Next Article in Journal
Stress-Induced Changes in Nucleocytoplasmic Localization of Crucial Factors in Gene Expression Regulation
Next Article in Special Issue
Genetic Structure and Diversity of Hatchery and Wild Populations of Yellow Catfish Tachysurus fulvidraco (Siluriformes: Bagridae) from Korea
Previous Article in Journal
Insights into the Adolescent Cystic Fibrosis Airway Microbiome Using Shotgun Metagenomics
Previous Article in Special Issue
Theoretical Analysis and Expression Profiling of 17β-Hydroxysteroid Dehydrogenase Genes in Gonadal Development and Steroidogenesis of Leopard Coral Grouper (Plectropomus leopardus)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring the Integrated Role of miRNAs and lncRNAs in Regulating the Transcriptional Response to Amino Acids and Insulin-like Growth Factor 1 in Gilthead Sea Bream (Sparus aurata) Myoblasts

by
Isabel García-Pérez
1,
Bruno Oliveira Silva Duran
2,
Maeli Dal-Pai-Silva
3 and
Daniel Garcia de la serrana
1,*
1
Department of Cell Biology, Physiology and Immunology, Faculty of Biology, University of Barcelona (UB), 08028 Barcelona, Spain
2
Department of Histology, Embryology and Cell Biology, Institute of Biological Sciences, Federal University of Goiás (UFG), Goiânia 74690-900, Brazil
3
Department of Structural and Functional Biology, Institute of Biosciences, São Paulo State University (UNESP), Botucatu 18618-689, Brazil
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(7), 3894; https://doi.org/10.3390/ijms25073894
Submission received: 23 December 2023 / Revised: 26 March 2024 / Accepted: 26 March 2024 / Published: 31 March 2024
(This article belongs to the Special Issue Fish Genomics and Developmental Biology)

Abstract

:
In this study, gilthead sea bream (Sparus aurata) fast muscle myoblasts were stimulated with two pro-growth treatments, amino acids (AA) and insulin-like growth factor 1 (Igf-1), to analyze the transcriptional response of mRNAs, microRNAs (miRNAs) and long non-coding RNAs (lncRNAs) and to explore their possible regulatory network using bioinformatic approaches. AA had a higher impact on transcription (1795 mRNAs changed) compared to Igf-1 (385 mRNAs changed). Both treatments stimulated the transcription of mRNAs related to muscle differentiation (GO:0042692) and sarcomere (GO:0030017), while AA strongly stimulated DNA replication and cell division (GO:0007049). Both pro-growth treatments altered the transcription of over 100 miRNAs, including muscle-specific miRNAs (myomiRs), such as miR-133a/b, miR-206, miR-499, miR-1, and miR-27a. Among 111 detected lncRNAs (>1 FPKM), only 30 were significantly changed by AA and 11 by Igf-1. Eight lncRNAs exhibited strong negative correlations with several mRNAs, suggesting a possible regulation, while 30 lncRNAs showed strong correlations and interactions with several miRNAs, suggesting a role as sponges. This work is the first step in the identification of the ncRNAs network controlling muscle development and growth in gilthead sea bream, pointing out potential regulatory mechanisms in response to pro-growth signals.

1. Introduction

The skeletal muscle of teleost fish is a very plastic tissue that integrates external and internal inputs to adapt to a changing environment. Most teleost can form new muscle fibers (hyperplasia) long after the end of metamorphosis well into adulthood (until about 44% of the maximal length of the species), overlapping with the growth of pre-existent fibers (hypertrophy) [1,2,3]. Muscle growth includes the activation of the satellite cells, their proliferation, fusion, differentiation, and maturation in a process known as myogenesis. The molecular regulation of myogenesis involves the coordinated work of transcription factors, growth factors, activation of different pathways, and fusion proteins [2,4,5,6,7,8].
Furthermore, muscle growth and development also depend on the balance between protein synthesis and degradation, which are controlled by various cellular signaling pathways, including the nutrient-sensitive mechanistic target of rapamycin (mTOR) network. mTOR is a protein kinase that acts as a central regulator of cellular metabolism, proliferation, and growth and is activated in response to various signals, including nutrients (e.g., amino acids (AA)) and growth factors (e.g., insulin-like growth factor 1 (Igf-1)). When AA are present in sufficient quantities, they activate mTOR at the lysosome membrane [9,10,11,12], which then stimulates protein synthesis by phosphorylating and activating downstream targets such as P70 ribosomal S6 kinase (P70S6K) and eukaryotic translation initiation factor 4e binding protein 1 (4EBP1). Studies with fish muscle showed that AA reduce transcription of muscle-specific ubiquitin ligases [13] and autophagy-related genes [14], leading to reduced protein breakdown [15,16]. Likewise, Igf-1 also activates mTOR throughout the induction of phosphatidylinositol 3-kinase (PI3K)/protein kinase B (AKT) pathway. It has been shown in different fish species and experimental setups that Igfs increase muscle growth by promoting myoblast proliferation and differentiation [17,18,19,20]. Thus, given the importance of the Igf system and AA in promoting protein synthesis and in the myogenesis process, studying their effects would help to better understand the potential regulation of muscle growth and development in fish, including aquaculture species like the gilthead sea bream (Sparus aurata).
In the last decades, research in mammalian models has demonstrated that the non-coding RNAs (ncRNAs) also play a key role in regulating myogenesis [21,22,23], but little is known about their role in fish muscle development [24,25,26,27]. The ncRNAs are a group of RNAs that, generally, do not codify for proteins but perform various regulatory functions in cellular processes. The ncRNAs include ribosomal RNAs (rRNAs), transfer RNAs (tRNAs), small nuclear RNAs (snRNAs), small nucleolar RNAs (snoRNA), microRNAs (miRNAs), small interfering RNA (siRNAs), piwi-interacting RNAs (piRNAs), circular RNAs (circRNAs) and long non-coding RNAs (lncRNAs). The miRNAs regulate gene expression by recognition of the complementary sequence present in the target mRNAs. When an mRNA is recognized by a specific miRNA, its degeneration, deadenylation, or impaired translation into a protein can be triggered, usually leading to a negative correlation between the expression of miRNAs and their target mRNAs [28,29,30]. Therefore, the miRNAs expand the complexity of transcriptomic regulation and are key players in the control of cellular functions [31]. Many miRNAs are ubiquitously expressed in most cell types and tissues, but some are tissue-specific like the myomiRs, muscle enriched, or striated muscle-specific miRNAs. MyomiRs are involved in myoblast proliferation, differentiation, or muscle regeneration, and each one is expressed differently throughout the myogenesis process [22,32,33]. It has been shown in C2C12 myoblasts that miR-1 and miR-206 promote myogenic differentiation by repressing the expression of Pax7 and histone deacetylase 4 (HDAC4) [34,35,36,37]; and also that miR-206 is involved in muscle regeneration and it is markedly upregulated in satellite cells following muscle injury [38,39]. On the other hand, miR-133a is known to have an important role in muscle cell proliferation, repressing serum response factor (SRF) [40]. However, miR-133b also participates in the promotion of myoblast differentiation and fusion [41,42]. In the case of miR-499 and miR-208b, they are involved in the specification and maintenance of slow-twitch phenotype [43,44,45]. The roles of these miRNAs were also investigated in fish skeletal muscle, with miR-1/206 and miR-133 families regulating myogenesis and development [46,47,48], sarcomeric organization [49], and protein balance [50,51,52]; and miR-499 inducing the establishment and maintenance of slow-twitch muscle fibers [48,53,54].
On the other hand, lncRNAs can increase or decrease the transcription and function of genes through different strategies, such as direct interaction with the DNA, RNA, or even proteins. Some lncRNAs can interact with the DNA and change the chromatin structure, modulating the access of transcription factors to the gene promotors or allowing the physical proximity to enhancers [55,56,57]. In addition, the lncRNAs can also interact directly with mRNAs, showing opposite functions that could induce mRNA degradation or prevent it by acting as sponges of miRNAs [26,58], or even regulate the gene function by direct interaction with proteins [59]. The number of lncRNAs characterized in human skeletal muscle has increased in recent years and now includes lncRNAs such as Neat1 [60], MUNC [61,62], linc-RAM [63], Irm [64], or H19 [65]. Recent research in fish has indicated that lncRNAs participate in many biological processes, including lipid metabolism [66], intestinal homeostasis [67], immune response [68], sex differentiation [69], and the smoltification process [70]. However, our knowledge of lncRNAs in fish skeletal muscle is very limited [24,26,27,71]. One of the major problems is the apparent low conservation of lncRNAs [72], which makes it very difficult to identify relevant lncRNAs in species other than humans, having to start from scratch the work in different species.
To our knowledge, the role of miRNAs and lncRNAs in regulating the transcriptional response of fish skeletal muscle to pro-growth signals such as AA and Igf-1 has not yet been investigated in fish. Hence, this work uses an RNA-Seq approach to address the present lack of knowledge by generating a transcriptome and microRNAome from gilthead sea bream myoblasts stimulated with AA or Igf-1, and study the interactions between mRNAs, miRNAs, and lncRNAs to better understand the role of ncRNAs in the myoblast’s transcriptional response to pro-growth signals.

2. Results

2.1. Identification of miRNAs and lncRNAs in Gilthead Sea Bream Myoblasts

Myoblasts from gilthead sea bream fast skeletal muscle were extracted and seeded on 6-well culture plates at a density of 1.5 × 106 cells/well and let to develop for 8 days (Figure 1). At day 8, myoblasts were still proliferating, but a significant proportion of them started to fuse and form myotubes, allowing us to investigate miRNAs and lncRNAs present during proliferation and differentiation in response to pro-growth signals. We detected 403 miRNAs expressed in the gilthead sea bream myoblasts, with 8.58% showing a high expression (over 10,000 normalized reads), 20% showing low expression (under 10 normalized reads), and 70% showing intermediate expression (between 10 and 10,000 normalized reads) (Figure 2A). We also identified more than 870 lncRNAs (over 0.001 FPKM), but only 111 had over 1 FPKM average expression, while 25 had over 10 FPKM (Figure 2B). It is interesting to notice that in both lncRNAs and miRNAs the transcriptomic landscape is dominated by a few of them (Figure 2). For instance, four miRNAs (miR-21, miR-146, miR-22, and miR-206) were found to have over 500,000 normalized reads (Figure 2A; Supplementary File S1). Other miRNAs known to be important in mammalian skeletal muscle (miR-26a, miR-27, miR-133a/b, miR-221/222, miR-1, or miR-499) were also relatively abundant but not at the same level (Figure 2A; Supplementary File S1). In the case of lncRNAs, one of them, ENSSAUG00010015132, showed ten times more expression (>900 average FPKM) than the second more expressed lncRNA (ENSSAUG00010022378; >80 average FPKM), which rapidly decreased to very low levels of expression for the others lncRNAs (Figure 2B; Supplementary File S1). The majority of lncRNAs were predicted to be either located in the cytoplasm (70%) or nucleus (28%) (Supplementary File S1). A BLAST search of the lncRNAs > 1 FPKM from gilthead sea bream against the human and mouse genome did not show any significant ortholog.

2.2. Transcriptomic Changes of mRNAs in Response to AA and Igf-1

To determine the effects of the treatments, a principal components analysis (PCA) was performed. The PCA analysis showed that the samples from each condition clustered together in three distinct groups. It is interesting to notice that the Igf-1 cluster was closer to the CTR cluster than the AA cluster, suggesting that the global transcriptomic profile of the myoblasts treated with Igf-1 was more similar to the CTR profile than to that of AA (Figure 3). Also, the replicates of the Igf-1 and CTR groups were closer to each other compared to the AA groups, indicating lower variability in the response to the treatments (Figure 3).
The transcriptional response of the gilthead sea bream myoblasts to AA was more intense than the response to only Igf-1 (Supplementary File S2). In response to AA, we found a total of 1184 upregulated and 611 downregulated mRNAs compared to the CTR myoblasts (Figure 4 and Figure 5). When Igf-1 was added, only 253 genes were upregulated and 132 downregulated compared to CTR myoblasts (Figure 4 and Figure 5). We also found 182 and 92 genes commonly upregulated and downregulated in response to AA and Igf-1, respectively (Figure 5).
The Gene Ontology analysis of the up and downregulated genes in response to the different treatments showed differences between the processes affected and their intensity. Several GO terms related to DNA replication and cell cycle (0007049; 0006260; 0003688), muscle differentiation (0042692; 0003012), and sarcomere and muscle cytoskeleton (0007010; 0045214; 0008092; 0043292; 0030017) were upregulated in response to AA; while GO terms such as transport activity (0034219; 0015293) or growth factor and cytokine activity (0008083; 0005125) were downregulated in this condition (Table 1). The addition of Igf-1 increased the expression of genes related to muscle development (0042692; 0055001; 0061061) and muscle cytoskeleton (0030016; 0030017; 0015629) (Table 1). Some GO terms were shared between AA and Igf-1, but the number of genes involved was significantly different, with many more genes modified by AA (Table 1; Figure 6).

2.3. Transcriptomic Analysis of ncRNAs

The total number of ncRNAs affected by the treatments was significantly smaller compared to the mRNAs. A total of 54 miRNAs were significantly upregulated in response to AA, such as miR-1 (log2FC = 2.62), miR-133a/b (log2FC = 2.54), miR-181b (log2FC = 1.80), miR-499 (log2FC = 1.54) or miR-206 (log2FC = 1.48); and 26 miRNAs were downregulated in response to AA, including miR-29d (log2FC = −2.79), miR-203a/b (log2FC = −1.38) or miR-122 (log2FC = −0.77) (Figure 4 and Figure 5; Supplementary File S3). Gene Ontology analysis based on human miRNA–mRNA interactions showed that miRNA modified by the presence of AA might control mRNA involved in protein and ATP binding and regulation of transcription (Figure 7). On the other hand, in response to Igf-1, only 20 miRNAs significantly increased their expression in response to Igf-1, such as miR-27c (log2FC = 1.67), miR-1 (log2FC = 1.56), miR-19a/b (log2FC = 1.06), or miR-133a/b (log2FC = 0.77); and a total of 26 miRNA appeared downregulated but most of them with a change log2FC < −1, such as miR-203a/b (log2FC = −0.83), miR-128 (log2FC = −0.83); miR-122 (log2FC = −0.71), miR-206 (log2FC = −0.72), miR-27a (log2FC = −0.48) and miR-221 (log2FC = −0.23) (Figure 4 and Figure 5; Supplementary File S3). Gene Ontology analysis based on human data predicted that those miRNAs controlled mRNAs involved in transmembrane transport, protein phosphorylation, signal transduction, and ATP binding (Figure 7).
The number of lncRNAs significantly modified was also small compared to mRNAs and miRNAs. In response to AA, only 17 lncRNAs appeared to be significantly upregulated with a log2FC between 1 and 2 (Figure 4 and Figure 5; Supplementary File S1). We also found 13 lncRNAs significantly downregulated in response to the presence of AA, showing a log2FC between −1 and −5. In response to Igf-1, only 4 lncRNAs were significantly increased with log2FC between 1.05 and 1.70. Similarly, only 7 lncRNAs appeared to be significantly downregulated in response to Igf-1 with log2FC between −1.20 and −12.40 (Figure 4 and Figure 5; Supplementary File S1). Due to the lack of information about GO terms associated with fish lncRNAs, no GO enrichment analysis was performed.

2.4. Predicted Interactions of miRNAs and lncRNAs with mRNAs Based on Transcriptomic Correlations and Bioinformatics Analysis

To better understand the changes in response to AA and Igf-1, correlation and binding analyses were performed between miRNAs, lncRNAs, and mRNAs. Significantly modified miRNAs, lncRNAs, and mRNAs were considered candidates for further consideration when correlations had a negative Pearson index lower than −0.80. We found up to 14,658 negative correlations between miRNAs and mRNAs and a total of 7488 negative correlations between significantly modified lncRNAs and mRNAs using all treatments (Supplementary File S4), indicating the possibility of co-regulation. To further investigate how miRNAs and lncRNAs might be involved in the variations of transcription observed in mRNAs, we estimated the probability of direct interaction between miRNAs or lncRNAs and mRNAs with a correlation lower than −0.80 using bioinformatic tools. While several strong interactions (<−25 kcal/mol) were found in response to AA (Supplementary File S5), only a handful of miRNAs dominate the majority of interactions observed, such as miR-17a, miR-128, miR-133a/b and miR-206. Similarly, in response to Igf-1, we found some miRNAs predicted to interact with multiple mRNAs, such as miR-34, miR-221, and miR-338 (Supplementary File S5). Gene Ontology enrichment analysis of the mRNAs predicted to both possibly correlate and interact with miRNAs was performed to determine the biological processes regulated by them. In the case of the AA treatment, miRNAs were involved in the downregulation of genes related to Igf binding, development, protein catabolism, sarcomere production, and DNA replication (Table 2). In the Igf-1 treatment, miRNAs were involved in the possible regulation of mRNAs related to the extracellular region and upregulation of genes related to development, DNA metabolic process, and cytoskeleton (Table 2).
We also found strong negative correlations between lncRNAs and those mRNAs significantly modified by treatments (Supplementary File S4), but only 8 lncRNAs simultaneously showed strong negative correlations (ρ < −0.80) and significant interactions (ndG < −0.10 kcal/mol) with some of the mRNAs identified to change in response to treatments such as acta1, rbm24b, h2az1, pin1, tcima, psmb3, tnni2, nupr1a, rgcc, and igfbp6a (Supplementary Table S1, Supplementary File S6).
The possibility of lncRNAs regulating mRNAs abundance by acting as miRNAs sponges was also investigated. Correlations of <−0.80 between lncRNAs and miRNAs were considered possible candidates (Table 3). From those, we found 30 lncRNAs with strong predicted interactions with miRNAs, which in turn possibly regulate multiple mRNAs, such as ENSSAUG00010001802 (interacting with miR-27a, miR-29d and miR-29b), ENSSAUG00010012228 (interacting with miR-338, miR-133a/b, miR-17a, miR-125a, miR-106, miR-217, and miR-206) or ENSSAUG00010017089 (interacting with miR-206, miR-106, miR-128, and miR-17a) (Table 3).

3. Discussion

Understanding the regulation of fish muscle development and growth is necessary to optimize aquaculture production because it is the most valuable part of the fish for the aquaculture industry. To thoroughly study the mechanisms orchestrating the myogenesis process, it is necessary to consider the complex networks integrating not only the transcription of genes but also of ncRNAs like miRNAs and lncRNAs [23,73]. For this purpose, fish myoblast cell culture is a very useful and powerful tool that allows the analysis of many signaling pathways and molecular networks under controlled conditions [74]. In this study, a cell culture of gilthead sea bream myoblasts was used to explore for the first time in fish the transcriptional response of mRNAs, miRNAs, and lncRNAs in response to AA and Igf-1, as well as their possible regulatory network.
Both pro-growth signals induced many transcriptional changes compared to untreated cells, but the AA group showed a higher number of transcriptionally modified mRNAs compared to Igf-1 (Figure 4). These results are in agreement with previous studies in pacu (Piaractus mesopotamicus) [25] and Atlantic salmon (Salmo salar) [75] that showed a better capacity of AA compared to Igf-1 alone to boost myoblast response, suggesting that the Igf-1 might need the assistance of AA to perform its function. Studies in L6 murine muscle cell lines have shown that blocking Igf-1 expression did not decrease the protein synthesis rate when induced by AA, indicating that Igf-1 transcription is a covariate to AA initiation of protein synthesis through an unknown process [76]. It is well known that Igf-1 performs its functions through the phosphorylation of Akt, which leads to the promotion of cell proliferation and protein synthesis by activating the mTOR complex 1 (mTORC1) [17,77,78]. The activation of mTORC1 can also be triggered by AA, but in this case is done through the Ragulator complex, a system believed to act independently of the Akt pathway [11,79,80]. Although it is presumed that the activation of mTORC1 by AA and Igf-1 occurs in an independent way, it might be possible that the lack of AA impairs the activation of this complex by the Igf-1/Akt pathway through a not yet described mechanism that needs further investigation.
Furthermore, there was a clear difference in the magnitude of transcriptional changes induced by both treatments: the upregulation of genes such as myoz1b, stac3, tnnt2c, igfbp2a, or usp28 was much higher in response to AA than in response to Igf-1, while downregulated genes such as plvapb, ccn5, or cav2 had their transcription less reduced in response to Igf-1 compared to AA. It is important to highlight that all these genes participate in the regulation of muscle growth by modulating mechanisms related to myogenesis and protein balance in the muscle fiber [8,81,82,83]. For instance, the upregulation of myoz1b, stac3, and tnnt2c at day 9 of culture with AA and Igf-1 confirms the correct development of myogenesis under these treatments because they are genes that encode for proteins involved in muscle contraction and are expected to increase their expression when myoblasts are fusing to form myotubes [82].
It is interesting to highlight that despite the big differences in the number of mRNAs modified and the magnitude of the changes, when GO analysis was performed for up and downregulated genes, both treatments regulated common processes related to muscle growth, differentiation, and sarcomere formation. This fact suggests that both AA and Igf-1 were able to promote the transcription of components of the molecular network controlling protein synthesis and sarcomere development. Moreover, it seemed that both treatments were able to increase DNA replication and cell proliferation (Figure 6).
Regarding the ncRNAs, we identified a comprehensive repertoire of miRNAs and lncRNAs present in gilthead seabream myoblasts with potential roles in regulating muscle growth. We found that the most expressed miRNAs in the gilthead sea bream myoblasts were miR-21, miR-146, miR-22b, and miR-206, with only the last one being a canonical myomiR [43,84], although the rest are also known to have roles on the control of skeletal muscle growth. For instance, in mammalian models, miR-21 is known to downregulate the transcription of pten [85,86], a component of the mTOR network, but also col1a1, col6a, and tgf-ß, components of the extracellular matrix [87]. On the other hand, miR-146 is known to promote myoblast differentiation through the regulation of smad4, notch1, and hmga2 [88], and miR-22b is also involved in myoblast differentiation by targeting tgfßr1 [89]. It is not surprising that these miRNAs promoting differentiation were highly expressed, considering that we used myoblasts developed for 8 days when myoblasts are slowing down proliferation and entering into the differentiation program, where TGF-ß is known to inhibit differentiation [90,91]. We only found a significant decrease in tgfb3 expression (FDR = 0) in response to AA (log2FC = −1.55), and less modulated in response to Igf-1 (log2FC = −0.90). Other components of the TGF-ß pathway, such as tgfb2, tgfb5, tgfb3, and tgfb1a, were non-significantly downregulated in response to both treatments.
Like mRNAs, more miRNAs changed their transcription in response to AA compared to Igf-1 (Figure 4). Not many miRNAs were downregulated by the pro-growth treatments, but we found low expression of miR-22b (when upregulated promotes differentiation) [89], miR-206 (promotes differentiation) [92], miR-221 (involved in proliferation and differentiation) [93] and miR-338 (function not known, but is differentially expressed in skeletal muscle of different species under different growth conditions) [25,94,95] in response to Igf-1. The fact that some differentiation-inducing miRNAs identified in mammals [21,22,43] appeared to be downregulated in the present experiment seems to be at odds with the results obtained, which suggests that both proliferation and differentiation were stimulated (as indicated by the GO enrichment analysis). However, we also found a significant increase of miRNAs that promote differentiation such as miR-1 (log2FC > 1.5; increased in response to both treatments), miR-206 (log2FC = 1.48; increased with AA), miR-499 (log2FC = 1.54; promotes differentiation toward slow phenotype, increased with AA), miR-181 (log2FC = 1.8 increased with AA) and miR-34 (log2FC = 1, inhibits proliferation, increased with Igf-1). At the same time, an upregulation of miRNAs generally associated with myoblast proliferation was also observed in response to AA, such as miR-128 (log2FC ≥ 0.78), or in response to both treatments, like the miR-133a/b (log2FC ≥ 0.66). The transcriptional changes of miRNAs and mRNAs involved in both myogenic proliferation and differentiation are likely due to the fact that the cell cultures used in the present study contain a mixture of cells at different stages, with still proliferative myoblasts but most cells differentiating.
Our analysis showed strong correlations between miRNAs and mRNAs differentially expressed in response to the treatments. However, many of the identified correlations (<−0.80) had relatively low predicted interactions (<−25 kcal/mol), suggesting that the mRNAs and miRNAs might be part of the same networks but not directly regulating each other. The strong correlations and significant interactions found were dominated by a small number of miRNAs: miR-133, miR-128 or miR-206 (upregulated) and miR-27a, miR-92a or miR-29d (downregulated) in the AA treatment; miR-128, miR-125, miR-338, miR-206 or miR-27a (downregulated) and miR-34 or miR-7147 (upregulated) in the Igf-1 treatment. The percentage of genes whose transcription seems to be potentially regulated by miRNAs was relatively low. However, we must take into consideration that in the present study, we have used quite stringent conditions, reducing the number of interactions identified. Likewise, the correlations were performed with only nine samples, and the strength of such correlations must be considered cautiously.
Unraveling the roles of lncRNAs in fish skeletal muscle based on transcriptomic data is quite challenging, and we can only hypothesize their possible functions using bioinformatic approaches. The study of lncRNAs in mammals has revealed their importance in the transcriptomic regulation of muscle development, and some lncRNAs have been shown to be critical in the control of muscle gene expression, including the linc-RAM (enhances myogenic differentiation by interacting with MyoD) [63], MUNC (increases MyoD, Myogenin, and Myh3 mRNAs and facilitates the function of MyoD) [61,62], OIP5-AS1 (interacts with MEF2C mRNA and promotes myogenic gene expression) [96], or Lnc-31 (binds to Rock1 mRNA and sustains myoblast proliferation) [97]. Similarly, lncRNAs can also exert their functions directly interacting with miRNAs, such as linc-MD1 and MDNCR (interact with miR-133) [98,99], Sirt1 AS (interacts with miR-34a) [100] or linc-smad7 (interacts with miR-125b) [101], acting as miRNAs sponges [102]. However, it is very difficult to translate the research done in mammalian models to other species due to the low degree of conservation found between lncRNAs [72]. Our data indicates that only a small fraction of the lncRNAs identified responded to the pro-growth signals, with most of them showing low expression, as previously observed in other studies [55,103]. It is interesting to notice that many of the lncRNAs previously identified in gilthead sea bream skeletal muscle [24] had very low levels of expression in myoblasts developed for 8 days, although one of them, the ENSSAUG00010020194, slightly increased transcription in response to pro-growth signals (log2FC < 1), but not significantly. Similarly, its predicted target (myod1) also slightly changed its transcription (log2FC < 1) in response to pro-growth signals, but not significantly.
Furthermore, our analysis revealed that a higher number of lncRNAs simultaneously exhibit strong negative correlations and interactions with miRNAs (Table 3) compared to mRNAs (Supplementary Table S1), which changed in response to the treatments. This fact may suggest that the contribution of lncRNAs to the modulation of transcription might be done mainly as miRNAs sponges rather than through direct interactions with mRNAs. Among the miRNAs that negatively correlate and interact with lncRNAs are those associated with multiple mRNAs modified by treatments: miR-338, miR-92, miR-34 miR-206, miR-133, miR-7147, miR-27, miR-29, miR-125 and miR-128 (Table 3). This indicates that highly expressed lncRNAs bind to the miRNAs, preventing the degradation of target mRNAs, which appear increased (and vice versa). Supplementary Figures S1 to S3 show examples of possible networks of mRNAs, miRNAs, and lncRNAs controlling some biological processes in response to AA and Igf-1. For example, Supplementary Figure S2A exposes a group of genes involved in muscle development that were upregulated with AA and could be affected by some miRNAs (miR-27a, miR-29d, miR-92a) that, in turn, might be sequestered by specific lncRNAs acting as sponges. These figures show part of the distinct levels in the transcriptional regulation and illustrate the complexity behind the interactions between different molecules.
Moreover, it is interesting to note that some of the interactions found in our study are also predicted for some human lncRNAs, such as linc-MD1 (miR-133) [98], Sirt1 AS (miR-34) [100], or lnc-mg (miR-125) [104]. The results suggest that some roles as sponges of lncRNAs in muscle might be conserved in teleost fish. However, it is important to highlight that we have found a relatively low conservation between lncRNAs with similar interactions in fish and mammals. For instance, ENSSAUG00010016143 (which interacts with acta1; Supplementary Table S1) had a 44% similarity with Myolinc [105] and not quite a good alignment, and the majority of lncRNAs identified to interact with miR-133 have less than 30% similarity with linc-MD1. Similarly, we did not find any clear conservation of the synteny between mammalian and fish lncRNAs with conserved targets, suggesting that while lncRNA interactions might be conserved, their evolution history is not clear.
Overall, this work is the first step in the identification of the network of mRNAs, miRNAs, and lncRNAs controlling muscle development and growth in gilthead sea bream, pointing out potential candidates with a high confidence value that might be of great interest for further experimental work. Moreover, this study contributes to a better understanding of the modulation of mRNAs and ncRNAs transcription by AA and Igf-1, along with their potential regulatory mechanisms in this species, and establishes the basis for future research focusing on the possible dose-dependent response of these pro-growth signals and exploring their synergistic effects.

4. Materials and Methods

4.1. Gilthead Sea Bream Primary Myoblast Cell Culture and Treatments

Myoblasts were isolated and cultured according to the protocol described by Fauconneau and Paboeuf (2000) [106] and adapted to gilthead sea bream by Montserrat et al. (2007) [107]. Briefly, fast-twitch muscles were collected from the epaxial region of gilthead sea bream fingerlings (≈ 5 g) and mechanically dissociated with scalpels, enzymatically digested with 0.2% collagenase type IA (Ref. C9891) and 0.1% trypsin (Ref. T4799), filtered with cell strainers (Ref. CLS431752 and CLS431750), centrifuged, resuspended and plated in poly-L-lysine/laminin (Ref. P6282 and L2020) pre-treated 6-well plates (Ref. 140675) with complete growth medium [DMEM (Ref. D7777), 9 mM NaHCO3 (Ref. S5761), 20 mM HEPES (Ref. H3375), 1.1 g/L NaCl (Ref. S5886), 1% antibiotic/antimycotic (Ref. A5955), and 10% fetal bovine serum (FBS; Ref. F7524), pH 7.4], at a density of 1.5 × 106 cells/well. All media, reagents, and cell strainers were obtained from Sigma-Aldrich (Tres Cantos, Madrid, Spain), and the culture plates were obtained from Thermo Fisher Scientific (Sant Cugat del Vallès, Barcelona, Spain). Myoblasts were incubated at 22 °C, with a full replacement of the culture medium every day. Myoblasts morphology was monitored regularly under an inverted microscope (Carl Zeiss, Oberkochen, Germany) and let to develop until the first myoblast fusion events were visible (around day 8 of culture). The present work was based on 3 independent cell cultures.
On day 8 of culture, myoblasts were incubated for 12 h in a free AA medium [Earle’s balanced salt solution 1× (Ref. E7510), 9 mM NaHCO3, 20 mM HEPES, 1.1 g/L NaCl, Vitamin Mix 1× (Ref. M6895), 1% antibiotic/antimycotic, and 4 g/L D-glucose (Ref. G8270)] to reduce gene expression to basal levels. Cells were incubated for additional 24 h in free AA medium (CTR group), medium with AA (AA group; DMEM, 9 mM NaHCO3, 20 mM HEPES, 1.1 g/L NaCl, and 1% antibiotic/antimycotic), or medium with recombinant Igf-1 (Igf-1 group) [free AA medium supplemented with Igf-1 from gilthead sea bream at 100 ng/mL (Ref. CYT-295, ProSpec, Rehovot, Israel), and 0.1 mg/mL of RIA grade bovine serum albumin (Ref. A7030, Sigma-Aldrich) as carrier protein]. The treatments were performed according to the protocol described by Bower and Johnston (2010) [75] and Garcia de la serrana and Johnston (2013) [108].

4.2. RNA Extraction, Sequencing, and Bioinformatic Analyses

After the treatments, gilthead sea bream myoblasts were washed thrice with PBS following medium removal. Total RNA was extracted using Trizol (Ref. 15596026, Thermo Fisher Scientific), followed by chloroform, isopropanol, and ethanol extraction as recommended by the manufacturer. Total RNA was resuspended in RNase-free water, and its concentration and integrity were estimated by spectrophotometry using Nanodrop 2200TM (Thermo Fisher Scientific) and a 1% (w/v) agarose gel, respectively.
The generation of DNA libraries and sequencing of mRNAs and miRNAs were performed by LC Sciences (Houston, TX, USA). Transcriptome was obtained through the NovaSeq 6000 platform (Illumina, San Diego, CA, USA) with 150 base pairs, paired-end, and 6 GB data per sample (40–50 million reads). microRNAome was obtained through the HiSeq 4000 platform (Illumina, San Diego, CA, USA) with 50 base pairs, single-end, and 10 million reads per sample. For transcriptome analysis, adapters and low-quality reads were removed using an in-house perl script and then mapped against the latest gilthead sea bream genome available (www.ensembl.org/index.html; accessed on 15 January 2023) using HISAT2 software v.2.2.1 [109]. Transcripts were assembled, followed by mRNA expression profiling analysis using StringTie v.2.2.0 [110] and expressing the results as FPKM (fragments per kilobase of exon per million fragments mapped). For the microRNAome, adapters and low-quality reads were removed using in-house perl scripts. Subsequently, unique sequences with length in 18–26 nucleotides were mapped to specific species precursors in miRBase 22.0 (www.mirbase.org, accessed on 18 December 2022) by BLAST search to identify known miRNAs and novel 5p- and 3p- derived miRNAs candidates. The remaining sequences were mapped to other selected species precursors (with the exclusion of specific species) in miRBase v.22.1 by BLAST search, and the mapped pre-miRNAs were further BLASTed against the specific species genomes to determine their genomic locations.
Gene Ontology (GO) analysis was performed using the STRING online tool against the zebrafish (Danio rerio) database (https://string-db.org/, accessed on 18 December 2022). Venn diagrams were obtained using plotting software (https://pnnl-comp-mass-spec.github.io/Venn-Diagram-Plotter/, v.1.6.7458, accessed on 20 July 2023).
Pearson correlation analysis was carried out using RStudio v.1.1.419 [111] to detect correlations between mRNAs-miRNAs and lncRNAs-miRNAs differentially expressed in response to the treatments. Sequences’ interactions were predicted using RNAhybrid v.2.2.1 (https://bibiserv.cebitec.uni-bielefeld.de/rnahybrid, accessed on 15 January 2023) [112], with a minimum free energy (MFE) threshold of <−25 kcal/mol. Possible interactions between lncRNAs and mRNAs were explored using LncTar software (www.cuilab.cn/lnctar, accessed on 15 January 2023), with a threshold of normalized binding free energy (ndG) < −0.10.

4.3. Validation of RNA-Seq Results by qPCR

To validate the expression profiles from the RNA-Seq analysis using qPCR, we selected mRNAs, miRNAs, and lncRNAs that showed significant differences between the experimental groups in the RNA-Seq analysis. We used samples of the three experimental conditions (CTR, AA, and Igf-1, explained in Section 4.1) from six independent cell cultures. Total RNA was extracted as previously described (Section 4.2). The qPCR analyses were carried out following the MIQE guidelines [113] in a CFX384™ Real-Time System (Bio-Rad, El Prat de Llobregat, Barcelona, Spain). The analysis was performed in triplicate, using for each well: 2.5 µL of iTAQ Universal SYBR® Green Supermix (Ref: 1725125, Bio-Rad), 1 µL of cDNA, 250 nM (final concentration) of forward and reverse primers and 1.25 µL of DEPC water. The reaction protocol was: 3 min at 95 °C, 40 × (10 s at 95 °C, 30 s at the annealing temperature of the primers, and fluorescence detection), followed by an amplicon dissociation analysis. In the case of the miRNAs, we designed primers to amplify pri-miRNAs sequences, to distinguish between the expression of different paralogs that have similar mature sequences. The genes analyzed were igfbp6, cav3, trim63, acta1, stac3, usp28, myoz1b, cpt1b, wnt4, and two reference genes, rps18 and tomm20b. The pri-miRNAs were pri-miR-1-2, pri-miR-133a-1, pri-miR-133a-2, pri-miR-133b, pri-miR-29a, pri-miR-206, pri-miR-221, and pri-miR-222. The lncRNAs were ENSSAUG00010012549, ENSSAUG00010001802, ENSSAUG00010004711, and ENSSAUG00010020194. The transcript abundance was calculated using the Bio-Rad CFX Manager™ 3.1 software, relative to the geometric mean of the reference genes [114]. Statistical analyses were performed using IBM SPSS Statistics v. 25 (IBM Corp., Armonk, NY, USA). The normality and homoscedasticity of the data were checked with a Shapiro–Wilk test and a Levene’s test, respectively. Groups were compared using one-way ANOVA followed by a Tukey’s post hoc test (significant differences considered at p-value < 0.05). All raw and processed data from these analyses and the primers used for the qPCRs are shown in Supplementary File S7. Transcript levels of genes, pri-miRNAs, and lncRNAs showed concordance between RNA-Seq and qPCR results, revealing similar expression patterns in both cases.

4.4. Statistics of RNA-Seq Data

Differences in transcription levels between treatments obtained from RNA-Seq data were biologically relevant when log2-fold change (log2FC) was ≤−1 and ≥1 and the corrected p-value (False Discovery Rate, FDR) was ≤0.05. In the case of miRNA-Seq data, only the FDR threshold was considered. For Gene Ontology analysis, differences between categories were compared against the zebrafish database and considered significant when FDR < 0.05. All graphs were generated using ggplot2 [115].

5. Conclusions

In summary, both AA and Igf-1 treatments induced the transcription of components related to myogenesis (proliferation and differentiation), sarcomere formation, and protein synthesis, but AA appeared to have a greater impact on the transcriptional response of genes and ncRNAs compared to Igf-1. Some of the miRNAs most regulated by the pro-growth signals were canonical myomiRs with known roles in myogenic mechanisms, such as miR-1, miR-133a/b, and miR-206, but also other miRNAs with more unknown functions in muscle, such as miR-203a/b or miR-122. In contrast, few lncRNAs responded to the treatments, with most of them showing very low expression, but interestingly, our study suggests that the lncRNAs act mainly as miRNAs sponges in response to AA and Igf-1. Furthermore, the results of the correlations and predicted interactions between mRNAs, miRNAs, and lncRNAs point out the importance and complexity of the network controlling muscle development and growth in response to pro-growth signals in gilthead sea bream fast muscle myoblasts. These results will serve as significant resources for future studies that further investigate the role of ncRNAs in the myogenesis processes of teleost.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25073894/s1.

Author Contributions

Conceptualization, D.G.d.l.s., M.D.-P.-S. and B.O.S.D.; methodology, D.G.d.l.s.; validation, I.G.-P.; formal analysis, D.G.d.l.s., B.O.S.D. and I.G.-P.; investigation, D.G.d.l.s. and I.G.-P.; resources, D.G.d.l.s., M.D.-P.-S. and B.O.S.D.; data curation, D.G.d.l.s. and B.O.S.D.; writing—original draft preparation, D.G.d.l.s., I.G.-P. and B.O.S.D.; writing—review and editing, D.G.d.l.s., I.G.-P., B.O.S.D. and M.D.-P.-S.; visualization: D.G.d.l.s., I.G.-P. and B.O.S.D.; supervision, D.G.d.l.s.; project administration, D.G.d.l.s. and M.D.-P.-S.; funding acquisition, D.G.d.l.s., B.O.S.D. and M.D.-P.-S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the São Paulo Research Foundation (FAPESP), grant numbers #2018/24575-6 and #2018/26428−0. This study was also financed by the National Council for Scientific and Technological Development (Conselho Nacional de Desenvolvimento Científico e Tecnológico; CNPq), grant numbers #306678/2021-7 and #403305/2021-7, and in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior—Brasil (CAPES)—Finance Code 001. I.G.-P. is supported by a predoctoral fellowship (PRE2019-089578) funded by MICIU/AEI/10.13039/501100011033 and by “ERDF A way of making Europe”. B.O.S.D. is supported by the grant #303991/2022-4, National Council for Scientific and Technological Development (CNPq).

Institutional Review Board Statement

The Ethics and Animal Care Committee of the University of Barcelona approved all the procedures involving fish manipulation and tissue collection (permit number CEEA OB 72/17). The animal-handling procedures were carried out in accordance with the current legislation established by the Council of the European Union (EU 2010/63) and the governments of Spain and Catalonia.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the current article and its corresponding Supplementary material. The raw and processed data of transcriptome and microRNAome analyses have been deposited on the Gene Expression Omnibus (GEO) DataSets, under the accession number GSE246665.

Acknowledgments

The authors would like to thank the personnel from the facilities at the School of Biology (CCiTUB) for the maintenance of the fish and Piscimar for providing the fish.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Mommsen, T.P. Paradigms of Growth in Fish. Comp. Biochem. Physiol.-B Biochem. Mol. Biol. 2001, 129, 207–219. [Google Scholar] [CrossRef] [PubMed]
  2. Johnston, I.A. Environment and Plasticity of Myogenesis in Teleost Fish. J. Exp. Biol. 2006, 209, 2249–2264. [Google Scholar] [CrossRef] [PubMed]
  3. Weatherley, 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, 851–859. [Google Scholar] [CrossRef]
  4. Dal-Pai-Silva, M.; Zanella, B.T.T.; Duran, B.O.S.; Almeida, F.L.A.; Mareco, E.A.; de Paula, T.G. Cellular and Molecular Features of Skeletal Muscle Growth and Plasticity. In Biology and Physiology ofFreshwater Neotropical Fish; Baldisserotto, B., Urbinati, E.C., Cyrino, J.E.P., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 163–183. [Google Scholar] [CrossRef]
  5. Garcia de la serrana, D.; Codina, M.; Capilla, E.; Jiménez-Amilburu, V.; Navarro, I.; Du, S.J.; Johnston, I.A.; Gutiérrez, J. Characterisation and Expression of Myogenesis Regulatory Factors during in Vitro Myoblast Development and in Vivo Fasting in the Gilthead Sea Bream (Sparus aurata). Comp. Biochem. Physiol.-A Mol. Integr. Physiol. 2014, 167, 90–99. [Google Scholar] [CrossRef]
  6. Hernández-Hernández, J.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] [PubMed]
  7. Perelló-Amorós, M.; Otero-Tarrazón, A.; Jorge-Pedraza, V.; García-Pérez, I.; Sánchez-Moya, A.; Gabillard, J.C.; Moshayedi, F.; Navarro, I.; Capilla, E.; Fernández-Borràs, J.; et al. Myomaker and Myomixer Characterization in Gilthead Sea Bream under Different Myogenesis Conditions. Int. J. Mol. Sci. 2022, 23, 14639. [Google Scholar] [CrossRef]
  8. Lehka, L.; Rędowicz, M.J. Mechanisms Regulating Myoblast Fusion: A Multilevel Interplay. Semin. Cell Dev. Biol. 2020, 104, 81–92. [Google Scholar] [CrossRef]
  9. Sancak, Y.; Bar-Peled, L.; Zoncu, R.; Markhard, A.L.; Nada, S.; Sabatini, D.M. Ragulator-Rag Complex Targets MTORC1 to the Lysosomal Surface and Is Necessary for Its Activation by Amino Acids. Cell 2010, 141, 290–303. [Google Scholar] [CrossRef]
  10. Sancak, Y.; Peterson, T.R.; Shaul, Y.D.; Lindquist, R.A.; Thoreen, C.C.; Bar-Peled, L.; Sabatini, D.M. The Rag GTPases Bind Raptor and Mediate Amino Acid Signaling to MTORC1. Science 2008, 320, 1496–1501. [Google Scholar] [CrossRef]
  11. Manifava, M.; Smith, M.; Rotondo, S.; Walker, S.; Niewczas, I.; Zoncu, R.; Clark, J.; Ktistakis, N.T. Dynamics of MTORC1 Activation in Response to Amino Acids. Elife 2016, 5, e19960. [Google Scholar] [CrossRef]
  12. Yao, Y.; Jones, E.; Inoki, K. Lysosomal Regulation of MTORC1 by Amino Acids in Mammalian Cells. Biomolecules 2017, 7, 51. [Google Scholar] [CrossRef] [PubMed]
  13. Vélez, E.J.; Azizi, S.; Verheyden, D.; Salmerón, C.; Lutfi, E.; Sánchez-Moya, A.; Navarro, I.; Gutiérrez, J.; Capilla, E. Proteolytic Systems’ Expression during Myogenesis and Transcriptional Regulation by Amino Acids in Gilthead Sea Bream Cultured Muscle Cells. PLoS ONE 2017, 12, e0187339. [Google Scholar] [CrossRef] [PubMed]
  14. Seiliez, I.; Gabillard, J.C.; Riflade, M.; Sadoul, B.; Dias, K.; Avérous, J.; Tesseraud, S.; Skiba, S.; Panserat, S. Amino Acids Downregulate the Expression of Several Autophagy-Related Genes in Rainbow Trout Myoblasts. Autophagy 2012, 8, 364–375. [Google Scholar] [CrossRef] [PubMed]
  15. Cleveland, B.M.; Weber, G.M. Effects of Insulin-like Growth Factor-I, Insulin, and Leucine on Protein Turnover and Ubiquitin Ligase Expression in Rainbow Trout Primary Myocytes. Am. J. Physiol.-Regul. Integr. Comp. Physiol. 2010, 298, 341–350. [Google Scholar] [CrossRef] [PubMed]
  16. Fuentes, E.N.; Valdés, J.A.; Molina, A.; Björnsson, B.T. Regulation of Skeletal Muscle Growth in Fish by the Growth Hormone-Insulin-like Growth Factor System. Gen. Comp. Endocrinol. 2013, 192, 136–148. [Google Scholar] [CrossRef] [PubMed]
  17. Duan, C.; Ren, H.; Gao, S. Insulin-like Growth Factors (IGFs), IGF Receptors, and IGF-Binding Proteins: Roles in Skeletal Muscle Growth and Differentiation. Gen. Comp. Endocrinol. 2010, 167, 344–351. [Google Scholar] [CrossRef] [PubMed]
  18. Wood, A.W.; Duan, C.; Bern, H.A. Insulin-like Growth Factor Signaling in Fish. Int. Rev. Cytol. 2005, 243, 215–285. [Google Scholar] [CrossRef]
  19. Vélez, E.J.; Azizi, S.; Millán-Cubillo, A.; Fernández-Borràs, J.; Blasco, J.; Chan, S.J.; Calduch-Giner, J.A.; Pérez-Sánchez, J.; Navarro, I.; Capilla, E.; et al. Effects of Sustained Exercise on GH-IGFs Axis in Gilthead Sea Bream (Sparus Aurata). Am. J. Physiol.-Regul. Integr. Comp. Physiol. 2016, 310, R313–R322. [Google Scholar] [CrossRef] [PubMed]
  20. Triantaphyllopoulos, K.A.; Cartas, D.; Miliou, H. Factors Influencing GH and IGF-I Gene Expression on Growth in Teleost Fish: How Can Aquaculture Industry Benefit? Rev. Aquac. 2020, 12, 1637–1662. [Google Scholar] [CrossRef]
  21. Güller, I.; Russell, A.P. MicroRNAs in Skeletal Muscle: Their Role and Regulation in Development, Disease and Function. J. Physiol. 2010, 588, 4075–4087. [Google Scholar] [CrossRef]
  22. Horak, M.; Novak, J.; Bienertova-Vasku, J. Muscle-Specific MicroRNAs in Skeletal Muscle Development. Dev. Biol. 2016, 410, 1–13. [Google Scholar] [CrossRef] [PubMed]
  23. Martone, J.; Mariani, D.; Desideri, F.; Ballarino, M. Non-Coding RNAs Shaping Muscle. Front. Cell Dev. Biol. 2020, 7, 1–15. [Google Scholar] [CrossRef] [PubMed]
  24. García-Pérez, I.; Molsosa-Solanas, A.; Perelló-Amorós, M.; Sarropoulou, E.; Blasco, J.; Gutiérrez, J.; Garcia de la serrana, D. The Emerging Role of Long Non-Coding RNAs in Development and Function of Gilthead Sea Bream (Sparus aurata) Fast Skeletal Muscle. Cells 2022, 11, 428. [Google Scholar] [CrossRef]
  25. Duran, B.O.S.; Zanella, B.T.T.; Perez, E.S.; Mareco, E.A.; Blasco, J.; Dal-Pai-Silva, M.; Garcia de la serrana, D. Amino Acids and IGF1 Regulation of Fish Muscle Growth Revealed by Transcriptome and MicroRNAome Integrative Analyses of Pacu (Piaractus mesopotamicus) Myotubes. Int. J. Mol. Sci. 2022, 23, 1180. [Google Scholar] [CrossRef] [PubMed]
  26. 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, 8416. [Google Scholar] [CrossRef]
  27. Ali, A.; Al-Tobasei, R.; Kenney, B.; Leeds, T.D.; Salem, M. Integrated Analysis of LncRNA and MRNA Expression in Rainbow Trout Families Showing Variation in Muscle Growth and Fillet Quality Traits. Sci. Rep. 2018, 8, 12111. [Google Scholar] [CrossRef]
  28. Bizuayehu, T.T.; Babiak, I. MicroRNA in Teleost Fish. Genome Biol. Evol. 2014, 6, 1911. [Google Scholar] [CrossRef]
  29. Filipowicz, W.; Bhattacharyya, S.N.; Sonenberg, N. Mechanisms of Post-Transcriptional Regulation by MicroRNAs: Are the Answers in Sight? Nat. Rev. Genet. 2008 92 2008, 9, 102–114. [Google Scholar] [CrossRef]
  30. Bartel, D.P. MicroRNA Target Recognition and Regulatory Functions. Cell 2009, 136, 215. [Google Scholar] [CrossRef]
  31. Bartel, D.P.; Chen, C.Z. Micromanagers of Gene Expression: The Potentially Widespread Influence of Metazoan MicroRNAs. Nat. Rev. Genet. 2004, 5, 396–400. [Google Scholar] [CrossRef]
  32. Koutsoulidou, A.; Mastroyiannopoulos, N.P.; Furling, D.; Uney, J.B.; Phylactou, L.A. Expression of MiR-1, MiR-133a, MiR-133b and MiR-206 Increases during Development of Human Skeletal Muscle. BMC Dev. Biol. 2011, 11, 34. [Google Scholar] [CrossRef] [PubMed]
  33. Mok, G.F.; Lozano-Velasco, E.; Münsterberg, A. MicroRNAs in Skeletal Muscle Development. Semin. Cell Dev. Biol. 2017, 72, 67–76. [Google Scholar] [CrossRef] [PubMed]
  34. Dey, B.K.; Gagan, J.; Dutta, A. MiR-206 and -486 Induce Myoblast Differentiation by Downregulating Pax7. Mol. Cell. Biol. 2011, 31, 203–214. [Google Scholar] [CrossRef] [PubMed]
  35. 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] [PubMed]
  36. Winbanks, C.E.; Beyer, C.; Hagg, A.; Qian, H.; Sepulveda, P.V.; Gregorevic, P. MiR-206 Represses Hypertrophy of Myogenic Cells but Not Muscle Fibers via Inhibition of HDAC4. PLoS ONE 2013, 8, e73589. [Google Scholar] [CrossRef] [PubMed]
  37. Winbanks, C.E.; Wang, B.; Beyer, C.; Koh, P.; White, L.; Kantharidis, P.; Gregorevic, P. TGF-β Regulates MiR-206 and MiR-29 to Control Myogenic Differentiation through Regulation of HDAC4. J. Biol. Chem. 2011, 286, 13805–13814. [Google Scholar] [CrossRef]
  38. Yuasa, K.; Hagiwara, Y.; Ando, M.; Nakamura, A.; Takeda, S.; Hijikata, T. MicroRNA-206 Is Highly Expressed in Newly Formed Muscle Fibers: Implications Regarding Potential for Muscle Regeneration and Maturation in Muscular Dystrophy. Cell Struct. Funct. 2008, 33, 163–169. [Google Scholar] [CrossRef] [PubMed]
  39. Liu, N.; Williams, A.H.; Maxeiner, J.M.; Bezprozvannaya, S.; Shelton, J.M.; Richardson, J.A.; Bassel-Duby, R.; Olson, E.N. MicroRNA-206 Promotes Skeletal Muscle Regeneration and Delays Progression of Duchenne Muscular Dystrophy in Mice. J. Clin. Investig. 2012, 122, 2054–2065. [Google Scholar] [CrossRef] [PubMed]
  40. Chen, J.F.; Mandel, E.M.; Thomson, J.M.; Wu, Q.; Callis, T.E.; 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. 2006, 38, 228–233. [Google Scholar] [CrossRef]
  41. Luo, Y.; Wu, X.; Ling, Z.; Yuan, L.; Cheng, Y.; Chen, J.; Xiang, C. MicroRNA133a Targets Foxl2 and Promotes Differentiation of C2C12 into Myogenic Progenitor Cells. DNA Cell Biol. 2015, 34, 29. [Google Scholar] [CrossRef]
  42. Feng, Y.; Niu, L.L.; Wei, W.; Zhang, W.Y.; Li, X.Y.; Cao, J.H.; Zhao, S.H. A Feedback Circuit between MiR-133 and the ERK1/2 Pathway Involving an Exquisite Mechanism for Regulating Myoblast Proliferation and Differentiation. Cell Death Dis. 2013, 4, e934. [Google Scholar] [CrossRef] [PubMed]
  43. McCarthy, J.J. The MyomiR Network in Skeletal Muscle Plasticity. Exerc. Sport Sci. Rev. 2011, 39, 150. [Google Scholar] [CrossRef] [PubMed]
  44. 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]
  45. van Rooij, E.; Quiat, D.; Johnson, B.A.; Sutherland, L.B.; Qi, X.; Richardson, J.A.; Kelm, R.J.; Olson, E.N. A Family of MicroRNAs Encoded by Myosin Genes Governs Myosin Expression and Muscle Performance. Dev. Cell 2009, 17, 662–673. [Google Scholar] [CrossRef] [PubMed]
  46. Yan, B.; Guo, J.T.; Zhao, L.H.; Zhao, J.L. MicroRNA Expression Signature in Skeletal Muscle of Nile Tilapia. Aquaculture 2012, 364–365, 240–246. [Google Scholar] [CrossRef]
  47. 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]
  48. Duran, B.O.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-Silvca, M. Differential MicroRNA Expression in Fast- and Slow-Twitch Skeletal Muscle of Piaractus mesopotamicus during Growth. PLoS ONE 2015, 10, e0144481. [Google Scholar] [CrossRef] [PubMed]
  49. 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. [Google Scholar] [CrossRef]
  50. Huang, M.B.; Xu, H.; Xie, S.J.; Zhou, H.; Qu, L.H. Insulin-like Growth Factor-1 Receptor Is Regulated by MicroRNA-133 during Skeletal Myogenesis. PLoS ONE 2011, 6, e29173. [Google Scholar] [CrossRef]
  51. de Paula, T.G.; Zanella, B.T.T.; de Almeida Fantinatti, B.E.; de Moraes, L.N.; Duran, B.O.S.; de Oliveira, C.B.; Salomäo, R.A.S.; Da Silva, R.N.; Padovani, C.R.; Dos Santos, V.B.; et al. Food Restriction Increase the Expression of MTORC1 Complex Genes in the Skeletal Muscle of Juvenile Pacu (Piaractus mesopotamicus). PLoS ONE 2017, 12, e0177679. [Google Scholar] [CrossRef]
  52. Yan, B.; Zhu, C.D.; Guo, J.T.; Zhao, L.H.; Zhao, J.L. MiR-206 Regulates the Growth of the Teleost Tilapia (Oreochromis niloticus) through the Modulation of IGF-1 Gene Expression. J. Exp. Biol. 2013, 216, 1265–1269. [Google Scholar] [CrossRef] [PubMed]
  53. 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] [PubMed]
  54. Duran, B.O.S.; Dal-Pai-Silva, M.; Garcia de la serrana, D. 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, jeb216390. [Google Scholar] [CrossRef]
  55. Statello, L.; Guo, C.-J.; Chen, L.-L.; Huarte, M. Gene Regulation by Long Non-Coding RNAs and Its Biological Functions. Nat. Rev. Mol. Cell Biol. 2021, 22, 96–118. [Google Scholar] [CrossRef] [PubMed]
  56. Neguembor, M.V.; Jothi, M.; Gabellini, D. Long Noncoding RNAs, Emerging Players in Muscle\ndifferentiation and Disease. Skelet. Muscle 2014, 4, 8. [Google Scholar] [CrossRef] [PubMed]
  57. Long, Y.; Wang, X.; Youmans, D.T.; Cech, T.R. How Do LncRNAs Regulate Transcription? Sci. Adv. 2017, 3, eaao2110. [Google Scholar] [CrossRef] [PubMed]
  58. Paraskevopoulou, M.D.; Hatzigeorgiou, A.G. Analyzing MiRNA–LncRNA Interactions. Methods Mol. Biol. 2016, 1402, 271–286. [Google Scholar] [CrossRef]
  59. Yoon, J.H.; Abdelmohsen, K.; Gorospe, M. Posttranscriptional Gene Regulation by Long Noncoding RNA. J. Mol. Biol. 2013, 425, 3723–3730. [Google Scholar] [CrossRef]
  60. Wang, S.; Zuo, H.; Jin, J.; Lv, W.; Xu, Z.; Fan, Y.; Zhang, J.; Zuo, B. Long Noncoding RNA Neat1 Modulates Myogenesis by Recruiting Ezh2. Cell Death Dis. 2019, 10, 505. [Google Scholar] [CrossRef]
  61. Mueller, A.C.; Cichewicz, M.A.; Dey, B.K.; Layer, R.; Reon, B.J.; Gagan, J.R.; Dutta, A. MUNC, a Long Noncoding RNA That Facilitates the Function of MyoD in Skeletal Myogenesis. Mol. Cell. Biol. 2015, 35, 498–513. [Google Scholar] [CrossRef]
  62. Cichewicz, M.A.; Kiran, M.; Przanowska, R.K.; Sobierajska, E.; Shibata, Y.; Dutta, A. MUNC, an Enhancer RNA Upstream from the MYOD Gene, Induces a Subgroup of Myogenic Transcripts in Trans Independently of MyoD. Mol. Cell. Biol. 2018, 38, e00655-17. [Google Scholar] [CrossRef]
  63. Yu, X.; Zhang, Y.; Li, T.; Ma, Z.; Jia, H.; Chen, Q.; Zhao, Y.; Zhai, L.; Zhong, R.; Li, C.; et al. Long Non-Coding RNA Linc-RAM Enhances Myogenic Differentiation by Interacting with MyoD. Nat. Commun. 2017, 8, 14016. [Google Scholar] [CrossRef] [PubMed]
  64. Sui, Y.; Han, Y.; Zhao, X.; Li, D.; Li, G. Long Non-Coding RNA Irm Enhances Myogenic Differentiation by Interacting with MEF2D. Cell Death Dis. 2019, 10, 181. [Google Scholar] [CrossRef] [PubMed]
  65. Li, Y.; Zhang, Y.; Hu, Q.; Egranov, S.D.; Xing, Z.; Zhang, Z.; Liang, K.; Ye, Y.; Pan, Y.; Chatterjee, S.S.; et al. Functional Significance of Gain-of-Function H19 LncRNA in Skeletal Muscle Differentiation and Anti-Obesity Effects. Genome Med. 2021, 13, 137. [Google Scholar] [CrossRef]
  66. Xu, H.; Cao, L.; Sun, B.; Wei, Y.; Liang, M. Transcriptomic Analysis of Potential “LncRNA-MRNA” Interactions in Liver of the Marine Teleost Cynoglossus Semilaevis Fed Diets with Different DHA/EPA Ratios. Front. Physiol. 2019, 10, 331. [Google Scholar] [CrossRef]
  67. Núñez-Acuña, G.; Détrée, C.; Gallardo-Escárate, C.; Gonçalves, A.T. Functional Diets Modulate LncRNA-Coding RNAs and Gene Interactions in the Intestine of Rainbow Trout Oncorhynchus mykiss. Mar. Biotechnol. 2017, 19, 287–300. [Google Scholar] [CrossRef]
  68. Wang, M.; Jiang, S.; Wu, W.; Yu, F.; Chang, W.; Li, P.; Wang, K. Non-Coding RNAs Function as Immune Regulators in Teleost Fish. Front. Immunol. 2018, 9, 2801. [Google Scholar] [CrossRef]
  69. Cai, J.; Li, L.; Song, L.; Xie, L.; Luo, F.; Sun, S.; Chakraborty, T.; Zhou, L.; Wang, D. Effects of Long Term Antiprogestine Mifepristone (RU486) Exposure on Sexually Dimorphic LncRNA Expression and Gonadal Masculinization in Nile Tilapia (Oreochromis niloticus). Aquat. Toxicol. 2019, 215, 105289. [Google Scholar] [CrossRef] [PubMed]
  70. Valenzuela-Muñoz, V.; Váldes, J.A.; Gallardo-Escárate, C. Transcriptome Profiling of Long Non-Coding RNAs During the Atlantic Salmon Smoltification Process. Mar. Biotechnol. 2021, 23, 308–320. [Google Scholar] [CrossRef]
  71. Wu, S.; Zhang, J.; Liu, B.; Huang, Y.; Li, S.; Wen, H.; Zhang, M.; Li, J.; Li, Y.; He, F. Identification and Characterization of LncRNAs Related to the Muscle Growth and Development of Japanese Flounder (Paralichthys Olivaceus). Front. Genet. 2020, 11, 1034. [Google Scholar] [CrossRef]
  72. Quinn, J.J.; Chang, H.Y. Unique Features of Long Non-Coding RNA Biogenesis and Function. Nat. Rev. Genet. 2016, 17, 47–62. [Google Scholar] [CrossRef] [PubMed]
  73. Zhao, Y.; Chen, M.; Lian, D.; Li, Y.; Li, Y.; Wang, J.; Deng, S.; Yu, K.; Lian, Z. Non-Coding RNA Regulates the Myogenesis of Skeletal Muscle Satellite Cells, Injury Repair and Diseases. Cells 2019, 8, 988. [Google Scholar] [CrossRef] [PubMed]
  74. Vélez, E.J.; Lutfi, E.; Azizi, S.; Montserrat, N.; Riera-Codina, M.; Capilla, E.; Navarro, I.; Gutiérrez, J. Contribution of in Vitro Myocytes Studies to Understanding Fish Muscle Physiology. Comp. Biochem. Physiol. B. Biochem. Mol. Biol. 2016, 199, 67–73. [Google Scholar] [CrossRef] [PubMed]
  75. Bower, N.I.; Johnston, I.A. Transcriptional Regulation of the IGF Signaling Pathway by Amino Acids and Insulin-like Growth Factors during Myogenesis in Atlantic Salmon. PLoS ONE 2010, 5, e11100. [Google Scholar] [CrossRef] [PubMed]
  76. Iresjö, B.M.; Diep, L.; Lundholm, K. Initiation of Muscle Protein Synthesis Was Unrelated to Simultaneously Upregulated Local Production of IGF-1 by Amino Acids in Non-Proliferating L6 Muscle Cells. PLoS ONE 2022, 17, e0270927. [Google Scholar] [CrossRef] [PubMed]
  77. Yoshida, T.; Delafontaine, P. Mechanisms of IGF-1-Mediated Regulation of Skeletal Muscle Hypertrophy and Atrophy. Cells 2020, 9, 1970. [Google Scholar] [CrossRef] [PubMed]
  78. Schiaffino, S.; Mammucari, C. Regulation of Skeletal Muscle Growth by the IGF1-Akt/PKB Pathway: Insights from Genetic Models. Skelet. Muscle 2011, 1, 1–14. [Google Scholar] [CrossRef]
  79. Deldicque, L.; Theisen, D.; Francaux, M. Regulation of MTOR by Amino Acids and Resistance Exercise in Skeletal Muscle. Eur. J. Appl. Physiol. 2005, 94, 1–10. [Google Scholar] [CrossRef] [PubMed]
  80. Kim, E. Mechanisms of Amino Acid Sensing in MTOR Signaling Pathway. Nutr. Res. Pract. 2009, 3, 64. [Google Scholar] [CrossRef]
  81. Horsley, V.; Pavlath, G.K. Forming a Multinucleated Cell: Molecules That Regulate Myoblast Fusion. Cells Tissues Organs 2004, 176, 67–78. [Google Scholar] [CrossRef]
  82. Murgia, M.; Nogara, L.; Baraldo, M.; Reggiani, C.; Mann, M.; Schiaffino, S. Protein Profile of Fiber Types in Human Skeletal Muscle: A Single-Fiber Proteomics Study. Skelet. Muscle 2021, 11, 24. [Google Scholar] [CrossRef] [PubMed]
  83. Garcia de la serrana, D.; Macqueen, D.J. Insulin-like Growth Factor-Binding Proteins of Teleost Fishes. Front. Endocrinol. 2018, 9, 1–12. [Google Scholar] [CrossRef] [PubMed]
  84. Salant, G.M.; Tat, K.L.; Goodrich, J.A.; Kugel, J.F. MiR-206 Knockout Shows It Is Critical for Myogenesis and Directly Regulates Newly Identified Target mRNAs. RNA Biol. 2020, 17, 956–965. [Google Scholar] [CrossRef] [PubMed]
  85. Fang, L.; Wang, X.; Sun, Q.; Papakonstantinou, E.; Sʼng, C.; Tamm, M.; Stolz, D.; Roth, M. IgE Downregulates PTEN through MicroRNA-21-5p and Stimulates Airway Smooth Muscle Cell Remodeling. Int. J. Mol. Sci. 2019, 20, 875. [Google Scholar] [CrossRef] [PubMed]
  86. Liu, L.; Pan, Y.; Zhai, C.; Zhu, Y.; Ke, R.; Shi, W.; Wang, J.; Yan, X.; Su, X.; Song, Y.; et al. Activation of Peroxisome Proliferation–Activated Receptor-γ Inhibits Transforming Growth Factor-Β1-Induced Airway Smooth Muscle Cell Proliferation by Suppressing Smad–MiR-21 Signaling. J. Cell. Physiol. 2019, 234, 669–681. [Google Scholar] [CrossRef] [PubMed]
  87. Dzobo, K.; Dandara, C. The Extracellular Matrix: Its Composition, Function, Remodeling, and Role in Tumorigenesis. Biomimetics 2023, 8, 146. [Google Scholar] [CrossRef] [PubMed]
  88. Khanna, N.; Ge, Y.; Chen, J. MicroRNA-146b Promotes Myogenic Differentiation and Modulates Multiple Gene Targets in Muscle Cells. PLoS ONE 2014, 9, e100657. [Google Scholar] [CrossRef]
  89. Wang, H.; Zhang, Q.; Wang, B.B.; Wu, W.J.; Wei, J.; Li, P.; Huang, R. MiR-22 Regulates C2C12 Myoblast Proliferation and Differentiation by Targeting TGFBR1. Eur. J. Cell Biol. 2018, 97, 257–268. [Google Scholar] [CrossRef] [PubMed]
  90. Liu, D.; Black, B.L.; Derynck, R. TGF-β Inhibits Muscle Differentiation through Functional Repression of Myogenic Transcription Factors by Smad3. Genes Dev. 2001, 15, 2950. [Google Scholar] [CrossRef]
  91. Girardi, F.; Taleb, A.; Ebrahimi, M.; Datye, A.; Gamage, D.G.; Peccate, C.; Giordani, L.; Millay, D.P.; Gilbert, P.M.; Cadot, B.; et al. TGFβ Signaling Curbs Cell Fusion and Muscle Regeneration. Nat. Commun. 2021, 12, 750. [Google Scholar] [CrossRef]
  92. Kim, H.K.; Lee, Y.S.; Sivaprasad, U.; Malhotra, A.; Dutta, A. Muscle-Specific MicroRNA MiR-206 Promotes Muscle Differentiation. J. Cell Biol. 2006, 174, 677–687. [Google Scholar] [CrossRef] [PubMed]
  93. Liu, B.; Shi, Y.; He, H.; Cai, M.; Xiao, W.; Yang, X.; Chen, S.; Jia, X.; Wang, J.; Lai, S. MiR-221 Modulates Skeletal Muscle Satellite Cells Proliferation and Differentiation. Vitr. Cell. Dev. Biol.-Anim. 2018, 54, 147–155. [Google Scholar] [CrossRef] [PubMed]
  94. Kappeler, B.I.G.; Regitano, L.C.A.; Poleti, M.D.; Cesar, A.S.M.; Moreira, G.C.M.; Gasparin, G.; Coutinho, L.L. MiRNAs Differentially Expressed in Skeletal Muscle of Animals with Divergent Estimated Breeding Values for Beef Tenderness. BMC Mol. Biol. 2019, 20, 1. [Google Scholar] [CrossRef]
  95. Marceca, G.P.; Nigita, G.; Calore, F.; Croce, C.M. MicroRNAs in Skeletal Muscle and Hints on Their Potential Role in Muscle Wasting During Cancer Cachexia. Front. Oncol. 2020, 10, 607196. [Google Scholar] [CrossRef]
  96. Yang, J.H.; Chang, M.W.; Pandey, P.R.; Tsitsipatis, D.; Yang, X.; Martindale, J.L.; Munk, R.; De, S.; Abdelmohsen, K.; Gorospe, M. Interaction of OIP5-AS1 with MEF2C MRNA Promotes Myogenic Gene Expression. Nucleic Acids Res. 2020, 48, 12943–12956. [Google Scholar] [CrossRef]
  97. Dimartino, D.; Colantoni, A.; Ballarino, M.; Martone, J.; Mariani, D.; Danner, J.; Bruckmann, A.; Meister, G.; Morlando, M.; Bozzoni, I. The Long Non-Coding RNA Lnc-31 Interacts with Rock1 MRNA and Mediates Its YB-1-Dependent Translation. Cell Rep. 2018, 23, 733–740. [Google Scholar] [CrossRef]
  98. Cesana, M.; Cacchiarelli, D.; Legnini, I.; Santini, T.; Sthandier, O.; Chinappi, M.; Tramontano, A.; Bozzoni, I. A Long Noncoding RNA Controls Muscle Differentiation by Functioning as a Competing Endogenous RNA. Cell 2011, 147, 358–369. [Google Scholar] [CrossRef] [PubMed]
  99. Li, H.; Yang, J.; Jiang, R.; Wei, X.; Song, C.; Huang, Y.; Lan, X.; Lei, C.; Ma, Y.; Hu, L.; et al. Long Non-Coding RNA Profiling Reveals an Abundant MDNCR That Promotes Differentiation of Myoblasts by Sponging MiR-133a. Mol. Ther.-Nucleic Acids 2018, 12, 610–625. [Google Scholar] [CrossRef]
  100. Wang, Y.; Pang, W.J.; Wei, N.; Xiong, Y.; Wu, W.J.; Zhao, C.Z.; Shen, Q.W.; Yang, G.S. Identification, Stability and Expression of Sirt1 Antisense Long Non-Coding RNA. Gene 2014, 539, 117–124. [Google Scholar] [CrossRef]
  101. Song, C.; Wang, J.; Ma, Y.; Yang, Z.; Dong, D.; Li, H.; Yang, J.; Huang, Y.; Plath, M.; Ma, Y.; et al. Linc-Smad7 Promotes Myoblast Differentiation and Muscle Regeneration via Sponging MiR-125b. Epigenetics 2018, 13, 591. [Google Scholar] [CrossRef]
  102. Wang, S.; Jin, J.; Xu, Z.; Zuo, B. Functions and Regulatory Mechanisms of LncRNAs in Skeletal Myogenesis, Muscle Disease and Meat Production. Cells 2019, 8, 1107. [Google Scholar] [CrossRef]
  103. Bhat, S.A.; Ahmad, S.M.; Mumtaz, P.T.; Malik, A.A.; Dar, M.A.; Urwat, U.; Shah, R.A.; Ganai, N.A. Long Non-Coding RNAs: Mechanism of Action and Functional Utility. Non-Coding RNA Res. 2016, 1, 43–50. [Google Scholar] [CrossRef]
  104. Zhu, M.; Liu, J.; Xiao, J.; Yang, L.; Cai, M.; Shen, H.; Chen, X.; Ma, Y.; Hu, S.; Wang, Z.; et al. Lnc-Mg Is a Long Non-Coding RNA That Promotes Myogenesis. Nat. Commun. 2017, 8, 14718. [Google Scholar] [CrossRef] [PubMed]
  105. Militello, G.; Hosen, M.R.; Ponomareva, Y.; Gellert, P.; Weirick, T.; John, D.; Hindi, S.M.; Mamchaoui, K.; Mouly, V.; Döring, C.; et al. A Novel Long Non-Coding RNA Myolinc Regulates Myogenesis through TDP-43 and Filip1. J. Mol. Cell Biol. 2018, 10, 102–117. [Google Scholar] [CrossRef]
  106. Fauconneau, B.; Paboeuf, G. Effect of Fasting and Refeeding on in Vitro Muscle Cell Proliferation in Rainbow Trout (Oncorhynchus mykiss). Cell Tissue Res. 2000, 301, 459–463. [Google Scholar] [CrossRef] [PubMed]
  107. Montserrat, N.; Sánchez-Gurmaches, J.; Garcia de la serrana, D.; Navarro, I.; Gutiérrez, J. IGF-I Binding and Receptor Signal Transduction in Primary Cell Culture of Muscle Cells of Gilthead Sea Bream: Changes throughout in Vitro Development. Cell Tissue Res. 2007, 330, 503–513. [Google Scholar] [CrossRef]
  108. Garcia de la serrana, D.; Johnston, I.A. Expression of Heat Shock Protein (Hsp90) Paralogues Is Regulated by Amino Acids in Skeletal Muscle of Atlantic Salmon. PLoS ONE 2013, 8, e74295. [Google Scholar] [CrossRef] [PubMed]
  109. Kim, D.; Langmead, B.; Salzberg, S.L. HISAT: A Fast Spliced Aligner with Low Memory Requirements. Nat. Methods 2015, 12, 357–360. [Google Scholar] [CrossRef]
  110. Pertea, M.; Pertea, G.M.; Antonescu, C.M.; Chang, T.C.; Mendell, J.T.; Salzberg, S.L. StringTie Enables Improved Reconstruction of a Transcriptome from RNA-Seq Reads. Nat. Biotechnol. 2015, 33, 290–295. [Google Scholar] [CrossRef]
  111. RStudio Team. RStudio: Integrated Development for R; RStudio, PBC: Boston, MA, USA, 2020; Available online: http://www.rstudio.com/ (accessed on 15 January 2023).
  112. Krüger, J.; Rehmsmeier, M. RNAhybrid: MicroRNA Target Prediction Easy, Fast and Flexible. Nucleic Acids Res. 2006, 34, W451-4. [Google Scholar] [CrossRef]
  113. Bustin, S.A.; Benes, V.; Garson, J.A.; Hellemans, J.; Huggett, J.; Kubista, M.; Mueller, R.; Nolan, T.; Pfaffl, M.W.; Shipley, G.L.; et al. The MIQE Guidelines: Minimum Information for Publication of Quantitative Real-Time PCR Experiments. Clin. Chem. 2009, 55, 611–622. [Google Scholar] [CrossRef] [PubMed]
  114. Pfaffl, M.W. A new mathematical model for relative quantification in real-time RT-PCR. Nucleic Acids Res. 2001, 29, e45. [Google Scholar] [CrossRef] [PubMed]
  115. Wickham, H. Ggplot2: Elegrant Graphics for Data Analysis, 2nd ed.; Springer: Cham, Switzerland, 2016. [Google Scholar] [CrossRef]
Figure 1. Representative bright-field images of muscle cells throughout the culture, from day 1 to day 9.
Figure 1. Representative bright-field images of muscle cells throughout the culture, from day 1 to day 9.
Ijms 25 03894 g001
Figure 2. miRNAs (A) and lncRNAs (B) transcription levels identified in 8 days developed gilthead sea bream myoblasts. Transcription levels of lncRNAs are expressed as FPKM (fragments per kilobase of exon per million fragments mapped), while miRNAs are expressed as normalized reads. The insert represents the expression of the first 100 miRNAs and lncRNAs.
Figure 2. miRNAs (A) and lncRNAs (B) transcription levels identified in 8 days developed gilthead sea bream myoblasts. Transcription levels of lncRNAs are expressed as FPKM (fragments per kilobase of exon per million fragments mapped), while miRNAs are expressed as normalized reads. The insert represents the expression of the first 100 miRNAs and lncRNAs.
Ijms 25 03894 g002
Figure 3. Principal component analysis (PCA) plot showing gene expression data grouped according to the CTR, AA, and Igf-1 treatments.
Figure 3. Principal component analysis (PCA) plot showing gene expression data grouped according to the CTR, AA, and Igf-1 treatments.
Ijms 25 03894 g003
Figure 4. Volcano plots of differentially expressed mRNAs, lncRNAs, and miRNAs detected in gilthead sea bream myoblasts in response to the treatments. Volcano plots of transcription results of the AA vs. CTR group (AC) and Igf-1 vs. CTR group (DF) for mRNAs (A,D), miRNAs (B,E), and lncRNAs (C,F). Red vertical lines represent log2FC of 1 and −1. Red horizontal lines represent a p-value of 0.05.
Figure 4. Volcano plots of differentially expressed mRNAs, lncRNAs, and miRNAs detected in gilthead sea bream myoblasts in response to the treatments. Volcano plots of transcription results of the AA vs. CTR group (AC) and Igf-1 vs. CTR group (DF) for mRNAs (A,D), miRNAs (B,E), and lncRNAs (C,F). Red vertical lines represent log2FC of 1 and −1. Red horizontal lines represent a p-value of 0.05.
Ijms 25 03894 g004
Figure 5. Venn diagrams of mRNAs, miRNAs, and lncRNAs significantly modified by AA and Igf-1. Venn diagrams showing the number of mRNAs, miRNAs, and lncRNAs upregulated (AC) and downregulated (DF) in response to the treatments. The numbers inside the blue bubbles and red bubbles represent the number of mRNAs (A,D), miRNAs (B,E), and lncRNAs (C,F) uniquely changed in response to AA and Igf-1, respectively. The number in the intersection of the two bubbles indicates the mRNAs, miRNAs, and lncRNAs that commonly changed in response to both treatments.
Figure 5. Venn diagrams of mRNAs, miRNAs, and lncRNAs significantly modified by AA and Igf-1. Venn diagrams showing the number of mRNAs, miRNAs, and lncRNAs upregulated (AC) and downregulated (DF) in response to the treatments. The numbers inside the blue bubbles and red bubbles represent the number of mRNAs (A,D), miRNAs (B,E), and lncRNAs (C,F) uniquely changed in response to AA and Igf-1, respectively. The number in the intersection of the two bubbles indicates the mRNAs, miRNAs, and lncRNAs that commonly changed in response to both treatments.
Ijms 25 03894 g005
Figure 6. Global Gene Ontology (GO) enrichment analysis of the genes that significantly changed their transcription in response to AA (A) or Igf-1 (B). The size of the dots represents the number of genes present in each GO term, while the color indicates the p-value associated with each GO term identified. The name of the enriched GO term is indicated on the left side of the panel, whereas the GO Rich Factor (ratio of the number of differentially expressed genes in the pathway to the total number of genes in the pathway) is indicated in the lower part of each panel.
Figure 6. Global Gene Ontology (GO) enrichment analysis of the genes that significantly changed their transcription in response to AA (A) or Igf-1 (B). The size of the dots represents the number of genes present in each GO term, while the color indicates the p-value associated with each GO term identified. The name of the enriched GO term is indicated on the left side of the panel, whereas the GO Rich Factor (ratio of the number of differentially expressed genes in the pathway to the total number of genes in the pathway) is indicated in the lower part of each panel.
Ijms 25 03894 g006
Figure 7. Global Gene Ontology (GO) enrichment analysis of the miRNAs that significantly changed their transcription in response to AA (A) or Igf-1 (B). The size of the dots represents the number of genes present in each GO term, while the color indicates the p-value associated with each GO term identified. The name of the enriched GO term is indicated on the left side of the panel, whereas the GO Rich Factor (ratio of the number of differentially expressed genes in the pathway to the total number of genes in the pathway) is indicated in the lower part of each panel.
Figure 7. Global Gene Ontology (GO) enrichment analysis of the miRNAs that significantly changed their transcription in response to AA (A) or Igf-1 (B). The size of the dots represents the number of genes present in each GO term, while the color indicates the p-value associated with each GO term identified. The name of the enriched GO term is indicated on the left side of the panel, whereas the GO Rich Factor (ratio of the number of differentially expressed genes in the pathway to the total number of genes in the pathway) is indicated in the lower part of each panel.
Ijms 25 03894 g007
Table 1. Gene Ontology analysis of the up and downregulated genes in response to AA and Igf-1.
Table 1. Gene Ontology analysis of the up and downregulated genes in response to AA and Igf-1.
AA vs. CTR
GO TermDescriptionFDR
UpregulatedBiological Process0007049Cell cycle5.87 × 10−45
0006260DNA replication2.12 × 10−19
0007010Cytoskeleton organization4.82 × 10−08
0003012Muscle system process1.87 × 10−07
0042692Muscle cell differentiation3.52 × 10−05
0045214Sarcomere organization4.19 × 10−07
Molecular Function0008092Cytoskeletal protein binding1.61 × 10−10
0003688DNA replication origin binding3.47 × 10−07
0005515Protein binding8.75 × 10−06
0005524ATP binding0.0074
0016787Hydrolase activity0.026
Cellular Component0043232Intracellular non-membrane-bounded organelle2.53 × 10−30
0043292Contractile fiber2.15 × 10−18
0030017Sarcomere6.87 × 10−17
0005654Nucleoplasm0.0279
DownregulatedBiological Process0032870Cellular response to hormone0.0043
0043473Pigmentation0.008
0034219Carbohydrate transmembrane transport0.0215
Molecular Function0015293Symporter activity0.0088
0008083Growth factor activity0.0430
0005125Cytokine activity0.0430
0005539Glycosaminoglycan binding0.0430
Cellular Component0005576Extracellular region0.0110
0110165Cellular anatomical entity0.0110
0031082BLOC complex0.0416
Igf-1 vs. CTR
UpregulatedBiological Process0042692Muscle cell differentiation0.0026
0055001Muscle cell development0.0120
0061061Muscle structure development0.0120
0009987Cellular process0.0120
Cellular component0030016Myofibril0.00029
0030017Sarcomere0.00029
0099512Supramolecular fiber0.00029
0015629Actin cytoskeleton0.0074
DownregulatedMolecular Function0005539Glycosaminoglycan binding0.0250
Cellular Component0005576Extracellular region0.0003
FDR: False discovery rate.
Table 2. Gene Ontology enrichment analysis of the up and downregulated genes that were predicted to correlate and interact with miRNAs.
Table 2. Gene Ontology enrichment analysis of the up and downregulated genes that were predicted to correlate and interact with miRNAs.
GO TermDescriptionFDR
AA vs. CTR
UpregulatedBiological Process0009888Tissue development0.034
0030163Protein catabolic process0.034
0097435Actin cytoskeleton organization0.034
0006260DNA replication0.034
Molecular Function0004298Threonine-type endopeptidase activity0.001
Cellular Component0005622Intracellular0.0007
0032991Protein-containing complex0.0086
DownregulatedBiological Process0043473Pigmentation0.0002
0019262N-acetylneuraminate catabolic process0.026
Molecular Function0016798Hydrolase activity, acting on glycosyl bonds0.031
0005520Insulin-like growth factor binding0.041
Cellular Component0110165Cellular anatomical entity0.010
0012505Endomembrane system0.039
0005773Vacuole0.000
Igf-1 vs. CTR
UpregulatedBiological Process0048731System development0.045
0006259DNA metabolic process0.045
0055001Muscle cell development0.008
Cellular Component0005856Cytoskeleton0.019
0030017Sarcomere0.000
DownregulatedCellular Component0005576Extracellular regions0.000
FDR: False discovery rate.
Table 3. Potential lncRNAs acting as miRNAs sponges. Predicted interactions between lncRNAs and miRNAs significantly modified in response to AA and Igf-1.
Table 3. Potential lncRNAs acting as miRNAs sponges. Predicted interactions between lncRNAs and miRNAs significantly modified in response to AA and Igf-1.
lncRNAs IDmiRNAsCorrelation IndexEnergy (ndG)
ENSSAUG00010001802miR-27a; miR-29d; miR-29b−0.88; −0.86; −0.87−29.1; −26.2; −27.1
ENSSAUG00010017848miR-122; miR-92a; miR-29a; miR-29d;
miR-29b; miR-203a; miR-25; miR-31
−0.88; −0.80; −0.86; −0.92;
−0.84; −0.91; −0.84; −0.81
−26.7; −32.5; −32.2; −32.2;
−26.6; −28.1; −27.7; −28.6
ENSSAUG00010024948miR-122; miR-92a; miR-10c; miR-10d;
miR-27a; miR-29b; miR-31
−0.92; −0.89; −0.87; −0.87;
−0.82; −0.90; −0.93
−26.0; −27.6; −25.5; −25.5;
−30.6; −32.3; −26.0
ENSSAUG00010012228miR-338; miR-133a; miR-133b; miR-206; miR-17a; miR-125a; miR-106; mir-217−0.80; −0.92; −0.91; −0.87;
−0.80; −0.82; −0.90; −0.91
−27.1; −26.1; −26.1; −28.6;
−27.8; −28.6; −28.1; −29.0
ENSSAUG00010000237miR-125b−0.83−28.5
ENSSAUG00010012182miR-7a; miR-338; miR-133a; miR-133b; miR-206; miR-106; miR-17a; miR-125a−0.93; −0.80; −0.92; −0.91;
−0.87; −0.90; −0.80; −0.82
−29.0; −27.1; −26.1; −26.1;
−28.6; −28.6; −27.8; −28.6
ENSSAUG00010012549miR-17a−0.80−26.5
ENSSAUG00010015941miR-206; miR-17a; miR-125b; mir-145;
miR-454
−0.86; −0.83; −0.84; −0.83;
−0.86
−26.4; −30.2; −30.2; −25.6;
−29.7
ENSSAUG00010016074miR-15a; miR-19b; miR-217; miR-34−0.89; −0.81; −0.85; −0.81−27; −25.3; −27.6; −27.5
ENSSAUG00010016143miR-133a−0.82−25.1
ENSSAUG00010017089miR-206; miR-106; miR-128; miR-17a−0.95; −0.82; −0.88; −0.91−28.6; −26.8; −30.7; −27.8
ENSSAUG00010016280miR-122; miR-92a; miR-25−0.86; −0.81; −0.80−29.6; −27.8; −25.1
ENSSAUG00010002983miR-15a−0.90−31.2
ENSSAUG00010008657miR-338; miR-15a; miR-34; miR-7147−0.88; −0.85; −0.88; −0.90−33.3; −29.4; −25.6; −25.8
ENSSAUG00010022074miR-128; miR-365; miR-454; miR-19a;
miR-15a; miR-34; miR-7147
−0.83; −0.83; −0.82; −0.82;
−0.80; −0.83; −0.86
−31.3; −29.7; −25.9; −27.5;
−32.9; −28.6; −27.2
ENSSAUG00010013187miR-30e; miR-29a; miR-29d; miR-22b;
miR-30a
−0.91; −0.90; −0.83; −0.80;
−0.90
−26.8; −28.8; −28.8; −28.3;
−29.5
ENSSAUG00010013622miR-30e; miR-29d; miR-8160ba; miR-30a−0.85; −0.84; −0.90; −0.82−27.0; −27.3; −25.2; −26.5;
ENSSAUG00010015504miR-27d; miR-30a−0.85; −0.81−29.1; −30.1
ENSSAUG00010016109miR-30e; miR-25; miR-27d; miR-27a−0.81; −0.82; −0.84; −0.85−31.7; −25.8; −25.5; −26.0
ENSSAUG00010001416miR-29b−0.83−25.2
ENSSAUG00010017066let-7g−0.81−28.0
ENSSAUG00010002786miR-10926; miR-29d; miR-8160ba−0.80; −0.83; −0.97−28.1; −28.7; −26.3
ENSSAUG00010004711miR-8160ba−0.89−27.1
ENSSAUG00010026349miR-10926; miR-22b; miR-29a; miR-29d; miR-551; miR-8160ba−0.81; −0.82; −0.82; −0.82;
−0.81; −0.82
−29.4; −27; −26.2; −26.2;
−25.6; −25.6
ENSSAUG00010009596miR-128; mir-365; miR-7550−0.95; −0.90; −0.83−27.7; −27.3; −25.5
ENSSAUG00010015789miR-128; miR-365; miR-125b−0.89; −0.82; −0.81−28.9; −32.7; −25.2
ENSSAUG00010020704miR-128; miR-365; miR-26b; miR-454;
miR-19a; miR-15a; miR-34; miR-7147
−0.83; −0.83; −0.86; −0.82;
−0.82; −0.80; −0.83; −0.85
−31.3; −29.2; −26.3; −25.9;
−27.5; −32.9; −28.6; −27.2
ENSSAUG00010010920miR-139; miR-27d; miR-8160ba−0.81; −0.85; −0.83−28.5; −26.1; −25.9
ENSSAUG00010003663miR-15a; miR-301b; miR-33b; miR-34;
miR-7147
−0.91; −0.81; −0.81; −0.84;
−0.85
−29.1; −25.3; −26.1; −33.2;
−28.1
ENSSAUG00010016209miR-27a; miR-122; miR-92a−0.85; −0.86; −0.81−26.0; −29.6; −27.8
The predicted interactions between lncRNAs and miRNAs shown are based on transcriptional correlations and bioinformatics analysis. Interactions with Pearson correlations lower than −0.80 and with predicted interaction energies lower than −25.0 kcal/mol are shown.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

García-Pérez, I.; Duran, B.O.S.; Dal-Pai-Silva, M.; Garcia de la serrana, D. Exploring the Integrated Role of miRNAs and lncRNAs in Regulating the Transcriptional Response to Amino Acids and Insulin-like Growth Factor 1 in Gilthead Sea Bream (Sparus aurata) Myoblasts. Int. J. Mol. Sci. 2024, 25, 3894. https://doi.org/10.3390/ijms25073894

AMA Style

García-Pérez I, Duran BOS, Dal-Pai-Silva M, Garcia de la serrana D. Exploring the Integrated Role of miRNAs and lncRNAs in Regulating the Transcriptional Response to Amino Acids and Insulin-like Growth Factor 1 in Gilthead Sea Bream (Sparus aurata) Myoblasts. International Journal of Molecular Sciences. 2024; 25(7):3894. https://doi.org/10.3390/ijms25073894

Chicago/Turabian Style

García-Pérez, Isabel, Bruno Oliveira Silva Duran, Maeli Dal-Pai-Silva, and Daniel Garcia de la serrana. 2024. "Exploring the Integrated Role of miRNAs and lncRNAs in Regulating the Transcriptional Response to Amino Acids and Insulin-like Growth Factor 1 in Gilthead Sea Bream (Sparus aurata) Myoblasts" International Journal of Molecular Sciences 25, no. 7: 3894. https://doi.org/10.3390/ijms25073894

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