Skip to main content

Myostatin gene role in regulating traits of poultry species for potential industrial applications

Abstract

The myostatin (MSTN) gene is considered a potential genetic marker to improve economically important traits in livestock, since the discovery of its function using the MSTN knockout mice. The anti-myogenic function of the MSTN gene was further demonstrated in farm animal species with natural or induced mutations. In poultry species, myogenesis in cell culture was regulated by modulation of the MSTN gene. Also, different expression levels of the MSTN gene in poultry models with different muscle mass have been reported, indicating the conserved myogenic function of the MSTN gene between mammalian and avian species. Recent advances of CRISPR/Cas9-mediated genome editing techniques have led to development of genome-edited poultry species targeting the MSTN gene to clearly demonstrate its anti-myogenic function and further investigate other potential functions in poultry species. This review summarizes research conducted to understand the function of the MSTN gene in various poultry models from cells to whole organisms. Furthermore, the genome-edited poultry models targeting the MSTN gene are reviewed to integrate diverse effects of the MSTN gene on different traits of poultry species.

Introduction

Poultry species provide a major source of dietary proteins from their meat and eggs. Poultry production is affected by various physiological processes which are regulated and influenced by many endogenous and exogenous factors. To benefit current poultry production and other economically important traits, genetic factors are considered potential major targets for regulating and improving most desired features in poultry species.

Nutritional and environmental factors are external aspects that can be readily modulated and thus have been actively studied in diverse poultry research. Genetic factors, however, are not easily accessible to investigate due to scarce research models, especially in avian species, despite their importance on regulation of economically important traits. Studies for genetic factors often require specific animal models to correlate their phenotypes with the genes of interest [1]. Improvement of analytic methods and tools revealed the association between specific traits and genetic variants and mutations such as single nucleotide polymorphisms in poultry species [2,3,4]. However, further experimental demonstrations are still required to conclude that the phenotypes are regulated by specific genes.

Researchers are currently using genome editing techniques as an innovative approach to artificially lead to phenotypic alterations in organisms. The emergence of the CRISPR/Cas9 system has revolutionized genome editing in almost all organisms, including avian species [5, 6]. By targeting specific genes of interest, genome-edited animal models show phenotypic changes caused by the modulation of the genes. These distinct phenotypes between the mutant and wild-type (WT) animals enable us to investigate the function of the gene directly and readily. Indeed, the function of the myostatin (MSTN) gene was first discovered in MSTN knockout mice by homologous recombination in embryonic stem cells and injection of the genome-edited cells into blastocysts [7], indicating the importance of genome editing methods in functional studies of genetic research.

Due to the increased muscle mass caused by the MSTN mutations in various farm animals [8,9,10,11,12,13], the MSTN gene has been considered a potential genetic marker to improve meat production in industry. Moreover, other benefits, such as decreased fat deposition [14,15,16] and improved bone quality [17, 18], also suggest immense potential of the MSTN gene in industry. Notably, generation of the genome-edited quail and chickens targeting the MSTN gene [19, 20] facilitates diverse research investigating the gene function in avian species. In this review, we discussed previously reported studies focusing on investigation of the MSTN gene function in different cells and poultry breeds across different species. In particular, functional studies using the MSTN mutant quail and chicken provide deeper insights into potential usage of the MSTN gene to improve various economic traits and well-being of poultry species.

Myogenic function of the MSTN gene in poultry species

Cell culture studies

The function of the MSTN gene has been investigated in several avian cell culture systems. When recombinant myostatin was treated into embryonic myoblasts, it inhibited proliferation and differentiation in chickens [21] and proliferation in turkeys [22]. The anti-myogenic effect of MSTN was also confirmed in breast muscle derived satellite cells from chicken and turkey after hatching [22, 23]. In addition, overexpression of the MSTN gene also inhibited proliferation and differentiation of the chicken embryonic myoblasts, resulting in reduction of the total myotube areas [24]. Moreover, inhibition of the MSTN gene expression by shRNA promoted proliferation of the duck embryonic myoblasts [25]. Also, knockout and overexpression of the MSTN gene in the quail myoblast cell line resulted in an increase and decrease of myotube formation, respectively [26, 27], indicating regulation of myogenesis by the MSTN gene.

Different poultry breeds

As a top-down approach, different breeds of poultry species showing specific traits can be used to investigate whether the gene of interest plays a role in the specific traits of the breeds. Broilers are any chicken breeds that are intentionally selected and specifically raised for meat production [28], showing an excessive amount of muscle mass at their market ages, generally 6 weeks of age. Unexpectedly, the MSTN gene expression of broilers is not consistently lower than layers throughout pre- and post-hatching periods [29,30,31]. In one study, broilers showed increased MSTN expression at 2 weeks of age but decreased MSTN expression at 4 and 7 weeks of age [32]. Another study showed higher MSTN expression in broilers only at 5 days of age compared to layers [30]. Although broilers have higher body weight compared to layers since the hatching day [33], similar expression of the MSTN gene between broilers and layers throughout the embryonic stages, when myoblasts are formed, proliferated and differentiated [34], indicates that the MSTN gene might not be the major driving force of the different muscle mass between broilers and layers. Indeed, more than 50 years of genetic selection of broilers for better growth rate and body weight resulted in genetic variants between broilers and layers [35], suggesting compound effects of multiple genes for the higher muscle mass in broilers compared to layers.

However, the MSTN gene is still considered one of major genes regulating body weight and muscle mass in different poultry breeds. As a small-sized slow growing breed, Daweishan mini chickens showed higher expression of the MSTN gene compared to broilers [36], suggesting that the MSTN gene negatively regulates its body and muscle growth. Similarly, a quail breed selected for low body weight also showed higher MSTN expression compared to random bred control (RBC) quail at embryonic periods [37]. Additionally, a heavy weight quail line showed a decrease in the MSTN expression compared to RBC quail at pre- and early post-hatching periods [38]. The association between higher and lower MSTN expression with smaller and bigger body weight, respectively, in different poultry breeds indicates that the MSTN gene is one of main genes that regulate body weight and muscle mass in poultry species.

Genome-engineered poultry models

Genome engineering method targeting the gene of interest is a direct approach to investigate the gene function on phenotypic changes of the animals. Currently, multiple genome-engineered poultry models targeting the MSTN gene have been generated via different methods (Table 1). Before the emergence of the CRISPR/Cas9 system, the function of the MSTN gene was investigated by overexpressing shRNA to inhibit its expression in transgenic chickens generated using a sperm-mediated gene transfer method [39, 40]. Although total dressing percentage and breast muscle percentage of the knockdown group were not different from the control group, body weight and leg muscle dressing percentage were increased in the knockdown chickens [40]. In another study, the MSTN–B form, a truncated peptide which inhibits the anti-myogenic function of MSTN–A form, was overexpressed in transgenic quail, using the lentivirus-mediated method [41]. Because expression of the MSTN–B form was specific in leg muscle of the transgenic quail, not throughout their body, transgenic quail showed increased muscle mass in their leg [41], suggesting improved muscle mass by inhibition of MSTN activity in poultry species.

Table 1 Currently reported whole-body genome-engineered avian models targeting the myostatin (MSTN) gene

As avian genome editing studies have been accelerated by the CRISPR/Cas9 system [5], genome-edited quail [19] and chickens [20] targeting the MSTN gene were generated using the adenovirus-mediated method and PGC-mediated method, respectively. The MSTN mutant quail showed significantly increased body weight compared to WT quail after 2 weeks of age in females and 3 weeks of age in males [19]. Increased number of muscle fibers, called hyperplasia, was shown in breast and leg muscles of the male and female mutant group compared to those of the WT group. In chickens, total body weight was not significantly increased by the MSTN mutation, although the mutant group showed a significantly higher growth rate compared to the WT group [20]. Nevertheless, leg and wing muscles, but not breast muscle, of the MSTN mutant chickens were heavier compared to those of the WT chickens [20]. Interestingly, muscle fiber hyperplasia was shown in leg and wing muscles of male mutant chickens, whereas female mutant chickens showed an increase in muscle fiber size only in the leg muscle [20]. These results suggest that the MSTN gene might not only act via different cellular mechanisms among muscle tissues at different locations but also presents sexual dimorphism in regulation of muscle mass in chickens.

In addition to the whole muscle mass, muscle fiber type I, but not fiber type II, was enlarged by the MSTN mutation in male chicken wing muscle [20]. Nonetheless, the number of type I fibers of the wing muscle in the MSTN mutant group was decreased [20]. Indeed, it has been suggested that one of the mechanisms of the MSTN gene on regulating muscle mass is by switching muscle fiber types, from slow oxidative type I and fast oxidative glycolytic type IIa to fast glycolytic type IIb [48]. Type II muscle fibers are responsible for an increase in muscle size, called hypertrophy, in the MSTN knockout mice and pigs [49, 50]. In breast muscles of duck and quail, which have both type IIa and IIb, IIb muscle fibers are larger in size and smaller in number than IIa fibers, and located on the outer edge of the muscle bundles [51, 52]. Therefore, the muscle fiber type transition from type IIa to IIb in the deep region of the breast muscle might partially support the increase in muscle mass of the MSTN mutant quail [52]. However, chicken breast muscle contains muscle fiber type IIb only [51], potentially limiting the effect of the MSTN mutation on regulating muscle fiber types and breast muscle mass. Because breast muscle is the largest muscle in chickens, the diminished effect of the MSTN mutation on body weight in chickens can be partially explained by the similar breast muscle weight between the MSTN mutant and WT chickens. Duck breast muscle is similar to quail breast muscle in terms of muscle fiber types [51]. Although functional studies have not yet been performed on the recently reported MSTN mutant duck model [47], it is expected that phenotypic effects would be similar to the results of the quail, but not the chicken.

Additional traits examined in the genome-edited poultry species

Fat deposition and feed intake

In addition to the increased muscle mass, decreased fat deposition is another major phenotypic change caused by the MSTN mutation in various farm animals [14, 15, 53, 54], suggesting an additional benefit of the MSTN mutation for production of leaner meat. Lean meat is not only favorable for consumers who are looking for healthier foods, but also beneficial to the producer by increasing meat yield. Excessive fat accumulation in broiler chickens is one of major problems in the modern broiler industry, because it is associated with meat yield, feed efficiency [55], and mortality [56]. In addition, excessive fat accumulation can also negatively affect laying performance and health in broiler breeder hens [57, 58].

Like in other MSTN mutant animals, fat deposition was reduced by the MSTN mutation in quail as shown in lower leg and abdominal fat percentages in both male and female quail [42], although leg and abdominal fat weight was significantly different only in female quail [19]. Decreased abdominal fat deposition was also shown in the MSTN knockout male chickens [20] while fat deposition of female chickens was not affected by the MSTN knockout [20]. Male chickens are bigger than female chickens, whereas female quail are bigger than male quail which could increase energy requirements for growing and maintaining their bigger bodies. When the body and muscle weight further increased by the MSTN mutation, it might redistribute energy expenditure more toward muscle accretion but away from fat deposition during the post-hatch growth. Along with the opposite sexual dimorphism of decreased fat deposition by MSTN mutation between quail and chickens, potential interspecies difference in various effects of the MSTN mutation among poultry species should be also investigated.

In broilers, the high fat line showed higher feed conversion ratio (FCR) compared to the lower fat line [55], indicating feed efficiency is affected by fat deposition in poultry species. As expected from the decreased fat deposition, the MSTN mutant quail showed lower FCR compared to WT quail in both male and female [42], confirming improvement of feed efficiency by the MSTN mutation. FCR was calculated by measuring feed intake (FI) and body weight gain (WG) during the growth period before sexual maturation. Although WG was higher in the mutant group compared to the WT group at all growth phases, FI was increased by the MSTN mutation at early- and mid-growth phases but not at late-growth phase. This prolonged WG without increased FI at late-growth phase resulted in decreased FCR at overall growth period in the MSTN mutant male and female quail [42]. Because feeding is a major part of production cost, improving feed efficiency by MSTN mutation could significantly reduce poultry production cost.

Bone size and quality

Positive effect of the MSTN mutation on bone quality, demonstrated in MSTN knockout mice [17], can be another beneficial feature of the MSTN gene in the poultry industry, because poor bone qualities of both broilers and layers raise serious welfare and economic issues [59,60,61]. MSTN mutant pigs and rabbits also showed morphological changes in their rib and pelvic bones, respectively [62, 63], further supporting the potential role of the MSTN gene on bone formation. Because genetic factors are considered as major contributors to bone quality, along with nutritional and environmental factors, genetic approaches to improve traits related to bone qualities has been suggested as a potential solution [64]. In the MSTN mutant quail, bone size and quality were precisely analyzed by Micro-Computed Tomography scanning [43, 44]. Although there might be some indirect effect on bone size by increased body weight and muscle mass, significant increase in the size of tibia bones, including length, width, and cortical thickness, of both male and female quail by MSTN mutation supports the direct role of the MSTN gene on bone size in quail.

Other bone quality related parameters, such as bone mineral content, bone mineral density, bone volume, and bone breaking strength (BBS) were also higher in the MSTN mutant group compared to the WT group of adult male quail [43]. MSTN mutant female quail before sexual maturation also showed improvement in their bone quality related parameters [44]. However, the benefit of MSTN mutation on bone quality disappeared in female quail after sexual maturation, when most bone quality related parameters were similar between the two groups [44]. In fact, after the onset of egg laying, birds undergo dramatic changes to form a medullary cavity inside the long bone functioning as an accessible source of calcium for eggshell formation [65]. Notably, diaphyseal medullary and metaphyseal trabecular bones in MSTN mutant female quail had lower total surface value compared to WT female quail after sexual maturation. Because BBS of adult female quail between the 2 groups was similar, these negative values might be related to other physiological functions, such as calcium mobilization for eggshell formation, rather than bone strength.

Egg production and eggshell quality

Fecund egg laying of poultry species makes them prosperous animals in the egg production industry. The effect of MSTN mutation on egg production traits was further confirmed in MSTN mutant quail [45], after an SNP in the MSTN gene, G2283A, was reported to be associated with body weight, age of first egg laying, and egg numbers [66]. The eggs of MSTN mutant quail were bigger and heavier compared to those of WT quail [45, 46]. Because the positive association between hens’ body weight and their egg size was reported in chickens [67, 68], heavier body weight of the MSTN mutant quail might have an indirect effect on bigger egg size in the mutant group compared to the WT group. Notably, egg weight was bigger in the MSTN mutant group than the WT group even relatively to their body weight [69], suggesting a potential direct effect on the egg weight by MSTN mutation in addition to the increased body weight of the mutant female quail. However, total egg production (egg weight (g) multiplied by produced egg numbers during the selected periods within the actively laying period) was similar between the two groups, because of lower number of eggs laid by the mutant group than WT group during the period [45]. Also, the onset of egg laying, another economic trait of layers, was delayed in the MSTN mutant female quail. Increased fat accumulation before sexual maturation in chickens and quail [70, 71] is required prior to initial egg laying. Lower fat deposition in the mutant female compared to the WT female [19] might be one of the contributors to this delayed onset of egg laying in the mutant group. Nevertheless, similar fertility and hatchability between the two groups indicates that egg quality and embryonic development are not negatively affected by MSTN mutation [45].

Like egg size, eggshell strength is also positively associated with avian body weight [72]. However, eggshell strength of the mutant eggs, analyzed by breaking strength, was lower compared to that of WT eggs, mainly due to decreased eggshell thickness [46]. Reduction in the thickness of the palisade layer in the mutant group played a major role in making the eggshell thinner and weaker than the WT group [46], because a palisade layer is the thickest and major calcified layer in the eggshell [73]. In addition, a decreased expression level of MSTN in the uterus, the specific region of avian oviduct for eggshell formation, in egg-laying hens compared to non-laying hens at the same age further supports a potential regulatory role of the MSTN gene in eggshell formation [74].

Other considerations of the MSTN gene function in poultry species

Alternative splicing isoforms

Even though most of the research investigating expression of the MSTN gene reports the gene expression as a whole, 5 different isoforms of the MSTN have been reported in poultry species [27, 66] (Fig. 1). Among them, MSTN–A and MSTN–B forms are dominantly expressed in poultry species and have opposing functions on muscle growth [27]. MSTN–A is an intact form having the anti-myogenic effect on muscle growth as shown in reduced myotube length and diameter in quail myoblast cells after its overexpression [27]. On the contrary, overexpression of the MSTN–B form positively affected myogenesis of quail myoblast cells resulting in longer and thicker myotubes [27]. It was proposed that the MSTN–B form directly binds to the premature form of MSTN to inhibit the maturation process required for activation of MSTN [27]. In natural quail lines, the expression of the MSTN–B form was detectable after post-hatching d 28 [38]. The expression patterns of both forms correspond to the overall MSTN expression pattern of different lines of quail selected for their body weight [37, 38]. In chickens, however, both forms are detected even at embryonic ages and overall MSTN expression level does not represent the expression patterns of both forms [30]. Although MSTN expression as a whole was not different between broilers and layers at embryonic and post-hatching periods except for the early post-hatching day, the ratio of B form to A form was constantly higher in broilers compared to layers, except for the late embryonic day [30]. Higher expression of the B form in broilers compared to layers might suggest different genetic variation in broilers to inhibit MSTN activity and increase muscle mass. Thus, different expression patterns among MSTN isoforms can be another possible way to regulate muscle mass using the MSTN gene in poultry species.

Fig. 1
figure 1

Schematic representation showing reported alternative splicing isoforms of the MSTN gene in major poultry species. Five isoforms including MSTN–A, B, C, D, and E forms have been identified in chickens, quail, and turkeys. In ducks, MSTN–A, B, D, and E forms are reported, but not MSTN–C form

Environmental factors

Climate change is a real problem in our society and livestock industry [75, 76]. Global warming can increase environmental and heat stresses that need to be resolved for sustainable production [77, 78]. Appropriate temperature setting is an important factor for optimal growth in poultry species [79, 80], and expression of the MSTN gene is also affected by different temperature in chickens. Under constant heat stress (31 ± 1 °C) compared to thermoneutral control (21 ± 1 °C) during the growth period, it is reported that MSTN expression is elevated and growth performance is negatively affected in broilers at market age [81]. On the contrary, four-day-old chicken embryos exposed to hot temperatures (44 ± 0.5 °C) once for an hour decreased MSTN expression level immediately [82], altering the expression pattern of the MSTN gene during embryogenesis [83]. Interestingly, 1 week old chick exposed to 4 °C for a day increased breast muscle weight and decreased MSTN expression [84], suggesting a potential association between body temperature and expression of MSTN gene during embryonic and early post-hatching growth. In addition to the thermal stress, environmental stress coming from high stocking density needs to be managed for regulating MSTN expression properly. Under high stocking density conditions, broilers showed not only a decrease in average daily gain and breast muscle yield, but also an increase in the expression of the MSTN gene [85]. Therefore, indirect effects from various environmental stresses on avian growth performance through the MSTN expression should also be considered in an industrial setting.

Conclusions

Generation of genome-edited poultry models targeting the MSTN gene enabled researchers to investigate various functions of the MSTN gene in poultry species. In addition to confirmation of the reported effects of the MSTN mutation on muscle mass, fat deposition, feed efficiency, and bone quality in animals, novel discoveries on changed egg production traits and eggshell qualities indicate the importance of the MSTN mutant poultry models (Fig. 2). It is notable that most phenotypic changes caused by the MSTN mutation in quail and chickens are positively related to economic traits and health of meat producing poultry species. Also, an increase in muscle size of the MSTN mutant quail did not affect meat quality in terms of related traits, such as pH, redness, yellowness, and drip loss [52]. Because genome-edited farm animals have been lately approved for human consumption [86, 87], the MSTN gene can be seriously considered not only as a genetic marker but also for generation of a superior genome-edited line of poultry. Considering the high demand and importance of poultry meat in our society, comprehensive understanding of the MSTN gene in various traits of poultry species through genome-edited models and diverse research will greatly benefit not only poultry industry but also consumers.

Fig. 2
figure 2

Illustrative representation of the effects of MSTN mutation on various traits in quail. MSTN mutant quail showed increased muscle mass and growth rate, decreased fat deposition, and improved bone quality, growth rate, and feed efficiency. However, MSTN mutant female quail did not show improvement in egg production and bone quality at egg laying period. In addition, eggshell quality was decreased by the MSTN mutation

Availability of data and materials

Not applicable.

Abbreviations

BBS:

Bone breaking strength

FCR:

Feed conversion ratio

FI:

Feed intake

MSTN:

Myostatin

RBC:

Random bred control

WG:

Weight gain

WT:

Wild-type

References

  1. Tixier-Boichard M. From phenotype to genotype: major genes in chickens. Worlds Poult Sci J. 2002;58:65–75.

    Article  Google Scholar 

  2. Wang J, Lei QX, Cao DG, Zhou Y, Han HX, Liu W, et al. Whole genome SNPs among 8 chicken breeds enable identification of genetic signatures that underlie breed features. J Integr Agric. 2023;22:2200–12.

    Article  CAS  Google Scholar 

  3. Wang Q, Li D, Guo A, Li M, Li L, Zhou J, et al. Whole-genome resequencing of Dulong Chicken reveal signatures of selection. Br Poult Sci. 2020;61:624–31.

    Article  CAS  PubMed  Google Scholar 

  4. Jeong H, Kim K, Caetano-Anollés K, Kim H, Kim B-K, Yi J-K, et al. Whole genome sequencing of Gyeongbuk Araucana, a newly developed blue-egg laying chicken breed, reveals its origin and genetic characteristics. Sci Rep. 2016;6:26484.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Lee J, Kim DH, Lee K. Current approaches and applications in avian genome editing. Int J Mol Sci. 2020;21:3937.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Lee KY, Lee HJ, Choi HJ, Han ST, Lee KH, Park KJ, et al. Highly elevated base excision repair pathway in primordial germ cells causes low base editing activity in chickens. FASEB J. 2020;34:15907–21.

    Article  CAS  PubMed  Google Scholar 

  7. McPherron AC, Lawler AM, Lee SJ. Regulation of skeletal muscle mass in mice by a new TGF-beta superfamily member. Nature. 1997;387:83–90.

    Article  CAS  PubMed  Google Scholar 

  8. Grobet L, Jose L, Martin R, Poncelet D, Pirottin D, Brouwers B, et al. A deletion in the bovine myostatin gene causes the double-muscled phenotype in cattle. Nat Genet. 1997;17:71–4.

    Article  CAS  PubMed  Google Scholar 

  9. Kambadur R, Sharma M, Smith TPL, Bass JJ. Mutations in myostatin (GDF8) in double-muscled Belgian Blue and Piedmontese cattle. Genome Res. 1997;7:910–6.

    Article  CAS  PubMed  Google Scholar 

  10. Bi Y, Hua Z, Liu X, Hua W, Ren H, Xiao H, et al. Isozygous and selectable marker-free MSTN knockout cloned pigs generated by the combined use of CRISPR/Cas9 and Cre/LoxP. Sci Rep. 2016;6:31729.

    Article  Google Scholar 

  11. He Z, Zhang T, Jiang L, Zhou M, Wu D, Mei J, et al. Use of CRISPR/Cas9 technology efficiently targetted goat myostatin through zygotes microinjection resulting in double-muscled phenotype in goats. Biosci Rep. 2018;38(6):BSR20180742.

    Article  CAS  Google Scholar 

  12. Wang X, Niu Y, Zhou J, Yu H, Kou Q, Lei A, et al. Multiplex gene editing via CRISPR/Cas9 exhibits desirable muscle hypertrophy without detectable off-target effects in sheep. Sci Rep. 2016;6:32271.

    Google Scholar 

  13. Lv Q, Yuan L, Deng J, Chen M, Wang Y, Zeng J, et al. Efficient generation of myostatin gene mutated rabbit by CRISPR/Cas9. Sci Rep. 2016;6:25029.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Ren H, Xiao W, Qin X, Cai G, Chen H, Hua Z, et al. Myostatin regulates fatty acid desaturation and fat deposition through MEF2C/miR222/SCD5 cascade in pigs. Commun Biol. 2020;3:612.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. McPherron AC, Lee S-J. Suppression of body fat accumulation in myostatin-deficient mice. J Clin Invest. 2002;109:595–601.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Deng B, Zhang F, Wen J, Ye S, Wang L, Yang Y, et al. The function of myostatin in the regulation of fat mass in mammals. Nutr Metab (Lond). 2017;14:29.

    Article  PubMed  Google Scholar 

  17. Suh J, Kim N-K, Lee S-H, Eom J-H, Lee Y, Park J-C, et al. GDF11 promotes osteogenesis as opposed to MSTN, and follistatin, a MSTN/GDF11 inhibitor, increases muscle mass but weakens bone. Proc Natl Acad Sci U S A. 2020;117:4910–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Bialek P, Parkington J, Li X, Gavin D, Wallace C, Zhang J, et al. A myostatin and activin decoy receptor enhances bone formation in mice. Bone. 2014;60:162–71.

    Article  CAS  PubMed  Google Scholar 

  19. Lee J, Kim D-H, Lee K. Muscle hyperplasia in Japanese quail by single amino acid deletion in MSTN propeptide. Int J Mol Sci. 2020;21:1504.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Kim GD, Lee JH, Song S, Kim SW, Han JS, Shin SP, et al. Generation of myostatin-knockout chickens mediated by D10A-Cas9 nickase. FASEB J. 2020;34:5688–96.

    Article  CAS  PubMed  Google Scholar 

  21. Yang W, Wang K, Chen Y, Zhang Y, Huang B, Zhu D-H. Functional characterization of recombinant myostatin and its inhibitory role to chicken muscle development. Acta Biochimica et Biophysica Sinica (Shanghai). 2003;35:1016–22.

    CAS  PubMed  Google Scholar 

  22. McFarland DC, Velleman SG, Pesall JE, Liu C. Effect of myostatin on turkey myogenic satellite cells and embryonic myoblasts. Comp Biochem Physiol A Mol Integr Physiol. 2006;144:501–8.

    Article  PubMed  Google Scholar 

  23. McFarland DC, Velleman SG, Pesall JE, Liu C. The role of myostatin in chicken (Gallus domesticus) myogenic satellite cell proliferation and differentiation. Gen Comp Endocrinol. 2007;151:351–7.

    Article  CAS  PubMed  Google Scholar 

  24. Zhang G, He M, Wu P, Zhang X, Zhou K, Li T, et al. Microrna-27b-3p targets the myostatin gene to regulate myoblast proliferation and is involved in myoblast differentiation. Cells. 2021;10(2):423.

    CAS  Google Scholar 

  25. Tao Z, Zhu C, Song C, Song W, Ji G, Shan Y, et al. Lentivirus-mediated RNA interference of myostatin gene affects MyoD and Myf5 gene expression in duck embryonic myoblasts. Br Poult Sci. 2015;56:551–8.

    CAS  PubMed  Google Scholar 

  26. Park J-W, Lee JH, Han JS, Shin SP, Park TS. Muscle differentiation induced by p53 signaling pathway-related genes in myostatin-knockout quail myoblasts. Mol Biol Rep. 2020;47:9531–40.

    Article  CAS  PubMed  Google Scholar 

  27. Shin S, Song Y, Ahn J, Kim E, Chen P, Yang S, et al. A novel mechanism of myostatin regulation by its alternative splicing variant during myogenesis in avian species. Am J Physiol Cell Physiol. 2015;309:C650–9.

    Article  PubMed  Google Scholar 

  28. Zuidhof MJ, Schneider BL, Carney VL, Korver DR, Robinson FE. Growth, efficiency, and yield of commercial broilers from 1957, 1978, and 2005. Poult Sci. 2014;93:2970–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Mott I, Ivarie R. Expression of myostatin is not altered in lines of poultry exhibiting myofiber hyper- and hypoplasia. Poult Sci. 2002;81:799–804.

    Article  CAS  PubMed  Google Scholar 

  30. Kim D-H, Choi YM, Lee J, Shin S, Kim S, Suh Y, et al. Differential expression of MSTN isoforms in muscle between broiler and layer chickens. Animals. 2022;12:539.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Scheuermann GN, Bilgili SF, Tuzun S, Mulvaney DR. Comparison of chicken genotypes: myofiber number in pectoralis muscle and myostatin ontogeny. Poult Sci. 2004;83:1404–12.

    Article  CAS  PubMed  Google Scholar 

  32. Bhattacharya TK, Chatterjee RN, Dushyanth K, Shukla R. Cloning, characterization and expression of myostatin (growth differentiating factor-8) gene in broiler and layer chicken (Gallus gallus). Mol Biol Rep. 2015;42:319–27.

    Article  CAS  PubMed  Google Scholar 

  33. Druyan S. The effects of genetic line (broilers vs. layers) on embryo development. Poult Sci. 2010;89:1457–67.

    Article  CAS  PubMed  Google Scholar 

  34. Miller JB, Stockdale FE. Developmental origins of skeletal muscle fibers: clonal analysis of myogenic cell lineages based on expression of fast and slow myosin heavy chains. Proc Natl Acad Sci U S A. 1986;83:3860–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Iqbal MA, Hadlich F, Reyer H, Oster M, Trakooljul N, Murani E, et al. RNA-Seq-based discovery of genetic variants and allele-specific expression of two layer lines and broiler chicken. Evol Appl. 2023;16:1135–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Dou T, Li Z, Wang K, Liu L, Rong H, Xu Z, et al. Regulation of myostatin expression is associated with growth and muscle development in commercial broiler and DMC muscle. Mol Biol Rep. 2018;45:511–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Kim D-H, Choi YM, Suh Y, Shin S, Lee J, Hwang S, et al. Research Note: increased myostatin expression and decreased expression of myogenic regulatory factors in embryonic ages in a quail line with muscle hypoplasia. Poult Sci. 2021;100:100978.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Kim D-H, Choi YM, Suh Y, Shin S, Lee J, Hwang S, et al. Research Note: association of temporal expression of myostatin with hypertrophic muscle growth in different Japanese quail lines. Poult Sci. 2020;99:2926–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Bhattacharya TK, Shukla R, Chatterjee RN, Dushyanth K. Knock down of the myostatin gene by RNA interference increased body weight in chicken. J Biotechnol. 2017;241:61–8.

    Article  CAS  PubMed  Google Scholar 

  40. Bhattacharya TK, Shukla R, Chatterjee RN, Bhanja SK. Comparative analysis of silencing expression of myostatin (MSTN) and its two receptors (ACVR2A and ACVR2B) genes affecting growth traits in knock down chicken. Sci Rep. 2019;9:7789.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Chen PR, Suh Y, Shin S, Woodfint RM, Hwang S, Lee K. Exogenous expression of an alternative splicing variant of myostatin prompts leg muscle fiber hyperplasia in Japanese quail. Int J Mol Sci. 2019;20:4617.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Lee J, Kim D-H, Brower AM, Schlachter I, Lee K. Research Note: Improved feed efficiency in quail with targeted genome editing in the myostatin gene. Poult Sci. 2021;100:101257.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Lee J, Tompkins Y, Kim D-H, Kim WK, Lee K. Increased sizes and improved qualities of tibia bones by myostatin mutation in Japanese quail. Front Physiol. 2023;13:1085935.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Lee J, Tompkins Y, Kim D-H, Kim WK, Lee K. The effects of myostatin mutation on the tibia bone quality in female Japanese quail before and after sexual maturation. Poult Sci. 2023;102:102734.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Lee J, Kim D-H, Brower AM, Schlachter I, Lee K. Effects of myostatin mutation on onset of laying, egg production, fertility, and hatchability. Animals. 2021;11(7):1935.

    Article  Google Scholar 

  46. Lee J, McCurdy C, Chae C, Hwang J, Karolak MC, Kim D-H, et al. Myostatin mutation in Japanese quail increased egg size but reduced eggshell thickness and strength. Animals. 2021;12:47.

    Article  PubMed  PubMed Central  Google Scholar 

  47. Lee J, Kim D-H, Karolak MC, Shin S, Lee K. Generation of genome-edited chicken and duck lines by adenovirus-mediated in vivo genome editing. Proc Natl Acad Sci U S A. 2022;119:e2214344119.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Hennebry A, Berry C, Siriett V, O’Callaghan P, Chau L, Watson T, et al. Myostatin regulates fiber-type composition of skeletal muscle by regulating MEF2 and MyoD gene expression. Am J Physiol Cell Physiol. 2009;296:C525–34.

    Article  CAS  PubMed  Google Scholar 

  49. Girgenrath S, Song K, Whittemore L-A. Loss of myostatin expression alters fiber-type distribution and expression of myosin heavy chain isoforms in slow- and fast-type skeletal muscle. Muscle Nerve. 2005;31:34–40.

    Article  CAS  PubMed  Google Scholar 

  50. Qian LL, Xie JY, Gao T, Cai CB, Jiang SW, Bi HF, et al. Targeted myostatin loss-of-function mutation increases type II muscle fibers in Meishan pigs. J Integr Agric. 2022;21:188–98.

    Article  CAS  Google Scholar 

  51. Ismail I, Joo S-T. Poultry meat quality in relation to muscle growth and muscle fiber characteristics. Korean J Food Sci Anim Resour. 2017;37:873–83.

    PubMed  PubMed Central  Google Scholar 

  52. Kim D-H, Lee B, Lee J, Bohrer BM, Choi YM, Lee K. Effects of a myostatin mutation in Japanese quail (Coturnix japonica) on the physicochemical and histochemical characteristics of the pectoralis major muscle. Front Physiol. 2023;14:1172884.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Boman IA, Klemetsdal G, Nafstad O, Blichfeldt T, Våge DI. Impact of two myostatin (MSTN) mutations on weight gain and lamb carcass classification in Norwegian White Sheep (Ovis aries). Genet Sel Evol. 2010;42:4.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Fiems LO. Double muscling in cattle: genes, husbandry, carcasses and meat. Animals. 2012;2:472–506.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Chen C, Su Z, Li Y, Luan P, Wang S, Zhang H, et al. Estimation of the genetic parameters of traits relevant to feed efficiency: result from broiler lines divergent for high or low abdominal fat content. Poult Sci. 2021;100:461–6.

    Article  PubMed  Google Scholar 

  56. Lippens M, Room G, De Groote G, Decuypere E. Early and temporary quantitative food restriction of broiler chickens. 1. Effects on performance characteristics, mortality and meat quality. Br Poult Sci. 2000;41:343–54.

    Article  CAS  PubMed  Google Scholar 

  57. Ma Y, Cheng B, Zhou S, Wang Y, Jing Y, Leng L, et al. Comparative analyses of laying performance and follicular development characteristics between fat and lean broiler lines. Poult Sci. 2024;103:103250.

    Article  PubMed  Google Scholar 

  58. Chen CY, Huang YF, Ko YJ, Liu YJ, Chen YH, Walzem RL, et al. Obesity-associated cardiac pathogenesis in broiler breeder hens: development of metabolic cardiomyopathy. Poult Sci. 2017;96:2438–46.

    Article  CAS  PubMed  Google Scholar 

  59. Julian RJ. Production and growth related disorders and other metabolic diseases of poultry–a review. Vet J. 2005;169:350–69.

    Article  CAS  PubMed  Google Scholar 

  60. Tompkins YH, Teng P, Pazdro R, Kim WK. Long bone mineral loss, bone microstructural changes and oxidative stress after eimeria challenge in broilers. Front Physiol. 2022;13:945740.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Webster AB. Welfare implications of avian osteoporosis. Poult Sci. 2004;83:184–92.

    Article  CAS  PubMed  Google Scholar 

  62. Qian L, Tang M, Yang J, Wang Q, Cai C, Jiang S, et al. Targeted mutations in myostatin by zinc-finger nucleases result in double-muscled phenotype in Meishan pigs. Sci Rep. 2015;5:14435.

    Article  Google Scholar 

  63. Zhang T, Lu Y, Song S, Lu R, Zhou M, He Z, et al. “Double-muscling” and pelvic tilt phenomena in rabbits with the cystine-knot motif deficiency of myostatin on exon 3. Biosci Rep. 2019;39:BSR20190207.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Fleming RH, McCormack HA, McTeir L, Whitehead CC. Relationships between genetic, environmental and nutritional factors influencing osteoporosis in laying hens. Br Poult Sci. 2006;47:742–55.

    Article  CAS  PubMed  Google Scholar 

  65. Whitehead CC. Overview of bone biology in the egg-laying hen. Poult Sci. 2004;83:193–9.

    Article  CAS  PubMed  Google Scholar 

  66. Huang KL, Wang JW, Han CC, Liu HH, Li L, Dai F, et al. Developmental expression and alternative splicing of the duck myostatin gene. Comp Biochem Physiol Part D Genomics Proteomics. 2011;6(3):238–43.

    Article  CAS  PubMed  Google Scholar 

  67. Bish CL, Beane WL, Ruszler PL, Cherry JA. Body weight influence on egg production. Poult Sci. 1985;64:2259–62.

    Article  Google Scholar 

  68. Leeson S, Summers JD. Effect of immature body weight on laying performance. Poult Sci. 1987;66:1924–8.

    Article  CAS  PubMed  Google Scholar 

  69. Kim D-H, Lee J, Choi YM, Lee K. Research Note: Growth promoting potential in Mstn mutant quail dependent and independent of increased egg size. Poult Sci. 2023;102:102260.

    Article  CAS  PubMed  Google Scholar 

  70. Szentirmai E, Milisits G, Donkó T, Budai Z, Ujvári J, Fülöp T, et al. Comparison of changes in production and egg composition in relation to in vivo estimates of body weight and composition of brown and white egg layers during the first egg-laying period. Br Poult Sci. 2013;54:587–93.

    Article  CAS  PubMed  Google Scholar 

  71. Oruwari BM, Cherry JA, Jones DE, Beane WL. Lipocyte hyperplasia and sexual maturation of Japanese quail ( Coturnix coturnix japonica ). Br J Nutr. 1986;55:621–30.

    Article  CAS  PubMed  Google Scholar 

  72. Le Roy N, Stapane L, Gautron J, Hincke MT. Evolution of the avian eggshell biomineralization protein toolkit – new insights from multi-omics. Front Genet. 2021;12:672433.

    Google Scholar 

  73. Gautron J, Stapane L, Le Roy N, Nys Y, Rodriguez-Navarro AB, Hincke MT. Avian eggshell biomineralization: an update on its structure, mineralogy and protein tool kit. BMC Mol Cell Biol. 2021;22:11.

    Google Scholar 

  74. Wasti S, Sah N, Kuehu DL, Kim YS, Jha R, Mishra B. Expression of follistatin is associated with egg formation in the oviduct of laying hens. Anim Sci J. 2020;91:e13396.

    Article  CAS  PubMed  Google Scholar 

  75. Godde CM, Mason-D’Croz D, Mayberry DE, Thornton PK, Herrero M. Impacts of climate change on the livestock food supply chain; a review of the evidence. Glob Food Sec. 2021;28:100488.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Filho WL, Setti AFF, Azeiteiro UM, Lokupitiya E, Donkor FK, Etim NN, et al. An overview of the interactions between food production and climate change. Sci Total Environ. 2022;838:156438.

    Article  CAS  PubMed  Google Scholar 

  77. Oke OE, Uyanga VA, Iyasere OS, Oke FO, Majekodunmi BC, Logunleko MO, et al. Environmental stress and livestock productivity in hot-humid tropics: alleviation and future perspectives. J Therm Biol. 2021;100:103077.

    Article  CAS  PubMed  Google Scholar 

  78. Renaudeau D, Collin A, Yahav S, de Basilio V, Gourdine JL, Collier RJ. Adaptation to hot climate and strategies to alleviate heat stress in livestock production. Animal. 2012;6:707–28.

    Article  CAS  PubMed  Google Scholar 

  79. Oluwagbenga EM, Tetel V, Schober J, Fraley GS. Chronic heat stress part 2: Increased stress and fear responses in F1 Pekin ducks raised from parents that experienced heat stress. Animals. 2023;13:1748.

    Article  PubMed  PubMed Central  Google Scholar 

  80. Mangan M, Siwek M. Strategies to combat heat stress in poultry production-a review. J Anim Physiol Anim Nutr (Berl). 2023. https://doi.org/10.1111/jpn.13916.

    Article  PubMed  Google Scholar 

  81. Li X, Zhang M, Feng J, Zhou Y. Myostatin and related factors are involved in skeletal muscle protein breakdown in growing broilers exposed to constant heat stress. Animals. 2021;11:1467.

    Article  PubMed  PubMed Central  Google Scholar 

  82. Gabriel JE, Alvares LE, Gobet MC, De Paz CCP, Packer IU, Macari M, et al. Expression of MyoD, myogenin, myostatin and Hsp70 transcripts in chicken embryos submitted to mild cold or heat. J Therm Biol. 2003;28:261–9.

    Article  CAS  Google Scholar 

  83. Kocamis H, Kirkpatrick-Keller DC, Richter J, Killefer J. The ontogeny of myostatin, follistatin and activin-B mRNA expression during chicken embryonic development. Growth Dev Aging. 1999;63:143–50.

    CAS  PubMed  Google Scholar 

  84. Ijiri D, Kanai Y, Hirabayashi M. Possible roles of myostatin and PGC-1alpha in the increase of skeletal muscle and transformation of fiber type in cold-exposed chicks: expression of myostatin and PGC-1α in chicks exposed to cold. Domest Anim Endocrinol. 2009;37(1):12–22.

    Article  CAS  PubMed  Google Scholar 

  85. Li XM, Zhang MH, Liu SM, Feng JH, Ma DD, Liu QX, et al. Effects of stocking density on growth performance, growth regulatory factors, and endocrine hormones in broilers under appropriate environments. Poult Sci. 2019;98:6611–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. CRISPR beef cattle get FDA green light. Nat Biotechnol. 2022;40:448. https://doi.org/10.1038/s41587-022-01297-z.

  87. Dolgin E. First GM pigs for allergies. Could xenotransplants be next? Nat Biotechnol. 2021;39(4):397–400.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank Michelle Milligan for proofreading this manuscript.

Funding

This work was funded by the United States Department of Agriculture National Institute of Food and Agriculture Grant (Project No. 2020-67030-31338).

Author information

Authors and Affiliations

Authors

Contributions

JL and KL: Conceptualization; JL and D-HK: Visualization, Writing – Original Draft, Review & Editing; KL: Funding, Supervision, Writing – Review & Editing. All authors have read and agreed to the published version of the manuscript.

Corresponding author

Correspondence to Kichoon Lee.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lee, J., Kim, DH. & Lee, K. Myostatin gene role in regulating traits of poultry species for potential industrial applications. J Animal Sci Biotechnol 15, 82 (2024). https://doi.org/10.1186/s40104-024-01040-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40104-024-01040-5

Keywords