Next Article in Journal
Clinical Characteristics of Molecularly Defined Renal Cell Carcinomas
Next Article in Special Issue
Molecular Detection of Nosema spp. in Three Eco Regions of Slovakia
Previous Article in Journal
Selective Concurrence of the Long Non-Coding RNA MALAT1 and the Polycomb Repressive Complex 2 to Promoter Regions of Active Genes in MCF7 Breast Cancer Cells
Previous Article in Special Issue
Acceleration of the Deamination of Cytosine through Photo-Crosslinking
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polymorphism of Genes and Their Impact on Beef Quality

Institute of Animal Science, Warsaw University of Life Sciences, Ciszewskiego 8, 02-786 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Curr. Issues Mol. Biol. 2023, 45(6), 4749-4762; https://doi.org/10.3390/cimb45060302
Submission received: 7 May 2023 / Revised: 24 May 2023 / Accepted: 26 May 2023 / Published: 30 May 2023
(This article belongs to the Collection Feature Papers in Current Issues in Molecular Biology)

Abstract

:
The single-nucleotide polymorphism (SNP) form of genes is a valuable source of information regarding their suitability for use as specific markers of desirable traits in beef cattle breeding. For several decades, breeding work focused on improving production efficiency through optimizing the feed conversion ratio and improving daily gains and meat quality. Many research teams previously undertook research work on single-nucleotide polymorphism in myostatin (MSTN), thyroglobulin (TG), calpain (CAPN), and calpastatin (CAST) proteins. The literature review focuses on the most frequently addressed issues concerning these genes in beef cattle production and points to a number of relevant studies on the genes’ polymorphic forms. The four genes presented are worth considering during breeding work as a set of genes that can positively influence productivity and production quality.

1. Introduction

Population growth and the enrichment of many countries are increasing the demand for quality products [1]. One of these is beef, which is an important part of many people’s diets. A number of research teams previously undertook work to study the determinants of beef quality and looked for a way to implement the results of their research. This paper undertakes the task of reviewing the current state of knowledge about four candidate genes and the determination of their impact on beef quality. The genes highlighted in this work are myostatin (MSTN), thyroglobulin 5 (TG5), u-calpain (CAPN1), and calpastatin (CAST). Beef has long been the third most consumed meat after poultry and pork [2], with this trend due to its high price and, thus, higher consumer demand for quality [3]. In developed countries, it probably previously reached its peak per capita consumption, and for ethical reasons and due to environmental concerns, its consumption is now slightly decreasing [4]. Customers in highly developed countries prefer meat with a lower fat content [5,6]; this contradicts their taste choices in blind tests, in which they stated that they preferred the taste of beef with a higher fat content [7]. Beef with a higher intramuscular fat (IMF) content is also more nutritious [8]. Polymorphic forms of TG5 gene can increase intramuscular fat content by 6.5% [9] and can be used to further improve of cattle performance. This state of affairs, therefore, is a challenge for breeders, who must meet the demand of customers whose actual preferences seem to contradict the choices they make. Null mutation in MSTN results in 20–25% lower muscle mass in the Belgian Blue breed [10], which makes a significant difference in terms of productivity. To increase meat tenderness, work on CAPN1 and CAST, which improve meat quality, should be carried out at the same time [11]. Genomic selection is extremely important due to its effectiveness in the process of improving meat quality. Juiciness, color, tenderness, and water-holding capacity are important elements of beef quality [12]. Meat production worldwide exceeded 337 million tons in 2020 [13], and this growth was accompanied by an increasing interest in higher quality products [14,15]. It is, therefore, necessary to conduct work on candidate genes, which can be an important tool in shaping breeding programs and will expand our knowledge about important issues, such as production efficiency and meat quality. The literature was selected based on keywords related to the topic of the paper in several bibliographic databases of the Warsaw University of Life Sciences, including Web of Science and Scopus.

2. Myostatin

Myostatin (MSTN), also known as GDF8 (growth and differentiation factor 8), is one of the most important regulators of skeletal muscle development [16]. It is a highly conserved gene that can be identified in many mammal livestock species [10]. This gene plays a crucial role in muscle size development [17]. Huang et al. reported that a lack of MSTN activity resulted in the overgrowth of skeletal muscles—this issue is called the double muscling (DM) trait [18]. The double muscling phenotype is described using a number of symbols: DM or N, DM or dm, D or n, C or N, A or a, and mh or + [19]. It is highly desired among cattle producers due to its positive impact on the meat content of carcasses. MSTN can generate both a significantly higher proportion of skeletal muscle on the carcasses of slaughtered animals and the expression of adipose tissue through inhibiting or promoting adipogenesis [20]. Muscles can be as much as 20–25% larger than those of individuals without the mutation [10], while a decrease in the proportion of organ mass [21,22] and a reduced proportion of fat on the carcass are also associated traits [23].
All members of the TGF -β (transforming growth factor -β) family are characterized by three distinct domains: the N-terminal signal domain, the C-terminal mature peptide, and the propeptide domain [10]. In cattle, these trait were mapped at chromosome 2 [24]. Other similar characteristics, such as the hydrophobic core of amino acids near the N-terminal signal domain, cysteine residues in the C terminal region, and the conserved RSRR proteolytic processing signal at the C-terminus, indicate that MSTN is a member of this family [25]. However, there is a difference that distinguishes MSTN from the rest of this family—the shorter nucleotide sequences at the C-terminus.
MSTN expression might be identified in various tissues. It can be identified in mammary glands, lymphocytes, spleen, and the cardiomyocytes in heart tissue, and it has an important role in skeletal muscle development [26]. It consists of three exons and two introns. The exons code for a 375 amino-acid (aa) latent protein, which later becomes biologically active through post-translational modification. Through forming disulfide bonds, the polypeptide undergoes intracellular homodimerization [27,28]. Two forms are produced: the N-terminal propeptide region and the C-terminal mature region. These forms initiate intracellular signaling cascades due to their ability to bind and activate the type II activin receptor located on the cell surface (ActRIIB and ActRIIA). Subsequently, the autophosphorylation process of ActRIIB l leads to the recruitment and activation of the low-affinity activin type I receptors ALK-4 or ALK-5. Through phosphorylating the transcription factors Smad2 and Smad3 with activated type I receptor kinase, it is possible for them to interact with Smad4 (co-Smad) and translocate to the nucleus in order to activate the transcription of the target gene [29]. The activated MSTN receptor is able to inhibit protein kinase B, which determines muscle protein synthesis and cell proliferation. The process of increasing the size of muscle fibers is called muscle fiber hypertrophy (or hypertrophy for short) and is strongly regulated through Protein kinase B (Akt). The formation of mature skeletal muscles is the result of myogenic differentiation. The high proliferation muscle precursors formed during embryogenesis differentiate into myoblasts. Myostatin determines the regulation of pre-natal muscle development processes through affecting myoblast proliferation, muscle precursors, and differentiation [30]. MSTN also affects the regulation of the marker that initiates the proliferation of the muscle precursor Pax3 in limb muscles. MSTN is additionally responsible for the increased expression of p21, which stops the proliferation of myoblasts that express myoblast determination protein 1 (MyoD), which is an important regulator of MSTN expression during myogenesis [20].
The phenomenon of muscle hypertrophy was previously identified in many mammalian species. However, muscle hypertrophy is not an accurate term because, in many cases, muscle growth is due to pre-natal hyperplasia [31]. The literature indicates that in species such as cattle, horses, sheep, and goats, changes in MSTN gene expression mainly cause hyperplasia, while hypertrophy is observed in mice [32]. Thus, the term is used casually. Muscles with larger surface areas increase their size significantly, while deeper muscles tend to decrease in size compared to normal muscles. When raising beef cattle, muscle size is very important, as it affects the conformation of the carcass and the proportions of the valuable elements that determine profits from the sale of the animal. Reduced body weights and body fats were observed in obese rats using sActRIIB or a polyclonal antibody to MSTN [33].
MSTN is an important component in myogenesis. However, this action is not its only function that affects production yield. Some sources report that it plays an important role in adipogenesis. MSTN can inhibit either adipogenesis in preadipocytes or can promote it in pluripotent stem cells. The deletion or inhibition of MSTN might improve muscle mass and reduce fat mass [34,35,36]. To determine the effect of MSTN on adipogenesis, white and brown adipocytes should be distinguished. White adipocytes are responsible for storing energy in large lipid droplets, while brown adipocytes contain much more numerous small droplets, which are used in non-shivering thermogenesis [37]. Studies report that MSTN can not only inhibit the adipogenesis of white adipocytes but also of brown adipocytes. This process involves Smad3-mediated β-catenin stabilization and TGF-β/Smad3 signaling [38]. Moreover, it was previously proven that under certain conditions of adipogenesis, mouse embryonic fibroblasts can differentiate into brown fat cells. MSTN-deficient primary mouse embryonic fibroblasts show differentiation into brown adipocytes with increased lipid metabolism [39]. MSTN inhibition can lead to a reduction in subcutaneous body fat in mammals. Transgenic mice whose propeptide cDNA sequence had suppressed MSTN showed reduced subcutaneous, epididymal, and retroperitoneal adipose tissue compared to normal animals [40]. McPherron and Lee [41] concluded that myostatin inhibition may be more effective at limiting adipose tissue gain than reducing it when soluble MSTN receptors from the extracellular domain of type IIB activin receptors were used in mice and induced through a high-fat diet. Reduced body weight and body fat were observed in obese rats using sActRIIB or a polyclonal antibody to MSTN. McPherron and Lee observed that white adipose tissue was converted to brown [41].
There are many reports on the effects of the various allelic variants through which breeding work was carried out to improve both slaughter yield and the quality of the meat itself, as well as reports of inactive MSTN on traits related to growth rate and carcass conformation [20,42,43]. The result of these works is the consolidation of the DM trait in the population, which is a desirable trait because of its beneficial effects on production efficiency and raising the quality of meat—a characteristic that is desired by consumers [44]. These animals have a lower proportion of bone and fat in the carcass, significantly higher proportions of muscle, a lower proportion of connective tissue, and improved meat tenderness. Many researchers focused their work on studying this phenomenon for specific breeds (Table 1).
One of the best-known cattle breeds with the DM trait is the Belgian Blue [49]. Many years of selection for this trait resulted in it being accentuated to an unprecedented level. According to a study by Grobet et al. [24], this breed has an 11-bp deletion (g.821–831 del11) in the open reading frame. There is the loss of three amino acids (275–277) and a shift in the reading frame after aa 274, resulting in a stop codon after aa 287.
The differences between DM and normal cattle can be seen in many aspects of the slaughter performance. The carcass’ lower fat and bone content, the lower collagen content of the muscles, and the significant increase in the size of some muscles relative to normal cattle all account for the value of this trait in beef cattle production conditions [50]. In Belgian Blue cattle, the semitendinosus muscle can be 1.6 times larger than in normal animals [51]. There are also significant changes in subcutaneous and intramuscular fat content. It is worth separating these two types of fat due to the consumption value of the meat. European customers prefer lean beef; thus, these changes are particularly welcome. The situation is, however, different in the markets of many countries where reduced intramuscular fat reduces the steaks’ attractiveness [52,53,54].
Myostatin plays a key role in the processes of adipogenesis and myogenesis. The deletion and inhibition of MSTN contributes mainly to an increase in size of individual skeletal muscles and a reduction in the proportion of fat in the carcass. These are important traits from the breeder’s point of view; thus, they are often used in crossbreeding to improve slaughter performance.
The DM phenotype is characterized by significant muscle hypertrophy relative to normal individuals. There is significant muscle prominence in the hindquarter and anterior quadrant areas, with clearly defined individual muscle parts separated by grooves. Breeding work led to the consolidation of the DM trait, which improves production results. It was only after some time that research began on the impact of this trait on animal health. Arthur et al. pointed out health problems in such animals, stating that lower fertility and lower calf viability were observed [55,56]. Dystocia is another problem found in DM cattle [57]. The well-muscled hindquarters and the effect of hyperplasia on calves prior to birth result in a higher frequency of dystocia. Belgian Blue cattle are the best example of this—almost every parturition ends with a cesarean section. It also turns out that these animals are more likely to become ill, as evidenced by an increased frequency of disease entities involving the respiratory, urinary, digestive, motor, and many other systems. For breeders, the effect on reproduction is also important. Animals with this trait are characterized by higher birth weights, which makes calving more difficult [54]. In addition, DM cattle have increased proportions of glycolytic muscle fibers, which are characterized by a susceptibility to fatigue; thus, these animals show a lower resistance to physical exertion and a faster onset of metabolic acidosis [58].

3. Thyroglobulin

Thyroglobulin (TG) is the main protein of the thyroid gland and makes up to 75% of the gland’s protein [59]. The thyroglobulin gene is considered to be a candidate gene that affects the ability to accumulate intramuscular fat; thus, it is important for breeders and further breeding work. Thyroglobulin production takes place in the thyroid gland’s follicular cells, and is secreted from the endoplasmic reticulum into a site where it undergoes iodination (incorporation of iodine into the tyrosine residues of thyroglobulin). It is stored inside thyroid follicles [60]. As a glycoprotein homodimer, it is a substrate in the production of the thyroid’s hormones and a carrier of triiodothyronine (T3) and tetraiodothyronine (T4) (called thyroxine). The influence of thyroid hormones is important for the regulation of metabolism and its effects on the growth, differentiation, and homeostasis of fat cell composition [61]. Thus, these are important hormones that affect the development of fat cells. Hormone release occurs due to the stimulation of the thyroid cells by the thyrotropic hormone (TSH). Further activity by the released hormones stimulates the hepatic processes of gluconeogenesis and lipogenesis, as well as the occurrence of glycogenolysis [62].
The TG5 gene is one of the longest genes in mammals. In cattle, it is located in the centromere region of the fourteenth chromosome, and consists of 37 exons. It is made up of two allelic variants, i.e., TG5C and TG5T, and three genotypes, i.e., TG5CT, TG5DW, and TG5T [63,64]. This gene affects the accumulation of body fat and is used for animal selection based on a single nucleotide polymorphism (SNP) located in the 5′ untranslated region of this gene [65].
Intramuscular fat content is an important factor in determining the quality of beef. This trait positively correlated with the juiciness and palatability of meat, and improves its flavor, tenderness, and nutritional value. Meat rich in intramuscular fat is characterized by a higher content of fat-soluble vitamins and unsaturated fatty acids [66]. This characteristic is referred to as meat marbling, and influences consumers’ interest during the purchase [67]. Most of the intramuscular fat is located between bundles of muscle fibers in the perimysium connective tissue [68].
Intramuscular fat deposition can be influenced by factors such as sex, weaning age, age and weight at slaughter, nutrition, and environmental factors. However, the trend in the quantitative change in intramuscular fat that is under the influence of the mentioned factors is related to breed [69]. Genetic potential largely determines the final marbling score (MS) [69].
In studies by Rincker et al. [70] and Casas et al., the TG5 SNP had no clear effect on beef marbling [70,71]. This result could be due to the rearing period being too short (<250 days) or other factors. Wood et al., in their meta-analysis based on 11 papers, indicated that there was a positive association between the polymorphic forms of TG5 and the degree of meat marbling [72]. A significant relationship between beef quality and TG5 for Charolaise and Angus cattle was also determined by Van Eenennaam et al. [73]. They indicated there was a significantly higher IMF content for TT genotypes compared to CC. Moreover, in the work of Barendse et al. on a sample containing 1750 cattle, it was indicated that TG5 can be used as an effective tool to improve marbling [9].
Park et al. [69], on the basis of papers written by Albrecht et al. [74] and Irie et al. [75], determined that the average IMF content in the longissimus dorsi (LD) muscle in the Japanese Wagyu breed was 36.5%; for the Korean Hanwoo breed, it was 13.7% [76,77,78,79], while for the Angus breed, it was 7.1% [80,81,82]. For the Hereford crossbreed, the figure was 6.9%. In research by Dubovskova et al., the presence of TT homozygote at 5% was determined in beef characterized by good marbling, and this also had the best results in terms of IMF content [83].

4. The Calpain–Calpastatin System

In the case of CAST and CAPN1, the influence on meat tenderness variability is more than 40% [84]. Thus, they are an extremely important element in the beef production process and have a very strong impact on the quality of the final product. Work carried out on beef tenderness is very important for improving meat quality. Out of the group of genes on which research has been conducted for decades, most of the work focused on calpains (CAPN) and calpastatins (CAST), which are the CAPN inhibitors. From 1993 to 2021, there were at least 175 English-language papers related to the topic [85]. This is a clear signal that work should be conducted to analyze the factors affecting meat quality, in particular tenderness (associated both with CAST [86] and CAPN1 [87,88]), which is the most important determinant of the customers’ willingness to buy. The variability in genes in the calpain–calpastatin system depends on the breed of cattle; therefore, the use of SNPs as genetic markers for animal selection to improve genetic progress is a promising direction [89,90]. Smith et al. indicated that meat tenderness is 46% determined by genetic factors and 54% determined by environmental factors [91].
Calpains are considered a candidate for being responsible for the meat tenderization process, alongside Takahashi’s calcium tenderization theory [92]. Many authors argue that the calpains that are dependent on the presence of calcium ions are responsible for this process. There is the consideration that these calpains may be responsible by virtue of their access to substrates, their ability to hydrolyze proteins, and because they are also found within the cells of muscle tissue. Evidence for their actions are indicated via the reduction in proteolysis under the influence of calcium ion chelators [93], as well as zinc chloride [94]. The main components of the calpain system are µ-calpain (CAPN1), m-calpain (CAPN2), calpain 3 (originally named p94, CAPN3), and their specific inhibitor, i.e., calpastatin (CAST), which blocks their activity [95]. Calpains are found in the cytoplasm of all vertebrate cells [96]. They are named after the calcium ion concentration required for their activation: µ-calpain requires 3–50 µM of calcium and m-calpain 0.4–0.8 mM to reach half of its maximum activity. In live animals, calcium concentration in the muscles is 0.2 µM [97], and only after slaughter does the calcium ion concentration rise to 100 µM [98], thus allowing µ-calpain activation. CAPN1 is considered to be the most important element in the maturation of meat due to its early activation stage, which occurs after slaughter. CAPN2 is activated later, when the calcium ion concentration increases further. CAPN2 is, therefore, important in the later stages of meat maturation. For CAPN3, no significant effect on the post-mortem proteolysis of meat was found [86], though some results may be promising [99,100,101]. The activity of µ-calpain and calpastatin fall sharply in the first few days after slaughter [102], which correlates with an increase in meat tenderness [103]. Boehm et al., Koohmaraie, and Pringle et al., confirmed that calpain plays a major role in this process [93,104,105]. CAPN1 4751 and CAPN1 316 were addressed by research teams in several studies (Table 2), and are largely responsible for meat tenderness in Bos taurus and Bos indicus cattle, as well as in their crosses. In the case of CAPN4751, a significant effect on tenderness was confirmed by Morris et al. [106], while in the case of CAPN316, the cutting force was decreased by about 20% [107] in Bos taurus crosses, which can be used as a meat quality predictor. For the CAPN1 530 marker, no significant effect on meat tenderness was observed in any breed.
Along with µ-Calpain, calpastatins are endogenous calcium-dependent proteinases that are responsible for mediating the proteolysis of myofibrillar proteins during meat aging processes [120]. CAPN1 is responsible for the proteolysis of cytoskeleton proteins and intermediate filaments. Endogenous proteases called calpains and their inhibitor (calpastatin) are thought to be responsible for initiating the degradation of myofibrillar proteins after slaughter [95].
The degradation of cytoskeletal and myofibrillar proteins largely influences changes in muscle cell integrity, which determines the degree to which the meat is tender and shapes organoleptic parameters. This process is shaped through the right aging conditions, such as temperature, time, and type of aging (dry or wet), which affect changes in meat pH over time. The activation of endogenous proteolytic enzymes is necessary to trigger these processes. One of the most important enzymes is µ-calpain, which digests desmin structures and is encoded by the CAPN1 gene. Barendse et al. indicated a strong epistatic effect between the CAST and CAPN1 genes, which occurs in most breeds [121]. The study observed that substituting alanine for glycine in CAPN1: c.947G > C had the greatest effect on meat tenderness in the Angus and Belmont Red cattle breeds.
Changes in consumer needs and eating habits require breeders to make breeding progress and continually improve product quality. A number of studies carried out on calpain also included calpastatin [122], which, as its inhibitor, plays an important role in the maturation of meat [115]. There are several forms of calpastatin, such as CAST, CAST1, CAST2, CAST3, and CAST4 [123]. Calpastatin is also dependent on calcium ions. In a study by Malheiros et al., which was conducted on Bos Indicus, significantly higher expressions of the CAST2 isoform were observed for hard meat and very hard meat compared to medium–hard meat [124]. For the CAST and CAST1 isoforms, no significant differences were observed between the experimental groups. In a study by Muroya et al., variations in expression were observed depending on the type of muscle [125]. Such results indicate that there are variable expressions of CAST isoforms in different muscles (Table 3), which may result in different calpain inhibition and, thus affect meat maturation differently.
In a study by Allais et al., which was conducted on three beef cattle breeds (Charolaise, Limousine, and Blonde d’Aquitaine), differences were found between the CAST SNP results for each breed [126]. For the Blonde d’Aquitaine breed, an increase in required cutting power and a decrease in tenderness were observed for the GA haplotype on the CAST-2 and Cast-3 markers. Casas et al. demonstrated the additive effect of the CAST-2 G allele in the GPE cycle7 group [111], which confirmed the findings of Allais et al. and, at the same time, indicated the positive effect of CAST TT on meat tenderness [126]. Similar conclusions were reached by Johnston and Graser in the case of required cutting strength for the CRC1 population, using Angus, Hereford, and Murray Grey breeds as examples [127]. The G allele was found to have a reduced effect in Charolaise x Angus, Brahman, and Hereford populations [73]. The important role of CAST and the relationship between CAST and CAPN1 in the regulation of beef tenderness was also confirmed by Tait et al. and Lee et al. [112,128].

5. Conclusions

The MSTN SNP, also known as DM, is associated with a mutation in the myostatin gene that affects muscle hypertrophy relative to normal individuals, and can be used to identify the double muscle phenotype in the further selection of individuals. The identification and isolation of the gene makes it possible to distinguish between heterozygous and homozygous individuals, which provides a significant advantage in achieving genetic progress in breeding and, thus, achieving more efficient production. DM is either hyperplasia, which increases the number of muscle fibers pre-natally, or hypertrophy, which manifests itself as an increase in muscle fiber diameter post-natally. The trait is widespread in some European cattle populations, particularly Belgian Blue cattle.
The thyroglobulin gene (TG5) is an important determinant of the degree of meat marbling. Intramuscular fat content is an important element from the consumer’s point of view, and enables the identification of individuals characterized by higher (TT genotype) or lower (CC genotype) proportions of fat in the muscle, which, depending on the market, can be desirable or undesirable. The TG5 SNP will allow breeding directions to be adapted to the needs of consumers in the market.
Calpain and calpastatin are the main determinants of the degree of tenderness in beef. Thus, they are an extremely important factor in the meat maturation process, which is the last stage of production; therefore, any loss in quality at this stage has the greatest consequences. CAST and CAPN1 SNP, through their influence, significantly enhance organoleptic qualities, and breeding work that utilizes them can significantly improve beef quality. It makes sense to identify those individuals that are characterized by the best meat tenderness, as thos strategy encourages customers to buy beef products.
Single nucleotide polymorphisms are a kind of signpost for beef cattle breeders. They are very helpful in the process of breeding progress and allow for more rational decision-making in the selection of individuals. They are also important in scientific and research work. However, it should be remembered that realizing the genetic potential of animals requires the highest possible level of welfare and optimal environmental conditions. Many environmental factors can negatively affect animal weight gain and meat quality. Among the most important factors that we can point to is heat stress, which can cause changes in the color of meat, reduce daily gains [129], and negatively affect reproduction [130]. It is important to provide proper rearing conditions to take advantage of the genetic potential of the animals. Careful analysis of SNPs and the study of their effects on animals is an important research direction, and it is important to carry out further work to learn as much as possible about the operation of such critically important genes in breeding.

Author Contributions

Conceptualization, P.K. and K.P.; methodology, P.K. and K.P.; software, P.K. and K.P.; validation, P.K., J.S. and K.P.; formal analysis, P.K. and K.P.; investigation, P.K. and J.S.; resources, M.G. and G.G.; data curation, J.S.; writing—original draft preparation, P.K. and K.P.; writing—review and editing, P.K. and K.P.; visualization, G.G.; supervision, M.G.; project administration, M.G.; funding acquisition, K.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during the study are included in this published article. The datasets used and/or analyzed in the current study are available from the corresponding author on reasonable request.

Acknowledgments

The paper is a part of the PhD thesis written by Piotr Kostusiak.

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. Kang, N.; Panzone, L.; Kuznesof, S. The role of cooking in consumers’ quality formation: An exploratory study of beef steaks. Meat Sci. 2022, 186, 108730. [Google Scholar] [CrossRef] [PubMed]
  2. Font-i-Furnols, M.; Guerrero, L. Consumer preference, behavior and perception about meat and meat products: An overview. Meat Sci. 2014, 98, 361–371. [Google Scholar] [CrossRef] [PubMed]
  3. Whitton, C.; Bogueva, D.; Marinova, D.; Phillips, C.J. Are we approaching peak meat consumption? Analysis of meat consumption from 2000 to 2019 in 35 countries and its relationship to gross domestic product. Animals 2021, 11, 3466. [Google Scholar] [CrossRef] [PubMed]
  4. Magalhaes, D.R.; Maza, M.T.; Prado, I.N.d.; Fiorentini, G.; Kirinus, J.K.; Campo, M.d.M. An exploratory study of the purchase and consumption of beef: Geographical and cultural differences between Spain and Brazil. Foods 2022, 11, 129. [Google Scholar] [CrossRef] [PubMed]
  5. Banović, M.; Chrysochou, P.; Grunert, K.G.; Rosa, P.J.; Gamito, P. The effect of fat content on visual attention and choice of red meat and differences across gender. Food Qual. Prefer. 2016, 52, 42–51. [Google Scholar] [CrossRef]
  6. Killinger, K.; Calkins, C.R.; Umberger, W.; Feuz, D.M.; Eskridge, K.M. Consumer visual preference and value for beef steaks differing in marbling level and color. J. Anim. Sci. 2004, 82, 3288–3293. [Google Scholar] [CrossRef]
  7. Morales, R.; Aguiar, A.; Subiabre, I.; Realini, C. Beef acceptability and consumer expectations associated with production systems and marbling. Food Qual. Prefer. 2013, 29, 166–173. [Google Scholar] [CrossRef]
  8. Frank, D.; Ball, A.; Hughes, J.; Krishnamurthy, R.; Piyasiri, U.; Stark, J.; Watkins, P.; Warner, R. Sensory and flavor chemistry characteristics of Australian beef: Influence of intramuscular fat, feed, and breed. J. Agric. Food Chem. 2016, 64, 4299–4311. [Google Scholar] [CrossRef]
  9. Barendse, W.; Bunch, R.; Thomas, M.; Armitage, S.; Baud, S.; Donaldson, N. The TG5 thyroglobulin gene test for a marbling quantitative trait loci evaluated in feedlot cattle. Aust. J. Exp. Agric. 2004, 44, 669–674. [Google Scholar] [CrossRef]
  10. McPherron, A.C.; Lee, S.-J. Double muscling in cattle due to mutations in the myostatin gene. Proc. Natl. Acad. Sci. USA 1997, 94, 12457–12461. [Google Scholar] [CrossRef]
  11. Steen, D.; Claeys, E.; Uytterhaegen, L.; De Smet, S.; Demeyer, D. Early post-mortem conditions and the calpain/calpastatin system in relation to tenderness of double-muscled beef. Meat Sci. 1997, 45, 307–319. [Google Scholar] [CrossRef]
  12. Li, X.; Zhang, D.; Ren, C.; Bai, Y.; Ijaz, M.; Hou, C.; Chen, L. Effects of protein posttranslational modifications on meat quality: A review. Compr. Rev. Food Sci. Food Saf. 2021, 20, 289–331. [Google Scholar] [CrossRef]
  13. FAO. FAOSTAT Database. 2019. Available online: https://www.fao.org/faostat/en/#data/QV (accessed on 7 March 2023).
  14. Huang, C.; Hou, C.; Ijaz, M.; Yan, T.; Li, X.; Li, Y.; Zhang, D. Proteomics discovery of protein biomarkers linked to meat quality traits in post-mortem muscles: Current trends and future prospects: A review. Trends Food Sci. Technol. 2020, 105, 416–432. [Google Scholar] [CrossRef]
  15. Kantono, K.; Hamid, N.; Ma, Q.; Chadha, D.; Oey, I. Consumers’ perception and purchase behaviour of meat in China. Meat Sci. 2021, 179, 108548. [Google Scholar] [CrossRef]
  16. Lee, J.; Kim, J.M.; Garrick, D. Increasing the accuracy of genomic prediction in pure-bred Limousin beef cattle by including cross-bred Limousin data and accounting for an F94L variant in MSTN. Anim. Genet. 2019, 50, 621–633. [Google Scholar] [CrossRef]
  17. Prihandini, P.W.; Primasari, A.; Aryogi, A.; Efendy, J.; Luthfi, M.; Pamungkas, D.; Hariyono, D.N.H. Genetic variation in the first intron and exon of the myostatin gene in several Indonesian cattle populations. Vet. World 2021, 14, 1197. [Google Scholar] [CrossRef] [PubMed]
  18. Huang, P.; Pang, D.; Wang, K.; Xu, A.; Yao, C.; Li, M.; You, W.; Wang, Q.; Yu, H. The possible role of complete loss of myostatin in limiting excessive proliferation of muscle cells (C2C12) via activation of microRNAs. Int. J. Mol. Sci. 2019, 20, 643. [Google Scholar] [CrossRef] [PubMed]
  19. Bellinge, R.; Liberles, D.; Iaschi, S.; O’brien, P.; Tay, G. Myostatin and its implications on animal breeding: A review. Anim. Genet. 2005, 36, 1–6. [Google Scholar] [CrossRef]
  20. Aiello, D.; Patel, K.; Lasagna, E. The myostatin gene: An overview of mechanisms of action and its relevance to livestock animals. Anim. Genet. 2018, 49, 505–519. [Google Scholar] [CrossRef] [PubMed]
  21. Ansay, M.; Hanset, R. Anatomical, physiological and biochemical differences between conventional and double-muscled cattle in the Belgian Blue and White breed. Livest. Prod. Sci. 1979, 6, 5–13. [Google Scholar] [CrossRef]
  22. Hanset, R. The major gene of muscular hypertrophy in the Belgian Blue cattle breed. In Breeding for Disease Resistance in Farm Animals; Owen, J., Axford, R., Eds.; CAB International: Wallingford, UK, 1991; pp. 467–478. [Google Scholar]
  23. Hanset, R.; Michaux, C.; Dessy-Doize, C.; Burtonboy, G. Muscle Hypertrophy of Genetic Origin and Its Use to Improve Beef Production; Springer: Dordrecht, The Netherlands, 1982. [Google Scholar]
  24. Grobet, L.; Royo Martin, L.J.; Poncelet, D.; Pirottin, D.; Brouwers, B.; Riquet, J.; Schoeberlein, A.; Dunner, S.; Ménissier, F.; Massabanda, J. A deletion in the bovine myostatin gene causes the double–muscled phenotype in cattle. Nat. Genet. 1997, 17, 71–74. [Google Scholar] [CrossRef] [PubMed]
  25. Sharma, M.; Kambadur, R.; Matthews, K.G.; Somers, W.G.; Devlin, G.P.; Conaglen, J.V.; Fowke, P.J.; Bass, J.J. Myostatin, a transforming growth factor-β superfamily member, is expressed in heart muscle and is upregulated in cardiomyocytes after infarct. J. Cell. Physiol. 1999, 180, 1–9. [Google Scholar] [CrossRef]
  26. Jiao, J.; Yuan, T.; Zhou, Y.; Xie, W.; Zhao, Y.; Zhao, J.; Ouyang, H.; Pang, D. Analysis of myostatin and its related factors in various porcine tissues. J. Anim. Sci. 2011, 89, 3099–3106. [Google Scholar] [CrossRef]
  27. Wolfman, N.M.; McPherron, A.C.; Pappano, W.N.; Davies, M.V.; Song, K.; Tomkinson, K.N.; Wright, J.F.; Zhao, L.; Sebald, S.M.; Greenspan, D.S. Activation of latent myostatin by the BMP-1/tolloid family of metalloproteinases. Proc. Natl. Acad. Sci. USA 2003, 100, 15842–15846. [Google Scholar] [CrossRef]
  28. McPherron, A.C.; Lawler, A.M.; Lee, S.-J. Regulation of skeletal muscle mass in mice by a new TGF-p superfamily member. Nature 1997, 387, 83–90. [Google Scholar] [CrossRef] [PubMed]
  29. Trendelenburg, A.U.; Meyer, A.; Rohner, D.; Boyle, J.; Hatakeyama, S.; Glass, D.J. Myostatin reduces Akt/TORC1/p70S6K signaling, inhibiting myoblast differentiation and myotube size. Am. J. Physiol.-Cell Physiol. 2009, 296, C1258–C1270. [Google Scholar] [CrossRef]
  30. Bryson-Richardson, R.J.; Currie, P.D. The genetics of vertebrate myogenesis. Nat. Rev. Genet. 2008, 9, 632–646. [Google Scholar] [CrossRef]
  31. Amthor, H.; Huang, R.; McKinnell, I.; Christ, B.; Kambadur, R.; Sharma, M.; Patel, K. The regulation and action of myostatin as a negative regulator of muscle development during avian embryogenesis. Dev. Biol. 2002, 251, 241–257. [Google Scholar] [CrossRef] [PubMed]
  32. 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. [Google Scholar] [CrossRef]
  33. Tang, L.; Zhao, T.; Kang, Y.; An, S.; Fan, X.; Sun, L. MSTN is an important myokine for weight-bearing training to attenuate bone loss in ovariectomized rats. J. Physiol. Biochem. 2022, 78, 61–72. [Google Scholar] [CrossRef]
  34. Xin, X.-B.; Yang, S.-P.; Li, X.; Liu, X.-F.; Zhang, L.-L.; Ding, X.-B.; Zhang, S.; Li, G.-P.; Guo, H. Proteomics insights into the effects of MSTN on muscle glucose and lipid metabolism in genetically edited cattle. Gen. Comp. Endocrinol. 2020, 291, 113237. [Google Scholar] [CrossRef] [PubMed]
  35. Kärst, S.; Strucken, E.M.; Schmitt, A.O.; Weyrich, A.; de Villena, F.P.; Yang, H.; Brockmann, G.A. Effect of the myostatin locus on muscle mass and intramuscular fat content in a cross between mouse lines selected for hypermuscularity. BMC Genom. 2013, 14, 16. [Google Scholar] [CrossRef] [PubMed]
  36. Hocquette, J.-F.; Bas, P.; Bauchart, D.; Vermorel, M.; Geay, Y. Fat partitioning and biochemical characteristics of fatty tissues in relation to plasma metabolites and hormones in normal and double-muscled young growing bulls. Comp. Biochem. Physiol. Part A Mol. Integr. Physiol. 1999, 122, 127–138. [Google Scholar] [CrossRef] [PubMed]
  37. Deng, B.; Zhang, F.; Wen, J.; Ye, S.; Wang, L.; Yang, Y.; Gong, P.; Jiang, S. The function of myostatin in the regulation of fat mass in mammals. Nutr. Metab. 2017, 14, 29. [Google Scholar] [CrossRef]
  38. Fournier, B.; Murray, B.; Gutzwiller, S.; Marcaletti, S.; Marcellin, D.; Bergling, S.; Brachat, S.; Persohn, E.; Pierrel, E.; Bombard, F. Blockade of the activin receptor IIb activates functional brown adipogenesis and thermogenesis by inducing mitochondrial oxidative metabolism. Mol. Cell. Biol. 2012, 32, 2871–2879. [Google Scholar] [CrossRef]
  39. Braga, M.; Pervin, S.; Norris, K.; Bhasin, S.; Singh, R. Inhibition of in vitro and in vivo brown fat differentiation program by myostatin. Obesity 2013, 21, 1180–1188. [Google Scholar] [CrossRef]
  40. Zhao, B.; Wall, R.J.; Yang, J. Transgenic expression of myostatin propeptide prevents diet-induced obesity and insulin resistance. Biochem. Biophys. Res. Commun. 2005, 337, 248–255. [Google Scholar] [CrossRef]
  41. McPherron, A.C.; Lee, S.-J. Suppression of body fat accumulation in myostatin-deficient mice. J. Clin. Investig. 2002, 109, 595–601. [Google Scholar] [CrossRef]
  42. Wiener, P.; Woolliams, J.; Frank-Lawale, A.; Ryan, M.; Richardson, R.; Nute, G.; Wood, J.; Homer, D.; Williams, J. The effects of a mutation in the myostatin gene on meat and carcass quality. Meat Sci. 2009, 83, 127–134. [Google Scholar] [CrossRef]
  43. Allais, S.; Levéziel, H.; Payet-Duprat, N.; Hocquette, J.-F.; Lepetit, J.; Rousset, S.; Denoyelle, C.; Bernard-Capel, C.; Journaux, L.; Bonnot, A. The two mutations, Q204X and nt821, of the myostatin gene affect carcass and meat quality in young heterozygous bulls of French beef breeds. J. Anim. Sci. 2010, 88, 446–454. [Google Scholar] [CrossRef]
  44. Purfield, D.; Evans, R.; Berry, D. Reaffirmation of known major genes and the identification of novel candidate genes associated with carcass-related metrics based on whole genome sequence within a large multi-breed cattle population. BMC Genom. 2019, 20, 72. [Google Scholar] [CrossRef] [PubMed]
  45. Bouyer, C.; Forestier, L.; Renand, G.; Oulmouden, A. Deep intronic mutation and pseudo exon activation as a novel muscular hypertrophy modifier in cattle. PLoS ONE 2014, 9, e97399. [Google Scholar] [CrossRef] [PubMed]
  46. Kambadur, R.; Sharma, M.; Smith, T.P.; Bass, J.J. Mutations in myostatin (GDF8) in double-muscled Belgian Blue and Piedmontese cattle. Genome Res. 1997, 7, 910–915. [Google Scholar] [CrossRef]
  47. Cappucio, I.; Marchitelli, C.; Serracchioli, A.; Nardone, A.; Filippini, F.; Ajmone-Marsan, P.; Valentini, A. A GT transversion introduces a stop codon at the mh locus in hypertrophic Marchigiana beef subjects. Anim. Genet 1998, 29, 51. [Google Scholar]
  48. Sellick, G.S.; Pitchford, W.; Morris, C.; Cullen, N.; Crawford, A.; Raadsma, H.; Bottema, C. Effect of myostatin F94L on carcass yield in cattle. Anim. Genet. 2007, 38, 440–446. [Google Scholar] [CrossRef]
  49. Casas, E.; Keele, J.; Shackelford, S.; Koohmaraie, M.; Sonstegard, T.; Smith, T.; Kappes, S.; Stone, R. Association of the muscle hypertrophy locus with carcass traits in beef cattle. J. Anim. Sci. 1998, 76, 468–473. [Google Scholar] [CrossRef]
  50. Dunner, S.; Miranda, M.E.; Amigues, Y.; Cañón, J.; Georges, M.; Hanset, R.; Williams, J.; Ménissier, F. Haplotype diversity of the myostatin gene among beef cattle breeds. Genet. Sel. Evol. 2003, 35, 103–118. [Google Scholar] [CrossRef] [PubMed]
  51. Wegner, J.; Albrecht, E.; Fiedler, I.; Teuscher, F.; Papstein, H.-J.; Ender, K. Growth-and breed-related changes of muscle fiber characteristics in cattle. J. Anim. Sci. 2000, 78, 1485–1496. [Google Scholar] [CrossRef]
  52. Mwashiuya, J.T.; Manyele, S.V.; Mwaluko, G. Assessment of Beef Quality Determinants based on consumer preferences. J. Serv. Sci. Manag. 2018, 11, 657. [Google Scholar] [CrossRef]
  53. Farmer, L.; Farrell, D. Beef-eating quality: A European journey. Animal 2018, 12, 2424–2433. [Google Scholar] [CrossRef]
  54. Egan, A.; Ferguson, D.; Thompson, J. Consumer sensory requirements for beef and their implications for the Australian beef industry. Aust. J. Exp. Agric. 2001, 41, 855–859. [Google Scholar] [CrossRef]
  55. Arthur, P.F.; Makarechian, M.; Price, M.A. Incidence of dystocia and perinatal calf mortality resulting from reciprocal crossing of double-muscled and normal cattle. Can. Vet. J. 1988, 29, 163. [Google Scholar] [PubMed]
  56. King, J.; Menissier, F. Muscle hypertrophy of genetic origin and its use to improve beef. In Current Topics in Veterinary Medicine and Animal Science; Springer: Dordrecht, The Netherlands, 1982. [Google Scholar]
  57. Arthur, P.; Makarechian, M.; Price, M.; Berg, R. Heterosis, maternal and direct effects in double-muscled and normal cattle: I. Reproduction and growth traits. J. Anim. Sci. 1989, 67, 902–910. [Google Scholar] [CrossRef] [PubMed]
  58. Rehfeldt, C.; Ott, G.; Gerrard, D.E.; Varga, L.; Schlote, W.; Williams, J.L.; Renne, U.; Bünger, L. Effects of the Compact mutant myostatin allele Mstn Cmpt-dl1Abc introgressed into a high growth mouse line on skeletal muscle cellularity. J. Muscle Res. Cell Motil. 2005, 26, 103. [Google Scholar] [CrossRef]
  59. Kowalewska-Luczak, I.; Kulig, H.; Szewczyk, K. Polimorfizm w genie tyreoglobuliny u bydła rasy jersey. Acta Sci. Polonorum. Zootech. 2010, 9, 129–134. [Google Scholar]
  60. van der Spek, A.H.; Fliers, E.; Boelen, A. The classic pathways of thyroid hormone metabolism. Mol. Cell. Endocrinol. 2017, 458, 29–38. [Google Scholar] [CrossRef]
  61. Ardicli, S.; Dincel, D.; Samli, H.; Senturk, N.; Karalar, B.; Unlu, S.; Soyudal, B.; Kubad, E.; Balci, F. Association of polymorphisms in lipid and energy metabolism-related genes with fattening performance in Simmental cattle. Anim. Biotechnol. 2022, 2, 1–13. [Google Scholar] [CrossRef]
  62. Dolmatova, I.; Sedykh, T.; Valitov, F.; Gizatullin, R.; Khaziev, D.; Kharlamov, A. Effect of the bovine TG5 gene polymorphism on milk-and meat-producing ability. Vet. World 2020, 13, 2046. [Google Scholar] [CrossRef]
  63. Gan, Q.-F.; Zhang, L.-P.; Li, J.-Y.; Hou, G.-Y.; Li, H.-D.; Gao, X.; Ren, H.-Y.; Chen, J.-B.; Xu, S.-Z. Association analysis of thyroglobulin gene variants with carcass and meat quality traits in beef cattle. J. Appl. Genet. 2008, 49, 251–255. [Google Scholar] [CrossRef]
  64. Ardicli, S.; Samli, H.; Dincel, D.; Ekiz, B.; Yalcintan, H.; Vatansever, B.; Balci, F. Relationship of the bovine IGF1, TG, DGAT1 and MYF5 genes to meat colour, tenderness and cooking loss. J. Hell. Vet. Med. Soc. 2018, 69, 1077–1087. [Google Scholar] [CrossRef]
  65. Carvalho, T.D.D.; Siqueira, F.; Torres Júnior, R.A.D.A.; Medeiros, S.R.D.; Feijó, G.L.D.; Souza Junior, M.D.D.; Blecha, I.M.Z.; Soares, C.O. Association of polymorphisms in the leptin and thyroglobulin genes with meat quality and carcass traits in beef cattle. Rev. Bras. De Zootec. 2012, 41, 2162–2168. [Google Scholar] [CrossRef]
  66. De la Fuente, J.; Diaz, M.; Alvarez, I.; Oliver, M.; i Furnols, M.F.; Sañudo, C.; Campo, M.; Montossi, F.; Nute, G.; Caneque, V. Fatty acid and vitamin E composition of intramuscular fat in cattle reared in different production systems. Meat Sci. 2009, 82, 331–337. [Google Scholar] [CrossRef] [PubMed]
  67. Testa, M.L.; Grigioni, G.; Panea, B.; Pavan, E. Color and marbling as predictors of meat quality perception of Argentinian consumers. Foods 2021, 10, 1465. [Google Scholar] [CrossRef] [PubMed]
  68. Listrat, A.; Lebret, B.; Louveau, I.; Astruc, T.; Bonnet, M.; Lefaucheur, L.; Picard, B.; Bugeon, J. How muscle structure and composition influence meat and flesh quality. Sci. World J. 2016, 2016, 3182746. [Google Scholar] [CrossRef] [PubMed]
  69. Park, S.J.; Beak, S.-H.; Kim, S.Y.; Jeong, I.H.; Piao, M.Y.; Kang, H.J.; Fassah, D.M.; Na, S.W.; Yoo, S.P.; Baik, M. Genetic, management, and nutritional factors affecting intramuscular fat deposition in beef cattle—A review. Asian-Australas. J. Anim. Sci. 2018, 31, 1043. [Google Scholar] [CrossRef] [PubMed]
  70. Rincker, C.B.; Pyatt, N.A.; Berger, L.L.; Faulkner, D.B. Relationship among GeneSTAR marbling marker, intramuscular fat deposition, and expected progeny differences in early weaned Simmental steers. J. Anim. Sci. 2006, 84, 686–693. [Google Scholar] [CrossRef]
  71. Casas, E.; White, S.N.; Riley, D.G.; Smith, T.P.L.; Brenneman, R.A.; Olson, T.A.; Johnson, D.D.; Coleman, S.W.; Bennett, G.L.; Chase, C.C., Jr. Assessment of single nucleotide polymorphisms in genes residing on chromosomes 14 and 29 for association with carcass composition traits in Bos indicus cattle1,2. J. Anim. Sci. 2005, 83, 13–19. [Google Scholar] [CrossRef] [PubMed]
  72. Wood, I.A.; Moser, G.; Burrell, D.L.; Mengersen, K.L.; Hetzel, D.J.S. A meta-analytic assessment of a Thyroglobulin marker for marbling in beef cattle. Genet. Sel. Evol. 2006, 38, 479–494. [Google Scholar] [CrossRef]
  73. Van Eenennaam, A.L.; Li, J.; Thallman, R.M.; Quaas, R.L.; Dikeman, M.E.; Gill, C.A.; Franke, D.E.; Thomas, M.G. Validation of commercial DNA tests for quantitative beef quality traits1,2. J. Anim. Sci. 2007, 85, 891–900. [Google Scholar] [CrossRef]
  74. Albrecht, E.; Gotoh, T.; Ebara, F.; Xu, J.; Viergutz, T.; Nürnberg, G.; Maak, S.; Wegner, J. Cellular conditions for intramuscular fat deposition in Japanese Black and Holstein steers. Meat Sci. 2011, 89, 13–20. [Google Scholar] [CrossRef]
  75. Irie, M.; Kouda, M.; Matono, H. Effect of ursodeoxycholic acid supplementation on growth, carcass characteristics, and meat quality of Wagyu heifers (Japanese Black cattle). J. Anim. Sci. 2011, 89, 4221–4226. [Google Scholar] [CrossRef]
  76. Jeong, J.; Kwon, E.; Im, S.; Seo, K.; Baik, M. Expression of fat deposition and fat removal genes is associated with intramuscular fat content in longissimus dorsi muscle of Korean cattle steers. J. Anim. Sci. 2012, 90, 2044–2053. [Google Scholar] [CrossRef] [PubMed]
  77. Cho, S.; Kang, G.; Seong, P.-N.; Park, B.; Kang, S.M. Effect of slaughter age on the antioxidant enzyme activity, color, and oxidative stability of Korean Hanwoo (Bos taurus coreanae) cow beef. Meat Sci. 2015, 108, 44–49. [Google Scholar] [CrossRef] [PubMed]
  78. Choi, C.; Jung, K.; Chung, K.; Yang, B.; Chin, K.; Suh, S.; Oh, D.; Jeon, M.; Baek, K.; Lee, S. Administration of zilpaterol hydrochloride alters feedlot performance, carcass characteristics, muscle, and fat profiling in finishing Hanwoo steers. Livest. Sci. 2013, 157, 435–441. [Google Scholar] [CrossRef]
  79. Jung, S.; Nam, K.C.; Lee, K.H.; Kim, J.J.; Jo, C. Meat quality traits of Longissimus dorsi muscle from carcasses of Hanwoo steers at different yield grades. Food Sci. Anim. Resour. 2013, 33, 305–316. [Google Scholar] [CrossRef]
  80. Greenwood, P.L.; Siddell, J.; Walmsley, B.; Geesink, G.; Pethick, D.; McPhee, M. Postweaning substitution of grazed forage with a high-energy concentrate has variable long-term effects on subcutaneous fat and marbling in Bos taurus genotypes. J. Anim. Sci. 2015, 93, 4132–4143. [Google Scholar] [CrossRef]
  81. Krone, K.; Ward, A.; Madder, K.; Hendrick, S.; McKinnon, J.; Buchanan, F. Interaction of vitamin A supplementation level with ADH1C genotype on intramuscular fat in beef steers. Animal 2016, 10, 403–409. [Google Scholar] [CrossRef]
  82. Dinh, T.; Blanton Jr, J.; Riley, D.; Chase Jr, C.; Coleman, S.; Phillips, W.; Brooks, J.; Miller, M.; Thompson, L. Intramuscular fat and fatty acid composition of longissimus muscle from divergent pure breeds of cattle. J. Anim. Sci. 2010, 88, 756–766. [Google Scholar] [CrossRef]
  83. Dubovskova, M.; Selionova, M.; Chizhova, L.; Surzhikova, E.; Gerasimov, N.; Mikhailenko, A.; Dolgashova, M. Use of genetic markers of meat productivity in breeding of Hereford breed bulls. Proc. IOP Conf. Series Earth Environ. Sci. 2019, 341, 012052. [Google Scholar] [CrossRef]
  84. Bernard, C.; Cassar-Malek, I.; Le Cunff, M.; Dubroeucq, H.; Renand, G.; Hocquette, J.-F. New indicators of beef sensory quality revealed by expression of specific genes. J. Agric. Food Chem. 2007, 55, 5229–5237. [Google Scholar] [CrossRef]
  85. Gonzales-Malca, J.A.; Tirado-Kulieva, V.A.; Abanto-López, M.S.; Aldana-Juárez, W.L.; Palacios-Zapata, C.M. Bibliometric Analysis of Research on the Main Genes Involved in Meat Tenderness. Animals 2022, 12, 2976. [Google Scholar] [CrossRef] [PubMed]
  86. Uzabaci, E.; Dincel, D. Associations Between c. 2832A < G Polymorphism of CAST Gene and Meat Tenderness in Cattle: A Meta-Analysis CAST Geninin c. 2832A< G Polimorfizmi ile Sığırlarda Et Gevrekliği Arasındaki İlişki: Bir Meta-Analizi. Kafkas Univ. Vet. Fak. Derg. 2022, 28, 613–620. [Google Scholar]
  87. Koohmaraie, M.; Geesink, G. Contribution of postmortem muscle biochemistry to the delivery of consistent meat quality with particular focus on the calpain system. Meat Sci. 2006, 74, 34–43. [Google Scholar] [CrossRef]
  88. Gagaoua, M.; Terlouw, E.C.; Mullen, A.M.; Franco, D.; Warner, R.D.; Lorenzo, J.M.; Purslow, P.P.; Gerrard, D.; Hopkins, D.L.; Troy, D. Molecular signatures of beef tenderness: Underlying mechanisms based on integromics of protein biomarkers from multi-platform proteomics studies. Meat Sci. 2021, 172, 108311. [Google Scholar] [CrossRef] [PubMed]
  89. Abd El-Hack, M.E.; Abdelnour, S.A.; Swelum, A.A.; Arif, M. The application of gene marker-assisted selection and proteomics for the best meat quality criteria and body measurements in Qinchuan cattle breed. Mol. Biol. Rep. 2018, 45, 1445–1456. [Google Scholar] [CrossRef]
  90. Brito Lopes, F.; Magnabosco, C.U.; Passafaro, T.L.; Brunes, L.C.; Costa, M.F.; Eifert, E.C.; Narciso, M.G.; Rosa, G.J.; Lobo, R.B.; Baldi, F. Improving genomic prediction accuracy for meat tenderness in Nellore cattle using artificial neural networks. J. Anim. Breed. Genet. 2020, 137, 438–448. [Google Scholar] [CrossRef]
  91. Smith, T.P.; Thallman, R.M.; Casas, E.; Shackelford, S.D.; Wheeler, T.L.; Koohmaraie, M. Theory and application of genome-based approaches to improve the quality and value of beef. Outlook Agric. 2003, 32, 253–265. [Google Scholar] [CrossRef]
  92. Takahashi, K. Mechanism of meat tenderization during post-mortem ageing: Calcium theory. In Proceedings of the International Congress of Meat Science and Technology, Yokohama, Japan, 1–6 August 1999; pp. 230–235. [Google Scholar]
  93. Koohmaraie, M. The role of Ca(2+)-dependent proteases (calpains) in post mortem proteolysis and meat tenderness. Biochimie 1992, 74, 239–245. [Google Scholar] [CrossRef]
  94. Watanabe, A.; Daly, C.; Devine, C. The effects of the ultimate pH of meat on tenderness changes during ageing. Meat Sci. 1996, 42, 67–78. [Google Scholar] [CrossRef]
  95. Bhat, Z.; Morton, J.D.; Mason, S.L.; Bekhit, A.E.-D.A. Role of calpain system in meat tenderness: A review. Food Sci. Hum. Wellness 2018, 7, 196–204. [Google Scholar] [CrossRef]
  96. Dransfield, E. Meat tenderness–the µ-calpain hypothesis. In Proceedings of the 45th International Congress of Meat Science and Technology, Yokohama, Japan, 1–6 August 1999. [Google Scholar]
  97. Kurebayashi, N.; Harkins, A.; Baylor, S. Use of fura red as an intracellular calcium indicator in frog skeletal muscle fibers. Biophys. J. 1993, 64, 1934–1960. [Google Scholar] [CrossRef]
  98. Jeacocke, R.E. The concentrations of free magnesium and free calcium ions both increase in skeletal muscle fibres entering rigor mortis. Meat Sci. 1993, 35, 27–45. [Google Scholar] [CrossRef]
  99. Ilian, M.A.; Bekhit, A.E.-D.; Bickerstaffe, R. The relationship between meat tenderization, myofibril fragmentation and autolysis of calpain 3 during post-mortem aging. Meat Sci. 2004, 66, 387–397. [Google Scholar] [CrossRef] [PubMed]
  100. Ilian, M.A.; Morton, J.D.; Bekhit, A.E.-D.; Roberts, N.; Palmer, B.; Sorimachi, H.; Bickerstaffe, R. Effect of preslaughter feed withdrawal period on longissimus tenderness and the expression of calpains in the ovine. J. Agric. Food Chem. 2001, 49, 1990–1998. [Google Scholar] [CrossRef]
  101. Yang, X.; Chen, J.; Jia, C.; Zhao, R. Gene expression of calpain 3 and PGC-1α is correlated with meat tenderness in the longissimus dorsi muscle of Sutai pigs. Livest. Sci. 2012, 147, 119–125. [Google Scholar] [CrossRef]
  102. Koohmaraie, M. Effect of pH, temperature, and inhibitors on autolysis and catalytic activity of bovine skeletal muscle μ-calpain. J. Anim. Sci. 1992, 70, 3071–3080. [Google Scholar] [CrossRef]
  103. Thomson, B.; Dobbie, P.; Singh, K.; Speck, P. Post-mortem kinetics of meat tenderness and the components of the calpain system in bull skeletal muscle. Meat Sci. 1996, 44, 151–157. [Google Scholar] [CrossRef]
  104. Boehm, M.L.; Kendall, T.L.; Thompson, V.F.; Goll, D.E. Changes in the calpains and calpastatin during postmortem storage of bovine muscle. J. Anim. Sci. 1998, 76, 2415–2434. [Google Scholar] [CrossRef]
  105. Pringle, T.; Harrelson, J.; West, R.; Williams, S.; Johnson, D. Calcium-activated tenderization of strip loin, top sirloin, and top round steaks in diverse genotypes of cattle. J. Anim. Sci. 1999, 77, 3230–3237. [Google Scholar] [CrossRef]
  106. Morris, C.; Cullen, N.; Hickey, S.; Dobbie, P.; Veenvliet, B.; Manley, T.; Pitchford, W.; Kruk, Z.; Bottema, C.; Wilson, T. Genotypic effects of calpain 1 and calpastatin on the tenderness of cooked M. longissimus dorsi steaks from Jersey× Limousin, Angus and Hereford-cross cattle. Anim. Genet. 2006, 37, 411–414. [Google Scholar] [CrossRef] [PubMed]
  107. White, S.; Casas, E.; Wheeler, T.; Shackelford, S.; Koohmaraie, M.; Riley, D.; Chase Jr, C.; Johnson, D.; Keele, J.; Smith, T. A new single nucleotide polymorphism in CAPN1 extends the current tenderness marker test to include cattle of Bos indicus, Bos taurus, and crossbred descent. J. Anim. Sci. 2005, 83, 2001–2008. [Google Scholar] [CrossRef] [PubMed]
  108. Basson, A.; Strydom, P.E.; van Marle-Köster, E.; Webb, E.C.; Frylinck, L. Sustained Effects of Muscle Calpain System Genotypes on Tenderness Phenotypes of South African Beef Bulls during Ageing up to 20 Days. Animals 2022, 12, 686. [Google Scholar] [CrossRef] [PubMed]
  109. Avilés, C.; Juárez, M.; Peña, F.; Domenech, V.; Clemente, I.; Molina, A. Association of single nucleotide polymorphisms in CAPN1 and CAST genes with beef tenderness from Spanish commercial feedlots. Czech. J. Anim. Sci. 2013, 58, 479–487. [Google Scholar] [CrossRef]
  110. Page, B.; Casas, E.; Heaton, M.; Cullen, N.; Hyndman, D.; Morris, C.; Crawford, A.; Wheeler, T.; Koohmaraie, M.; Keele, J. Evaluation of single-nucleotide polymorphisms in CAPN1 for association with meat tenderness in cattle. J. Anim. Sci. 2002, 80, 3077–3085. [Google Scholar] [CrossRef]
  111. Casas, E.; White, S.; Wheeler, T.; Shackelford, S.; Koohmaraie, M.; Riley, D.; Chase Jr, C.; Johnson, D.; Smith, T. Effects of calpastatin and μ-calpain markers in beef cattle on tenderness traits. J. Anim. Sci. 2006, 84, 520–525. [Google Scholar] [CrossRef]
  112. Lee, S.-H.; Kim, S.-C.; Chai, H.-H.; Cho, S.-H.; Kim, H.-C.; Lim, D.; Choi, B.-H.; Dang, C.-G.; Sharma, A.; Gondro, C. Mutations in calpastatin and μ-calpain are associated with meat tenderness, flavor and juiciness in Hanwoo (Korean cattle): Molecular modeling of the effects of substitutions in the calpastatin/μ-calpain complex. Meat Sci. 2014, 96, 1501–1508. [Google Scholar] [CrossRef] [PubMed]
  113. Rubio Lozano, M.S.; Alfaro-Zavala, S.; Sifuentes-Rincón, A.M.; Parra-Bracamonte, G.M.; Braña Varela, D.; Medina, R.D.M.; Pérez Linares, C.; Ríos Rincón, F.; Sánchez Escalante, A.; Torrescano Urrutia, G. Meat tenderness genetic and genomic variation sources in commercial beef cattle. J. Food Qual. 2016, 39, 150–156. [Google Scholar] [CrossRef]
  114. Smith, T.; Thomas, M.; Bidner, T.; Paschal, J.; Franke, D. Single nucleotide polymorphisms in Brahman steers and their association with carcass and tenderness traits. Genet. Mol. Res 2009, 8, 39–46. [Google Scholar] [CrossRef]
  115. Pinto, L.; Ferraz, J.B.S.; Meirelles, F.V.; Eler, J.P.; Rezende, F.M.d.; Carvalho, M.; Almeida, H.; Silva, R. Association of SNPs on CAPN 1 and CAST genes with tenderness in Nellore cattle. Genet. Mol. Res. 2010, 9, 1431–1442. [Google Scholar] [CrossRef]
  116. Pinto, L.F.B.; Ferraz, J.B.S.; Pedrosa, V.B.; Eler, J.P.; Meirelles, F.V.; Bonin, M.d.N.; Rezende, F.M.D.; Carvalho, M.E.; Cucco, D.D.C.; Silva, R.C.G.D. Single nucleotide polymorphisms in CAPN and leptin genes associated with meat color and tenderness in Nellore cattle. Genet. Mol. Res. 2011, 10, 2057–2064. [Google Scholar]
  117. Rosa, A.F.; Moncau, C.T.; Poleti, M.D.; Fonseca, L.D.; Balieiro, J.C.; Silva, S.L.; Eler, J.P. Proteome changes of beef in Nellore cattle with different genotypes for tenderness. Meat Sci. 2018, 138, 1–9. [Google Scholar] [CrossRef] [PubMed]
  118. Kök, S.; Atalay, S. The Use of various SNPs in CAST and CAPN1 genes to determine the meat tenderness in Turkish grey cattle. Kafkas Univ. Vet. Fak. Derg. 2018, 24, 1–8. [Google Scholar]
  119. Calvo, J.; Iguácel, L.; Kirinus, J.; Serrano, M.; Ripoll, G.; Casasús, I.; Joy, M.; Pérez-Velasco, L.; Sarto, P.; Albertí, P. A new single nucleotide polymorphism in the calpastatin (CAST) gene associated with beef tenderness. Meat Sci. 2014, 96, 775–782. [Google Scholar] [CrossRef]
  120. Schenkel, F.S.; Miller, S.P.; Jiang, Z.; Mandell, I.B.; Ye, X.; Li, H.; Wilton, J.W. Association of a single nucleotide polymorphism in the calpastatin gene with carcass and meat quality traits of beef cattle1. J. Anim. Sci. 2006, 84, 291–299. [Google Scholar] [CrossRef]
  121. Barendse, W.; Harrison, B.E.; Hawken, R.J.; Ferguson, D.M.; Thompson, J.M.; Thomas, M.B.; Bunch, R.J. Epistasis Between Calpain 1 and Its Inhibitor Calpastatin Within Breeds of Cattle. Genetics 2007, 176, 2601–2610. [Google Scholar] [CrossRef] [PubMed]
  122. Liu, J.; Fu, R.; Liu, R.; Zhao, G.; Zheng, M.; Cui, H.; Li, Q.; Song, J.; Wang, J.; Wen, J. Protein profiles for muscle development and intramuscular fat accumulation at different post-hatching ages in chickens. PLoS ONE 2016, 11, e0159722. [Google Scholar] [CrossRef] [PubMed]
  123. Cônsolo, N.R.B.; Ferrari, V.B.; Mesquita, L.G.; Goulart, R.S.; e Silva, L.F.P. Zilpaterol hydrochloride improves beef yield, changes palatability traits, and increases calpain-calpastatin gene expression in Nellore heifers. Meat Sci. 2016, 121, 375–381. [Google Scholar] [CrossRef]
  124. Malheiros, J.M.; Enríquez-Valencia, C.E.; da Silva Duran, B.O.; de Paula, T.G.; Curi, R.A.; de Vasconcelos Silva, J.A.I.; Dal-Pai-Silva, M.; de Oliveira, H.N.; Chardulo, L.A.L. Association of CAST2, HSP90AA1, DNAJA1 and HSPB1 genes with meat tenderness in Nellore cattle. Meat Sci. 2018, 138, 49–52. [Google Scholar] [CrossRef] [PubMed]
  125. Muroya, S.; Neath, K.E.; Nakajima, I.; Oe, M.; Shibata, M.; Ojima, K.; Chikuni, K. Differences in mRNA expression of calpains, calpastatin isoforms and calpain/calpastatin ratios among bovine skeletal muscles. Anim. Sci. J. 2012, 83, 252–259. [Google Scholar] [CrossRef]
  126. Allais, S.; Journaux, L.; Levéziel, H.; Payet-Duprat, N.; Raynaud, P.; Hocquette, J.-F.; Lepetit, J.; Rousset, S.; Denoyelle, C.; Bernard-Capel, C. Effects of polymorphisms in the calpastatin and µ-calpain genes on meat tenderness in 3 French beef breeds. J. Anim. Sci. 2011, 89, 1–11. [Google Scholar] [CrossRef]
  127. Johnston, D.; Graser, H.-U. Estimated gene frequencies of GeneSTAR markers and their size of effects on meat tenderness, marbling, and feed efficiency in temperate and tropical beef cattle breeds across a range of production systems. J. Anim. Sci. 2010, 88, 1917–1935. [Google Scholar] [CrossRef] [PubMed]
  128. Tait Jr, R.; Shackelford, S.; Wheeler, T.; King, D.; Keele, J.; Casas, E.; Smith, T.; Bennett, G. CAPN1, CAST, and DGAT1 genetic effects on preweaning performance, carcass quality traits, and residual variance of tenderness in a beef cattle population selected for haplotype and allele equalization. J. Anim. Sci. 2014, 92, 5382–5393. [Google Scholar] [CrossRef] [PubMed]
  129. Wang, J.; Li, J.; Wang, F.; Xiao, J.; Wang, Y.; Yang, H.; Li, S.; Cao, Z. Heat stress on calves and heifers: A review. J. Anim. Sci. Biotechnol. 2020, 11, 79. [Google Scholar] [CrossRef] [PubMed]
  130. Dobson, H.; Smith, R. What is stress, and how does it affect reproduction? Anim. Reprod. Sci. 2000, 60, 743–752. [Google Scholar] [CrossRef]
Table 1. MSTN gene polymorphism in cattle breeds.
Table 1. MSTN gene polymorphism in cattle breeds.
SNP
Reference Breed Position Mutation
[24]Belgian Blue c.821 Del11
[10]Blonde d’Aquitainec.821Del11
[45] g.3811 T > G
[46]Charolaisec.610 C > T
[46]Limousinec.821Del11
[47]c.610 C > T
[48]g.433 C > A
[47] Marchigiana g.874 G > T
[46]Piedmontese c.938 G > A
Table 2. CAPN gene polymorphism in cattle breeds.
Table 2. CAPN gene polymorphism in cattle breeds.
ReferenceBreedMuscleCAPN SNP
[108] Angus, Charolaise, Brahman, and NguniLongissimus thoracis et lumborumCAPN1 184+, CAPN1 187+, CAPN1 4751+, and CAPN2 780+
[109] Charolaise, Limousine, and RetintaLongissimus dorsiCAPN1+
[106] Jersey–Limousine cross, Angus, and Hereford crossLongissimus dorsiCAPN1: c.947C > G+
[110] Piedmontese–Angus cross and Jersey-Limousine crossLongissimus thoracis38 SNPs+
[111] Angus, Red Angus, Beefmaster, Brangus, Hereford, Bonsmara, Romosinuano, Brahman, Limousine, Charolaise, Gelbvieh, and SimmentalNo dataCAPN1+
[107] Brangus, Beefmaster, Bonsmara, Brahman, Romosinuano, Hereford, and AngusLongissimusCAPN1 316+, CAPN1 4753+, and CAPN1 530+
[112] HanwooLongissimus lumborumCAPN1:c.1589G > A+, CAPN1:c.658C > T+, CAPN1:c.948G > C+, and CAPN1:c.580A > G+
[113] B. taurus, B. indicus, and crossesLongissimus dorsiCAPN1 316+ and CAPN1 4751+
[114] BrahmanLongissimus dorsiCAPN316+ and CAPN4751+
[115] NelloreLongissimus dorsiCAPN1 316+, CAPN1 4751+, CAPN1 530+, and CAPN1 4753+
[116] NelloreLongissimus dorsiCAPN1 4751
[117] NelloreLongissimus dorsiCAPN1 4751+
[118] Turkish GreyLongissimus dorsiCAPN1 316+ and CAPN1 4751+
[119] Parda de Montaña and PirenaicaLongissimus thoracisCAPN1 316, CAPN1 530, and CAPN1 4751
No association with meat tenderness, (−); association with meat tenderness, (+).
Table 3. CAST gene polymorphism in cattle breeds.
Table 3. CAST gene polymorphism in cattle breeds.
ReferenceBreedMuscleCAST SNP
[108]Angus, Charolaise, Brahman, and NguniLongissimus thoracis et lumborumCAST 736+ and CAST 763+
[109]Charolaise, Limousine, and RetintaLongissimus dorsiCAST+
[106]Jersey–Limousine cross, Angus–Hereford, and other crossesLongissimus dorsiCAST: c.2959A > G+
[111]Angus, Red Angus, Beefmaster, Brangus, Hereford, Bonsmara, Romosinuano, Brahman Limousine, Charolaise, Gelbvieh, and SimmentalNo dataCAST+
[112]HanwooLongissimus lumborumCAST:c.182A > G+, CAST:c.1985G > C+, and CAST:c.1526T > C+
[113]B. taurus, B. indicus, and crossesLongissimus dorsiCAST-T1
[114]BrahmanLongissimus dorsiCAST+
[115]NelloreLongissimus dorsiUOGCAST+ and WSUCAST+
[117]NelloreLongissimus dorsiUOGCAST+
[118]Turkish GreyLongissimus dorsiUOGCAST+
[119]Parda de Montaña and PirenaicaLongissimus thoracisCAST1+, CAST2+, CAST3, CAST4+, and CAST5
No association with meat tenderness, (−); association with meat tenderness, (+).
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

Kostusiak, P.; Slósarz, J.; Gołębiewski, M.; Grodkowski, G.; Puppel, K. Polymorphism of Genes and Their Impact on Beef Quality. Curr. Issues Mol. Biol. 2023, 45, 4749-4762. https://doi.org/10.3390/cimb45060302

AMA Style

Kostusiak P, Slósarz J, Gołębiewski M, Grodkowski G, Puppel K. Polymorphism of Genes and Their Impact on Beef Quality. Current Issues in Molecular Biology. 2023; 45(6):4749-4762. https://doi.org/10.3390/cimb45060302

Chicago/Turabian Style

Kostusiak, Piotr, Jan Slósarz, Marcin Gołębiewski, Grzegorz Grodkowski, and Kamila Puppel. 2023. "Polymorphism of Genes and Their Impact on Beef Quality" Current Issues in Molecular Biology 45, no. 6: 4749-4762. https://doi.org/10.3390/cimb45060302

Article Metrics

Back to TopTop