy, studies in rainbow trout have determined that miRNAs are important for the mobilization of muscle nutrient reserves during the energy intensive spawning period. Through comparing the transcriptome of sterile and spawning rainbow trout, 28 miRNAs were associated with the lncRNA-mRNA-microRNA gene network described as the muscle ”degradome” [119]. The steroid hormone, 17β-estradiol, is up-regulated only during sexual maturation in female rainbow trout and has been implicated as a biological factor directly regulating miRNA in muscle [90]. Furthermore, 17β-estradiol-induced responses in muscle included regulation of miR-133, miR-206, and miR-499, along with differential regulation of MRFs like Pax7 and Myod. In an additional area of hormone-related research, efforts to understand the effects of thyroid hormone influence on histone deacetylase (Hdac4) lead to the discovery of miRNA regulation of muscle development during metamorphosis in the Japanese halibut [105]. This study presented evidence of an interaction between signaling pathways, epigenetic regulators like Hdac4, and miRNAs (miR-1 and miR-133).

5. Conclusions

Within the last decade significant advancements in the knowledge of molecular regulation of gene expression have established that miRNAs and DNA methylation are significant mechanisms affecting expression of myogenic regulatory factors and muscle plasticity in fish. Genetic, environmental, and physiological factors cause differential expression of miRNAs and DNA/gene methylation in muscle that are directly linked or correlated with growth responses. These findings are reported in model species like the zebrafish and stickleback as well as fish species significant for aquaculture such as rainbow trout, Atlantic salmon, Chinese perch, and tilapia. Continued advancements in our understanding of the molecular mechanisms regulating muscle plasticity is central for the development of novel genetic markers to aid in selective breeding for enhanced growth. It is also valuable for the development of husbandry strategies that improve the capacity for muscle growth through, for example, diet, photoperiod, or temperature manipulation that enhance aquaculture production efficiency and advance global food security.

Author Contributions

Writing—original draft preparation, P.K. and B.M.C.; writing—review and editing, J.Y. and B.M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. AlamiDurante, H.; Fauconneau, B.; Rouel, M.; Escaffre, A.M.; Bergot, P. Growth and multiplication of white skeletal muscle fibres in carp larvae in relation to somatic growth rate. J. Fish Biol. 1997, 50, 1285–1302. [Google Scholar] [CrossRef]
  2. Higgins, P.J. Metabolic differences between Atlantic salmon (Salmo salar) parr and smolts. Aquaculture 1985, 45, 33–53. [Google Scholar] [CrossRef]
  3. Koumans, J.T.M.; Akster, H.A.; Booms, G.H.R.; Osse, J.W.M. Growth of Carp (Cyprinus-Carpio) White Axial Muscle—Hyperplasia and Hypertrophy in Relation to the Myonucleus Sarcoplasm Ratio and the Occurrence of Different Subclasses of Myogenic Cells. J. Fish Biol. 1993, 43, 69–80. [Google Scholar] [CrossRef]
  4. Mommsen, T.P. Paradigms of growth in fish. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2001, 129, 207–219. [Google Scholar] [CrossRef]
  5. Patruno, M.; Radaelli, G.; Mascarello, F.; Carnevali, M.C. Muscle growth in response to changing demands of functions in the teleost Sparus aurata (L.) during development from hatching to juvenile. Anat. Embryol. 1998, 198, 487–504. [Google Scholar] [CrossRef]
  6. Stickland, N.C. Growth and development of muscle fibres in the rainbow trout (Salmo gairdneri). J. Anat. 1983, 137, 323–333. [Google Scholar]
  7. Weatherly, A.H.; Gill, H.S.; Lobo, A.F. Recruitment and maximal diameter of axial muscle fibres in teleosts and their relationship to somatic growth and ultimate size. J. Fish Biol. 1988, 33, 351–359. [Google Scholar] [CrossRef]
  8. Seale, P.; Sabourin, L.A.; Girgis-Gabardo, A.; Mansouri, A.; Gruss, P.; Rudnicki, M.A. Pax7 is required for the specification of myogenic satellite cells. Cell 2000, 102, 777–786. [Google Scholar] [CrossRef] [Green Version]
  9. Asfour, H.A.; Allouh, M.; Said, R.S. Myogenic regulatory factors: The orchestrators of myogenesis after 30 years of discovery. Exp. Biol. Med. 2018, 243, 118–128. [Google Scholar] [CrossRef]
  10. Hernandez-Hernandez, M.; García-González, E.G.; Brun, C.E.; Rudnicki, M.A. The myogenic regulatory factors, determinants of muscle development, cell identity and regeneration. Semin. Cell Dev. Biol. 2017, 72, 10–18. [Google Scholar] [CrossRef]
  11. Zammit, P.S. Function of the myogenic regulatory factors Myf5, MyoD, Myogenin and MRF4 in skeletal muscle, satellite cells and regenerative myogenesis. Semin. Cell Dev. Biol. 2017, 72, 19–32. [Google Scholar] [CrossRef] [PubMed]
  12. Zanou, N.; Gailly, P. Skeletal muscle hypertrophy and regeneration: Interplay between the myogenic regulatory factors (MRFs) and insulin-like growth factors (IGFs) pathways. Cell. Mol. Life Sci. 2013, 70, 4117–4130. [Google Scholar] [CrossRef] [PubMed]
  13. Watabe, S. Myogenic regulatory factors and muscle differentiation during ontogeny in fish. J. Fish Biol. 1999, 55, 1–18. [Google Scholar] [CrossRef]
  14. Seger, C.; Hargrave, M.; Wang, X.; Chai, R.J.; Elworthy, S.; Ingham, P.W. Analysis of Pax7 expressing myogenic cells in zebrafish muscle development, injury, and models of disease. Dev. Dyn. 2011, 240, 2440–2451. [Google Scholar] [CrossRef]
  15. Hollway, G.E.; Bryson-Richardson, R.J.; Berger, S.; Cole, N.J.; Hall, T.E.; Currie, P.D. Whole-Somite Rotation Generates Muscle Progenitor Cell Compartments in the Developing Zebrafish Embryo. Dev. Cell 2007, 12, 207–219. [Google Scholar] [CrossRef] [Green Version]
  16. Froehlich, J.M.; Galt, N.J.; Charging, M.J.; Meyer, B.M.; Biga, P.R. In vitro indeterminate teleost myogenesis appears to be dependent on Pax3. Vitr. Cell. Dev. Biol. Anim. 2013, 49, 371–385. [Google Scholar] [CrossRef] [Green Version]
  17. Kuang, S.; Chargé, S.B.; Seale, P.; Huh, M.; Rudnicki, M.A. Distinct roles for Pax7 and Pax3 in adult regenerative myogenesis. J. Cell Biol. 2006, 172, 103–113. [Google Scholar] [CrossRef] [Green Version]
  18. Olguin, H.C.; Olwin, B.B. Pax-7 up-regulation inhibits myogenesis and cell cycle progression in satellite cells: A potential mechanism for self-renewal. Dev. Biol. 2004, 275, 375–388. [Google Scholar] [CrossRef] [Green Version]
  19. Oustanina, S.; Hause, G.; Braun, T. Pax7 directs postnatal renewal and propagation of myogenic satellite cells but not their specification. Embo J. 2004, 23, 3430–3439. [Google Scholar] [CrossRef] [Green Version]
  20. Relaix, F.; Montarras, D.; Zaffran, S.; Gayraud-Morel, B.; Rocancourt, D.; Tajbakhsh, S.; Mansouri, A.; Cumano, A.; Buckingham, M. Pax3 and Pax7 have distinct and overlapping functions in adult muscle progenitor cells. J. Cell Biol. 2005, 172, 91–102. [Google Scholar] [CrossRef]
  21. Günther, S.; Kim, J.; Kostin, S.; Lepper, C.; Fan, C.-M.; Braun, T. Myf5-Positive Satellite Cells Contribute to Pax7-Dependent Long-Term Maintenance of Adult Muscle Stem Cells. Cell Stem Cell 2013, 13, 590–601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Von Maltzahn, J.; Jones, A.E.; Parks, R.J.; Rudnicki, M.A. Pax7 is critical for the normal function of satellite cells in adult skeletal muscle. Proc. Natl. Acad. Sci. USA 2013, 110, 16474–16479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zammit, P.S.; Golding, J.P.; Nagata, Y.; Hudon, V.; Partridge, T.A.; Beauchamp, J.R. Muscle satellite cells adopt divergent fates: A mechanism for self-renewal? J. Cell Biol. 2004, 166, 347–357. [Google Scholar] [CrossRef] [PubMed]
  24. Megeney, L.A.; Kablar, B.; Garrett, K.; Anderson, E.J.; Rudnicki, A.M. MyoD is required for myogenic stem cell function in adult skeletal muscle. Genes Dev. 1996, 10, 1173–1183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. White, J.D.; Scaffidi, A.; Davies, M.; McGeachie, J.; Rudnicki, M.A.; Grounds, M.D. Myotube formation is delayed but not prevented in MyoD-deficient skeletal muscle: Studies in regenerating whole muscle grafts of adult mice. J. Histochem. Cytochem. 2000, 48, 1531–1543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Füchtbauer, E.-M.; Westphal, H. MyoD and myogenin are coexpressed in regenerating skeletal muscle of the mouse. Dev. Dyn. 1992, 193, 34–39. [Google Scholar] [CrossRef] [PubMed]
  27. Reuveni, Z.Y.-; Rivera, A.J. Temporal expression of regulatory and structural muscle proteins during myogenesis of satellite cells on isolated adult rat fibers. Dev. Biol. 1994, 164, 588–603. [Google Scholar] [CrossRef] [Green Version]
  28. Tsumagari, K.; Baribault, C.; Terragni, J.; Varley, K.E.; Gertz, J.; Pradhan, S.; Badoo, M.; Crain, C.M.; Song, L.; Crawford, G.E.; et al. Early de novo DNA methylation and prolonged demethylation in the muscle lineage. Epigenetics 2013, 8, 317–332. [Google Scholar] [CrossRef] [Green Version]
  29. Carrio, E.; Díez-Villanueva, A.; Lois, S.; Mallona, I.; Cases, I.; Forn, M.; Peinado, M.A.; Suelves, M. Deconstruction of DNA Methylation Patterns During Myogenesis Reveals Specific Epigenetic Events in the Establishment of the Skeletal Muscle Lineage. Stem Cells 2015, 33, 2025–2036. [Google Scholar] [CrossRef]
  30. Brunk, B.P.; Goldhamer, D.J.; Emerson, C.P., Jr. Regulated Demethylation of the myoD Distal Enhancer during Skeletal Myogenesis. Dev. Biol. 1996, 177, 490–503. [Google Scholar] [CrossRef] [Green Version]
  31. Lucarelli, M.; Fuso, A.; Strom, R.; Scarpa, S. The Dynamics of Myogenin Site-specific Demethylation Is Strongly Correlated with Its Expression and with Muscle Differentiation. J. Biol. Chem. 2000, 276, 7500–7506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Miyata, K.; Miyata, T.; Nakabayashi, K.; Okamura, K.; Naito, M.; Kawai, T.; Takada, S.; Kato, K.; Miyamoto, S.; Hata, K.; et al. DNA methylation analysis of human myoblasts during in vitro myogenic differentiation: De novo methylation of promoters of muscle-related genes and its involvement in transcriptional down-regulation. Hum. Mol. Genet. 2015, 24, 410–423. [Google Scholar] [CrossRef] [PubMed]
  33. Terragni, J.; Zhang, G.; Sun, Z.; Pradhan, S.; Song, L.; Crawford, G.E.; Lacey, M.; Ehrlich, M. Notch signaling genes Myogenic DNA hypomethylation and 5-hydroxymethylcytosine. Epigenetics 2014, 9, 842–850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Jiang, B.-H.; Aoki, M.; Zheng, J.Z.; Li, J.; Vogt, P.K. Myogenic signaling of phosphatidylinositol 3-kinase requires the serine-threonine kinase Akt/protein kinase B. Proc. Natl. Acad. Sci. USA 1999, 96, 2077–2081. [Google Scholar] [CrossRef] [Green Version]
  35. Florini, J.R.; Ewton, D.Z.; Coolican, S.A. Growth Hormone and the Insulin-Like Growth Factor System in Myogenesis. Endocr. Rev. 1996, 17, 481–517. [Google Scholar] [CrossRef] [Green Version]
  36. Hotchkiss, R.D. The quantitative separation of purines, pyrimidines, and nucleosides by paper chromatography. J. Biol. Chem. 1948, 175, 315–332. [Google Scholar]
  37. Johnson, T.B.; Coghill, R.D. researches on pyrimidines. C111. The discovery of 5-methyl-cytosine in tuberculinic acid, the nucleic acid of the tubercle bacillus1. J. Am. Chem. Soc. 1925, 47, 2838–2844. [Google Scholar] [CrossRef]
  38. Holliday, R.; Pugh, J. DNA modification mechanisms and gene activity during development. Science 1975, 187, 226–232. [Google Scholar] [CrossRef]
  39. Riggs, A. X inactivation, differentiation, and DNA methylation. Cytogenet. Genome Res. 1975, 14, 9–25. [Google Scholar] [CrossRef]
  40. Dong, A. Structure of human DNMT2, an enigmatic DNA methyltransferase homolog that displays denaturant-resistant binding to DNA. Nucleic Acids Res. 2001, 29, 439–448. [Google Scholar] [CrossRef] [Green Version]
  41. Mhanni, A.A.; Yoder, J.A.; Dubesky, C.; McGowan, R.A. Cloning and sequence analysis of a zebrafish cDNA encoding DNA (cytosine-5)-methyltransferase-1. Genesis 2001, 30, 213–219. [Google Scholar] [CrossRef] [PubMed]
  42. Shimoda, N.; Yamakoshi, K.; Miyake, A.; Takeda, H. Identification of a gene required for de novo DNA methylation of the zebrafish no tail gene. Dev. Dyn. 2005, 233, 1509–1516. [Google Scholar] [CrossRef] [PubMed]
  43. Okano, M.; Xie, S.; Li, E. Cloning and characterization of a family of novel mammalian DNA (cytosine-5) methyltransferases. Nat. Genet. 1998, 19, 219–220. [Google Scholar] [CrossRef] [PubMed]
  44. Cummings, D.J.; Tait, A.; Goddard, J.M. Methylated bases in DNA from Paramecium aurelia. Biochim. Biophys. Acta (BBA) 1974, 374, 1–11. [Google Scholar] [CrossRef]
  45. Hattman, S.; Kenny, C.; Berger, L.; Pratt, K. Comparative study of DNA methylation in three unicellular eucaryotes. J. Bacteriol. 1978, 135, 1156–1157. [Google Scholar] [CrossRef] [Green Version]
  46. Tomizawa, S.; Kobayashi, H.; Watanabe, T.; Andrews, S.; Hata, K.; Kelsey, G.; Sasaki, H. Dynamic stage-specific changes in imprinted differentially methylated regions during early mammalian development and prevalence of non-CpG methylation in oocytes. Development 2011, 138, 811–820. [Google Scholar] [CrossRef] [Green Version]
  47. Xie, W.; Barr, C.L.; Kim, A.; Yue, F.; Lee, A.Y.; Eubanks, J.; Dempster, E.L.; Ren, B. Base-Resolution Analyses of Sequence and Parent-of-Origin Dependent DNA Methylation in the Mouse Genome. Cell 2012, 148, 816–831. [Google Scholar] [CrossRef] [Green Version]
  48. Ramsahoye, B.H.; Biniszkiewicz, D.; Lyko, F.; Clark, V.; Bird, A.P.; Jaenisch, R. Non-CpG methylation is prevalent in embryonic stem cells and may be mediated by DNA methyltransferase 3a. Proc. Natl. Acad. Sci. USA 2000, 97, 5237–5242. [Google Scholar] [CrossRef] [Green Version]
  49. Bell, A.C.; Felsenfeld, G. Methylation of a CTCF-dependent boundary controls imprinted expression of the Igf2 gene. Nat. Cell Biol. 2000, 405, 482–485. [Google Scholar] [CrossRef]
  50. Lin, S.-Y.; Riggs, A.D. Lac Operator Analogues: Bromodeoxyuridine Substitution in the lac Operator Affects the Rate of Dissociation of the lac Repressor. Proc. Natl. Acad. Sci. USA 1972, 69, 2574–2576. [Google Scholar] [CrossRef] [Green Version]
  51. Lin, S.-Y.; Lin, D.; Riggs, A.D. Histones bind more tightly to bromodeoxyuridine-substituted DNA than to normal DNA. Nucleic Acids Res. 1976, 3, 2183–2192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Kallos, J.; Fasy, T.M.; Hollander, V.P.; Bick, M.D. Estrogen receptor has enhanced affinity for bromodeoxyuridine-substituted DNA. Proc. Natl. Acad. Sci. USA 1978, 75, 4896–4900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Hendrich, B.; Bird, A. Identification and characterization of a family of mammalian methyl-CpG binding proteins. Mol. Cell. Biol. 1998, 18, 6538–6547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Liu, Y.; Sun, L.; Jost, J.-P. In Differentiating Mouse Myoblasts DNA Methyltransferase Is Posttranscriptionally and Posttranslationally Regulated. Nucleic Acids Res. 1996, 24, 2718–2722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Pallafacchina, G.; François, S.; Regnault, B.; Czarny, B.; Dive, V.; Cumano, A.; Montarras, D.; Buckingham, M. An adult tissue-specific stem cell in its niche: A gene profiling analysis of in vivo quiescent and activated muscle satellite cells. Stem Cell Res. 2010, 4, 77–91. [Google Scholar] [CrossRef] [PubMed]
  56. Liu, L.; Cheung, T.H.; Charville, G.W.; Hurgo, B.M.C.; Leavitt, T.; Shih, J.; Brunet, A.; Rando, T.A. Chromatin Modifications as Determinants of Muscle Stem Cell Quiescence and Chronological Aging. Cell Rep. 2013, 4, 189–204. [Google Scholar] [CrossRef] [Green Version]
  57. Ryall, J.G.; Dell’Orso, S.; Derfoul, A.; Juan, A.; Zare, H.; Feng, X.; Clermont, D.; Koulnis, M.; Gutierrez-Cruz, G.; Fulco, M.; et al. The NAD+-Dependent SIRT1 Deacetylase Translates a Metabolic Switch into Regulatory Epigenetics in Skeletal Muscle Stem Cells. Cell Stem Cell 2015, 16, 171–183. [Google Scholar] [CrossRef] [Green Version]
  58. Constantinides, P.G.; Jones, P.A.; Gevers, W. Functional striated muscle cells from non-myoblast precursors following 5-azacytidine treatment. Nat. Cell Biol. 1977, 267, 364–366. [Google Scholar] [CrossRef]
  59. Szyf, M.; Rouleau, J.; Theberge, J.; Bozovic, V. Induction of myogenic differentiation by an expression vector encoding the DNA methyltransferase cDNA sequence in the antisense orientation. J. Biol. Chem. 1992, 267, 12831–12836. [Google Scholar]
  60. Scarpa, S.; Lucarelli, M.; Palitti, F.; Carotti, D.; Strom, R. Simultaneous myogenin expression and overall DNA hypomethylation promote in vitro myoblast differentiation. Cell Growth Differ. Mol. Boil. J. Am. Assoc. Cancer Res. 1996, 7, 1051–1058. [Google Scholar]
  61. Podgorniak, T.; Brockmann, S.; Konstantinidis, I.; Fernandes, J.M.O. Differences in the fast muscle methylome provide insight into sex-specific epigenetic regulation of growth in Nile tilapia during early stages of domestication. Epigenetics 2019, 14, 818–836. [Google Scholar] [CrossRef] [PubMed]
  62. Biga, P.R.; Latimer, M.N.; Froehlich, J.M.; Gabillard, J.-C.; Seiliez, I. Distribution of H3K27me3, H3K9me3, and H3K4me3 along autophagy-related genes highly expressed in starved zebrafish myotubes. Biol. Open 2017, 6, 1720–1725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Craig, P.M.; Moon, T.W. Methionine restriction affects the phenotypic and transcriptional response of rainbow trout (Oncorhynchus mykiss) to carbohydrate-enriched diets. Br. J. Nutr. 2012, 109, 402–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Canada, P.; Engrola, S.; Conceição, L.E.; Valente, L.M. Improving growth potential in Senegalese sole (Solea senegalensis) through dietary protein. Aquaculture 2019, 498, 90–99. [Google Scholar] [CrossRef]
  65. Kumkhong, S.; Marandel, L.; Plagnes-Juan, E.; Veron, V.; Boonanuntanasarn, S.; Panserat, S. Glucose Injection Into Yolk Positively Modulates Intermediary Metabolism and Growth Performance in Juvenile Nile Tilapia (Oreochromis niloticus). Front. Physiol. 2020, 11, 286. [Google Scholar] [CrossRef]
  66. Burgerhout, E.; Mommens, M.; Johnsen, H.; Aunsmo, A.; Santi, N.; Andersen, Ø. Genetic background and embryonic temperature affect DNA methylation and expression of myogenin and muscle development in Atlantic salmon (Salmo solar). PLoS ONE 2017, 12, e0179918. [Google Scholar] [CrossRef] [Green Version]
  67. Anastasiadi, D.; Díaz, N.; Piferrer, F. Small ocean temperature increases elicit stage-dependent changes in DNA methylation and gene expression in a fish, the European sea bass. Sci. Rep. 2017, 7, 1–12. [Google Scholar] [CrossRef] [Green Version]
  68. Campos, C.; Valente, L.M.; Conceição, L.E.; Engrola, S.; Sousa, V.; Rocha, E.; Fernandes, J.M.O. Incubation temperature induces changes in muscle cellularity and gene expression in Senegalese sole (Solea senegalensis). Gene 2013, 516, 209–217. [Google Scholar] [CrossRef]
  69. Campos, C.; Valente, L.M.P.; Conceição, L.E.; Engrola, S.; Fernandes, J.M.O. Temperature affects methylation of the myogenin putative promoter, its expression and muscle cellularity in Senegalese sole larvae. Epigenetics 2013, 8, 389–397. [Google Scholar] [CrossRef] [Green Version]
  70. Giannetto, A.; Nagasawa, K.; Fasulo, S.; Fernandes, J.M.O. Influence of photoperiod on expression of DNA (cytosine-5) methyltransferases in Atlantic cod. Gene 2013, 519, 222–230. [Google Scholar] [CrossRef]
  71. Koganti, P.P.; Wang, J.; Cleveland, B.; Yao, J. 17 beta-Estradiol Increases Non-CpG Methylation in Exon 1 of the Rainbow Trout (Oncorhynchus mykiss) MyoD Gene. Mar. Biotechnol. 2017, 19, 321–327. [Google Scholar] [CrossRef] [PubMed]
  72. Simó-Mirabet, P.; Perera, E.; Calduch-Giner, J.A.; Pérez-Sánchez, J. Local DNA methylation helps to regulate muscle sirtuin 1 gene expression across seasons and advancing age in gilthead sea bream (Sparus aurata). Front. Zool. 2020, 17, 1–18. [Google Scholar] [CrossRef] [PubMed]
  73. Fuentes, E.N.; Zuloaga, R.; Nardocci, G.; De La Reguera, C.F.; Simonet, N.G.; Fumeron, R.; Valdés, J.A.; Molina, A.; Alvarez, M. Skeletal muscle plasticity induced by seasonal acclimatization in carp involves differential expression of rRNA and molecules that epigenetically regulate its synthesis. Comp. Biochem. Physiol. Part B Biochem. Mol. Biol. 2014, 172, 57–66. [Google Scholar] [CrossRef] [PubMed]
  74. Seiliez, I.; Froehlich, J.M.; Marandel, L.; Gabillard, J.-C.; Biga, P.R. Evolutionary history and epigenetic regulation of the three paralogous pax7 genes in rainbow trout. Cell Tissue Res. 2014, 359, 715–727. [Google Scholar] [CrossRef] [Green Version]
  75. Nilsen, T.W. Mechanisms of microRNA-mediated gene regulation in animal cells. Trends Genet. 2007, 23, 243–249. [Google Scholar] [CrossRef]
  76. Hwang, H.-W.; Wentzel, E.A.; Mendell, J.T. A Hexanucleotide Element Directs MicroRNA Nuclear Import. Science 2007, 315, 97–100. [Google Scholar] [CrossRef]
  77. Liao, J.-Y.; Ma, L.-M.; Guo, Y.-H.; Zhang, Y.-C.; Zhou, H.; Shao, P.; Chen, Y.-Q.; Qu, L.-H. Deep Sequencing of Human Nuclear and Cytoplasmic Small RNAs Reveals an Unexpectedly Complex Subcellular Distribution of miRNAs and tRNA 3′ Trailers. PLoS ONE 2010, 5, e10563. [Google Scholar] [CrossRef] [Green Version]
  78. Valadi, H.; Ekström, K.; Bossios, A.; Sjöstrand, M.; Lee, J.J.; Tvall, J.O.L.O. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat. Cell Biol. 2007, 9, 654–659. [Google Scholar] [CrossRef] [Green Version]
  79. Gallo, A.; Tandon, M.; Alevizos, I.; Illei, G.G. The Majority of MicroRNAs Detectable in Serum and Saliva Is Concentrated in Exosomes. PLoS ONE 2012, 7, e30679. [Google Scholar] [CrossRef] [Green Version]
  80. Lv, L.-L.; Cao, Y.; Liu, D.; Xu, M.; Liu, H.; Tang, R.-N.; Ma, K.-L.; Liu, B.-C. Isolation and Quantification of MicroRNAs from Urinary Exosomes/Microvesicles for Biomarker Discovery. Int. J. Biol. Sci. 2013, 9, 1021–1031. [Google Scholar] [CrossRef]
  81. Zhou, R.; Wu, Y.; Tao, M.; Zhang, C.; Liu, S. MicroRNA profiles reveal female allotetraploid hybrid fertility. BMC Genet. 2015, 16, 119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Kosaka, N.; Iguchi, H.; Yoshioka, Y.; Takeshita, F.; Matsuki, Y.; Ochiya, T. Secretory mechanisms and intercellular transfer of MicroRNAs in living cells. J. Biol. Chem. 2010, 285, 17442–17452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Kim, V.N.; Han, J.; Siomi, M.C. Biogenesis of small RNAs in animals. Nat. Rev. Mol. Cell Biol. 2009, 10, 126–139. [Google Scholar] [CrossRef] [PubMed]
  84. O’Rourke, J.R.; Georges, S.A.; Seay, H.R.; Tapscott, S.J.; McManus, M.T.; Goldhamer, D.J.; Swanson, M.S.; Harfe, B.D. Essential role for Dicer during skeletal muscle development. Dev. Biol. 2007, 311, 359–368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Ge, Y.; Chen, J. MicroRNAs in skeletal myogenesis. Cell Cycle 2011, 10, 441–448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Horak, M.; Novák, J.; Bienertova-Vasku, J. Muscle-specific microRNAs in skeletal muscle development. Dev. Biol. 2016, 410, 1–13. [Google Scholar] [CrossRef] [PubMed]
  87. Rao, P.K.; Kumar, R.M.; Farkhondeh, M.; Baskerville, S.; Lodish, H.F. Myogenic factors that regulate expression of muscle-specific microRNAs. Proc. Natl. Acad. Sci. USA 2006, 103, 8721–8726. [Google Scholar] [CrossRef] [Green Version]
  88. Nachtigall, P.G.; Dias, M.C.; Carvalho, R.F.; Martins, C.; Pinhal, D. MicroRNA-499 Expression Distinctively Correlates to Target Genes sox6 and rod1 Profiles to Resolve the Skeletal Muscle Phenotype in Nile Tilapia. PLoS ONE 2015, 10, e0119804. [Google Scholar] [CrossRef] [Green Version]
  89. Salem, M.; Xiao, C.; Womack, J.; Rexroad, C.E.; Yao, J. A MicroRNA Repertoire for Functional Genome Research in Rainbow Trout (Oncorhynchus mykiss). Mar. Biotechnol. 2009, 12, 410–429. [Google Scholar] [CrossRef]
  90. Koganti, P.P.; Wang, J.; Cleveland, B.; Ma, H.; Weber, G.M.; Yao, J. Estradiol regulates expression of miRNAs associated with myogenesis in rainbow trout. Mol. Cell. Endocrinol. 2017, 443, 1–14. [Google Scholar] [CrossRef] [Green Version]
  91. Mishima, Y.; Abreu-Goodger, C.; Staton, A.A.; Stahlhut, C.; Shou, C.; Cheng, C.; Gerstein, M.; Enright, A.J.; Giraldez, A.J. Zebrafish miR-1 and miR-133 shape muscle gene expression and regulate sarcomeric actin organization. Genes Dev. 2009, 23, 619–632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Fu, Y.; Shi, Z.; Wu, M.; Zhang, J.; Jia, L.; Chen, X. Identification and Differential Expression of MicroRNAs during Metamorphosis of the Japanese Flounder (Paralichthys olivaceus). PLoS ONE 2011, 6, e22957. [Google Scholar] [CrossRef]
  93. Xia, J.H.; He, X.P.; Bai, Z.Y.; Yue, G.H. Identification and Characterization of 63 MicroRNAs in the Asian Seabass Lates calcarifer. PLoS ONE 2011, 6, e17537. [Google Scholar] [CrossRef] [PubMed]
  94. Huang, C.W.; Li, Y.H.; Hu, S.Y.; Chi, J.R.; Lin, G.H.; Lin, C.C.; Gong, H.-Y.; Chen, J.-Y.; Chen, R.H.; Chang, S.J.; et al. Differential expression patterns of growth-related microRNAs in the skeletal muscle of Nile tilapia (Oreochromis niloticus)1. J. Anim. Sci. 2012, 90, 4266–4279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Chu, W.-Y.; Liu, L.-S.; Li, Y.-L.; Chen, L.; Wang, K.-Z.; Li, H.-H.; Du, S.-J.; Zhang, J.-S. Systematic Identification and Differential Expression Profiling of MicroRNAs from White and Red Muscles of Siniperca chuatsi. Curr. Mol. Med. 2013, 13, 1397–1407. [Google Scholar] [CrossRef] [PubMed]
  96. Yan, X.; Ding, L.; Li, Y.; Zhang, X.; Liang, Y.; Sun, X.; Teng, C.-B. Identification and Profiling of MicroRNAs from Skeletal Muscle of the Common Carp. PLoS ONE 2012, 7, e30925. [Google Scholar] [CrossRef] [PubMed]
  97. Chen, J.-F.; Mandel, E.M.; Thomson, J.M.; Wu, Q.; Callis, E.T.; Hammond, S.M.; Conlon, F.L.; Wang, D.-Z. The Role of MicroRNA-1 and MicroRNA-133 in Skeletal Muscle Proliferation and Differentiation. Nat. Genet. 2005, 38, 228–233. [Google Scholar] [CrossRef]
  98. Chen, J.-F.; Tao, Y.; Li, J.; Deng, Z.; Yan, Z.; Xiao, X.; Wang, D.-Z. microRNA-1 and microRNA-206 regulate skeletal muscle satellite cell proliferation and differentiation by repressing Pax7. J. Cell Biol. 2010, 190, 867–879. [Google Scholar] [CrossRef] [Green Version]
  99. Rosenberg, M.I.; Georges, S.A.; Asawachaicharn, A.; Analau, E.; Tapscott, S.J. MyoD inhibits Fstl1 and Utrn expression by inducing transcription of miR-206. J. Cell Biol. 2006, 175, 77–85. [Google Scholar] [CrossRef] [Green Version]
  100. Yan, B.; Zhu, C.-D.; Guo, J.; Zhao, L.-H.; Zhao, J. miR-206 regulates the growth of the teleost tilapia (Oreochromis niloticus) through the modulation of IGF-1 gene expression. J. Exp. Biol. 2012, 216, 1265–1269. [Google Scholar] [CrossRef] [Green Version]
  101. Yan, B.; Guo, J.; Zhu, C.-D.; Zhao, L.-H.; Zhao, J. miR-203b: A novel regulator of MyoD expression in tilapia skeletal muscle. J. Exp. Biol. 2012, 216, 447–451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Chen, L.; Wu, P.; Guo, X.-H.; Hu, Y.; Li, Y.-L.; Shi, J.; Wang, K.-Z.; Chu, W.-Y.; Zhang, J. miR-143: A novel regulator of MyoD expression in fast and slow muscles of Siniperca chuatsi. Curr. Mol. Med. 2014, 14, 370–375. [Google Scholar] [CrossRef] [PubMed]
  103. Zhao, Z.; Yu, X.; Jia, J.; Yang, G.; Sun, C.; Li, W. miR-181b-5p May Regulate Muscle Growth in Tilapia by Targeting Myostatin b. Front. Endocrinol. 2019, 10, 812. [Google Scholar] [CrossRef] [PubMed]
  104. Lou, Z.; Zhao, Y.; Zhang, Y.; Zheng, B.; Feng, H.; Hosain, M.A.; Xue, L. MiR-2014-5p and miR-1231-5p regulate muscle growth of Larimichthys crocea by targeting MSTN gene. Comp. Biochem. Physiol. Part B Biochem. Mol. Biol. 2020, 252, 110535. [Google Scholar] [CrossRef]
  105. Zhang, R.; Li, R.; Lin, Y. Identification and characterization of microRNAs in the muscle of Schizothorax prenanti. Fish Physiol. Biochem. 2017, 121, 207–1064. [Google Scholar] [CrossRef]
  106. Duran, B.O.D.S.; Fernandez, G.J.; Mareco, E.A.; Moraes, L.N.; Salomão, R.A.S.; De Paula, T.G.; Santos, V.B.; Carvalho, R.F.; Dal-Pai-Silva, M. Differential microRNA Expression in Fast- and Slow-Twitch Skeletal Muscle of Piaractus mesopotamicus during Growth. PLoS ONE 2015, 10, e0141967. [Google Scholar] [CrossRef]
  107. Paneru, B.; Al-Tobasei, R.; Kenney, B.; Leeds, T.D.; Salem, M. RNA-Seq reveals MicroRNA expression signature and genetic polymorphism associated with growth and muscle quality traits in rainbow trout. Sci. Rep. 2017, 7, 1–15. [Google Scholar] [CrossRef]
  108. Tu, J.; Tian, C.; Zhao, P.; Sun, J.; Wang, M.; Fan, Q.; Yuan, Y.C. Identification and profiling of growth-related microRNAs in Chinese perch (Siniperca chuatsi). BMC Genom. 2017, 18, 489. [Google Scholar] [CrossRef] [Green Version]
  109. Yi, S.; Gao, Z.; Zhao, H.-H.; Zeng, C.; Luo, W.; Chen, B.-X.; Wang, W. Identification and characterization of microRNAs involved in growth of blunt snout bream (Megalobrama amblycephala) by Solexa sequencing. BMC Genom. 2013, 14, 754. [Google Scholar] [CrossRef] [Green Version]
  110. Johnston, I.A.; Lee, H.-T.; MacQueen, D.J.; Paranthaman, K.; Kawashima, C.; Anwar, A.; Kinghorn, J.R.; Dalmay, T. Embryonic temperature affects muscle fibre recruitment in adult zebrafish: Genome-wide changes in gene and microRNA expression associated with the transition from hyperplastic to hypertrophic growth phenotypes. J. Exp. Biol. 2009, 212, 1781–1793. [Google Scholar] [CrossRef] [Green Version]
  111. Campos, C.; Valente, L.M.; Conceição, L.E.; Engrola, S.; Fernandes, J.M. Molecular regulation of muscle development and growth in Senegalese sole larvae exposed to temperature fluctuations. Aquaculture 2014, 432, 418–425. [Google Scholar] [CrossRef] [Green Version]
  112. Zhu, X.; Chen, D.; Hu, Y.; Wu, P.; Wang, K.; Zhang, J.; Chu, W.; Zhang, J. The microRNA Signature in Response to Nutrient Restriction and Refeeding in Skeletal Muscle of Chinese Perch (Siniperca chuatsi). Mar. Biotechnol. 2014, 17, 180–189. [Google Scholar] [CrossRef] [PubMed]
  113. Zhu, X.; Chu, W.-Y.; Wu, P.; Yi, T.; Chen, T.; Zhang, J. MicroRNA signature in response to nutrient restriction and re-feeding in fast skeletal muscle of grass carp (Ctenopharyngodon idella). Dongwuxue Yanjiu 2014, 35, 404–410. [Google Scholar] [PubMed]
  114. Latimer, M.; Sabin, N.; Le Cam, A.; Seiliez, I.; Biga, P.; Gabillard, J.-C. miR-210 expression is associated with methionine-induced differentiation of trout satellite cells. J. Exp. Biol. 2017, 220, 2932–2938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Latimer, M.; Cleveland, B.; Biga, P.R. Dietary methionine restriction: Effects on glucose tolerance, lipid content and micro-RNA composition in the muscle of rainbow trout. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2018, 208, 47–52. [Google Scholar] [CrossRef] [PubMed]
  116. Li, H.; An, X.; Bao, L.; Li, Y.; Pan, Y.; He, J.; Liu, L.; Zhu, X.; Zhang, J.; Cheng, J.; et al. MiR-125a-3p-KLF15-BCAA Regulates the Skeletal Muscle Branched-Chain Amino Acid Metabolism in Nile Tilapia (Oreochromis niloticus) During Starvation. Front. Genet. 2020, 11, 852. [Google Scholar] [CrossRef]
  117. Mennigen, J.A.; Skiba-Cassy, S.; Panserat, S. Ontogenetic expression of metabolic genes and microRNAs in rainbow trout alevins during the transition from the endogenous to the exogenous feeding period. J. Exp. Biol. 2013, 216, 1597–1608. [Google Scholar] [CrossRef] [Green Version]
  118. Bizuayehu, T.T.; Furmanek, T.; Karlsen, Ø.; Van Der Meeren, T.; Edvardsen, R.B.; Rønnestad, I.; Hamre, K.; Johansen, S.D.; Babiak, I. First feed affects the expressions of microRNA and their targets in Atlantic cod. Br. J. Nutr. 2016, 115, 1145–1154. [Google Scholar] [CrossRef] [Green Version]
  119. Paneru, B.; Ali, A.; Al-Tobasei, R.; Kenney, B.; Salem, M. Crosstalk among lncRNAs, microRNAs and mRNAs in the muscle ‘degradome’ of rainbow trout. Sci. Rep. 2018, 8, 1–15. [Google Scholar] [CrossRef] [Green Version]
  120. Flynt, A.S.; Li, N.; Thatcher, E.J.; Solnica-Krezel, L.; Patton, J.G. Zebrafish miR-214 modulates Hedgehog signaling to specify muscle cell fate. Nat. Genet. 2007, 39, 259–263. [Google Scholar] [CrossRef] [Green Version]
  121. Lin, C.-Y.; Chen, J.-S.; Loo, M.-R.; Hsiao, C.-C.; Chang, W.-Y.; Tsai, H.-J. MicroRNA-3906 Regulates Fast Muscle Differentiation through Modulating the Target Gene homer-1b in Zebrafish Embryos. PLoS ONE 2013, 8, e70187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Duran, B.O.D.S.; Dal-Pai-Silva, M.; De La Serrana, D.G. Rainbow trout slow myoblast cell culture as a model to study slow skeletal muscle, and the characterization of mir-133 and mir-499 families as a case study. J. Exp. Biol. 2020, 223. [Google Scholar] [CrossRef] [PubMed]
  123. Wang, X.; Ono, Y.; Tan, S.C.; Chai, R.J.; Parkin, C.; Ingham, P.W. Prdm1a and miR-499 act sequentially to restrict Sox6 activity to the fast-twitch muscle lineage in the zebrafish embryo. Development 2011, 138, 4399–4404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Chu, W.; Zhang, F.; Song, R.; Li, Y.; Wu, P.; Chen, L.; Cheng, J.; Du, S.; Zhang, J. Proteomic and microRNA Transcriptome Analysis revealed the microRNA-SmyD1 network regulation in Skeletal Muscle Fibers performance of Chinese perch. Sci. Rep. 2017, 7, 16498. [Google Scholar] [CrossRef] [PubMed]
  125. Campos, C.; Sundaram, A.Y.M.; Valente, L.M.P.; Conceição, L.E.; Engrola, S.; Fernandes, J.M.O. Thermal plasticity of the miRNA transcriptome during Senegalese sole development. BMC Genom. 2014, 15, 525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Ando, S.; Hatano, M.; Zama, K. Protein degradation and protease activity of chum salmon (Oncorhynchus keta) muscle during spawning migration. Fish Physiol. Biochem. 1986, 1, 17–26. [Google Scholar] [CrossRef] [PubMed]
  127. Cleveland, B.M.; Kenney, P.B.; Manor, M.L.; Weber, G.M. Effects of feeding level and sexual maturation on carcass and fillet characteristics and indices of protein degradation in rainbow trout (Oncorhynchus mykiss). Aquaculture 2012, 338, 228–236. [Google Scholar] [CrossRef]
Table 1. Established roles of DNA methylation for myogenic regulatory factors (MRF).
Table 1. Established roles of DNA methylation for myogenic regulatory factors (MRF).
Gene/MRFFunctional RoleMethylation StatusReference
Pax7Migration and early lineage commitmentHypermethylation in myogenic cells and mature muscle fibers[28]
Myf5Proliferation and differentiation of MPC into myoblastsHypermethylation of enhancer region in embryonic stem cells
Hypomethylation in myoblasts, myotubes, and skeletal muscle
[29]
MyodProliferation and differentiation of MPC into myoblastsHypomethylation in distal enhancer region[30]
MyogeninDifferentiation of myoblasts into myotubesDemethylation in differentiated muscle
Hypermethylated in myoblasts and non-myogenic cells
[31]
ObsnFormation of skeletal muscleHypomethylation in muscle tissue
Hypermethylated in myoblasts and myotubes
[28]
Myh7bExpressed intronic microRNA miR499Hypomethylation[28]
Gene promotersMyotube formationHypermethylation of ID4 and ZNF238 binding sites[32]
Notch1Proliferation of muscle satellite cellsHypomethylation and its ligands Dll1 and Jag2 in skeletal lineage cells[33]
Table 2. Research in fish investigating epigenetic regulation of muscle plasticity.
Table 2. Research in fish investigating epigenetic regulation of muscle plasticity.
Factor Affecting Muscle PlasticityFish Specie(s)Reference
Genetic variation in growthNile tilapia[61]
TemperatureAtlantic salmon,
European sea bass, Senegalese sole,
stickleback
[66,67,68,69]
Nutritionzebrafish,
rainbow trout,
Senegalese sole
[62,63,64,65]
PhotoperiodAtlantic cod[70]
17β-estradiolrainbow trout[71]
Seasonal acclimationcommon carp
gilthead sea bream
[72,73]
Table 3. Research in fish supporting miRNA-mediated regulation of muscle plasticity.
Table 3. Research in fish supporting miRNA-mediated regulation of muscle plasticity.
Biological Factors Affecting Muscle PlasticityFish Specie(s)Reference
TemperatureSenegalese sole
zebrafish
[110,111]
Genetic variation in growthrainbow trout
tilapia
Chinese perch,
blunt snout sea bream
[94,107,108,109]
Nutritionrainbow trout,
Chinese perch,
grass carp,
Nile tilapia,
Atlantic cod
[112,113,114,115,116,117,118]
Spawning or 17β-estradiolrainbow trout[90,119]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Koganti, P.; Yao, J.; Cleveland, B.M. Molecular Mechanisms Regulating Muscle Plasticity in Fish. Animals 2021, 11, 61. https://0-doi-org.brum.beds.ac.uk/10.3390/ani11010061

AMA Style

Koganti P, Yao J, Cleveland BM. Molecular Mechanisms Regulating Muscle Plasticity in Fish. Animals. 2021; 11(1):61. https://0-doi-org.brum.beds.ac.uk/10.3390/ani11010061

Chicago/Turabian Style

Koganti, Prasanthi, Jianbo Yao, and Beth M. Cleveland. 2021. "Molecular Mechanisms Regulating Muscle Plasticity in Fish" Animals 11, no. 1: 61. https://0-doi-org.brum.beds.ac.uk/10.3390/ani11010061

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop