Next Article in Journal
Oocyte Quality, In Vitro Fertilization and Embryo Development of Alpaca Oocytes Collected by Ultrasound-Guided Follicular Aspiration or from Slaughterhouse Ovaries
Previous Article in Journal
Vibrio metschnikovii, a Potential Pathogen in Freshwater-Cultured Hybrid Sturgeon
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification of Candidate Genes Regulating Carcass Depth and Hind Leg Circumference in Simmental Beef Cattle Using Illumina Bovine Beadchip and Next-Generation Sequencing Analyses

1
Key Laboratory of Animal Genetics Breeding and Reproduction, Ministry of Agriculture and Rural Affairs, Institute of Animal Sciences, Chinese Academy of Agricultural Sciences, Beijing 100193, China
2
Department of Animal Science, Faculty of Agriculture, Shahid Bahonar University of Kerman, Kerman 77951, Iran
3
Center for Quantitative Genetics and Genomics, Aarhus University, 8210 Aarhus, Denmark
*
Author to whom correspondence should be addressed.
Animals 2022, 12(9), 1103; https://doi.org/10.3390/ani12091103
Submission received: 19 February 2022 / Revised: 14 April 2022 / Accepted: 21 April 2022 / Published: 25 April 2022
(This article belongs to the Section Animal Genetics and Genomics)

Abstract

:

Simple Summary

The performance of genome-wide association studies with high-density single-nucleotide polymorphism panels and next-generation sequencing is a robust approach to the identification of genetic variants that can elucidate variation in complex diseases and economically important traits in farm animals. We identified many markers associated with crucial traits, such as muscle development in Simmental beef cattle. This information can be used to pinpoint candidate genes affecting traits of interest. Our results contribute to the clarification of molecular mechanisms underlying processes associated with hind-leg meat yield and carcass depth in beef cattle.

Abstract

Genome-wide association studies are a robust means of identifying candidate genes that regulate economically important traits in farm animals. The aim of this study is to identify single-nucleotide polymorphisms (SNPs) and candidate genes potentially related to carcass depth and hind leg circumference in Simmental beef cattle. We performed Illumina Bovine HD Beadchip (~670 k SNPs) and next-generation sequencing (~12 million imputed SNPs) analyses of data from 1252 beef cattle, to which we applied a linear mixed model. Using a statistical threshold (p = 0.05/number of SNPs identified) and adopting a false discovery rate (FDR), we identified many putative SNPs on different bovine chromosomes. We identified 12 candidate genes potentially annotated with the markers identified, including CDKAL1 and E2F3, related to myogenesis and skeletal muscle development. The identification of such genes in Simmental beef cattle will help breeders to understand and improve related traits, such as meat yield.

1. Introduction

Farm animals, particularly cattle, are of major economic importance in many parts of the world [1]. Beef cattle breeds are being developed genetically to maximize the efficiency of meat production [2] using breeding programs based on traits related to meat yield [3]. Recently, beef breeders have been inclined toward the augmentation of lean muscle mass growth to meet rapidly changing consumer demand for lean meat [4]. Thus, in the past several years, researchers have made more efforts to increase our understanding of genetic influences on economically important traits covering different aspects of meat production, especially muscle growth and development.
Simmental beef cattle, distributed in nearly all parts of the world, are bred primarily due to their high levels of growth and lean meat production under an appropriate diet [5,6,7]. Many studies have been conducted to examine the economically crucial traits of this breed, such as meat production [8], meat quality [9], carcass characteristics [10], and growth [11]. An understanding of the effects of genetic variation on meat quality, meat yield, and carcass quality will aid breeders’ implementation of methods to effectively address beef market demands.
The search for candidate genes via genetic association studies, conducted for almost all species [12], may be practical for the extension of our understanding of genes affecting economically important traits. Candidate genes may affect particular biological actions related to different aspects of a trait of interest, and their identification is an important step in efforts to identify quantitative trait loci (QTLs) that regulate the genetic variation of traits of interest [13].
The performance of genome-wide association studies (GWASs) using high-density single nucleotide polymorphism (SNP) panels is a robust approach to the identification of different parts of the genome and genetic variants that can elucidate variation in complex diseases and prominent or economically important traits in farm animals [14,15,16] and humans [17,18,19]. GWAS results can be sensitive to different factors, such as the sample size and the number of genes affecting traits of interest [20]. Next-generation sequencing (NGS) can be employed to access numerous genetic variants in close proximity to genes affecting traits of interest. Sharma et al. [21] used NGS to identify 18 mutations associated with Mendelian diseases in Hanwoo cattle; in addition, in the same cattle, 33 genes were found to be associated with domestication [22]. In Simmental beef cattle, NGS has been used to identify important genes that control hind leg width, such as PLXNA4, ASB15, SLC13A1, NDUFA5, IQUB, LMOD2, and WASL [23], and a promising candidate gene that controls net meat weight (MTPN) [24]. However, due to the high SNP density in NGS data, the calculation of marker effects is challenging [23,24,25].
Hindquarters from beef carcasses include valuable cuts, and their properties are thus important components of meat-related traits [23]. The hindquarter consists of connective tissues; the femur, tibia, and fibula; muscles including the semitendinosus, vastus lateralis, tensor fascia lata, gluteus medius, gastrocnemius, semimembranosus, and biceps femoris. Most of these muscles are skeletal and supply large amounts of meat. Muscle quantity has been reported to be closely related to the hind leg dimension [26]. In addition, the hind-leg muscle thickness correlated positively with the growth rate of European beef cattle [26].
In the past several years, genetic selection for carcass conformation traits has been practiced in most beef cattle populations. Carcass depth is an important carcass conformation trait that covers the principal muscle and bone tissues and thus is profitable in the beef industry. Han et al. [27] illustrated that it was associated with the IGF2 gene, a key gene for muscle development and growth in farm animals [28]. However, the literature on carcass depth is limited, and more research is needed.
In this GWAS, we detected SNPs for the identification of candidate genes (via QTLs and markers) associated with hind leg circumference and carcass depth in Simmental beef cattle. Our results can contribute to the clarification of the molecular mechanisms underlying hind-leg meat yield and carcass depth.

2. Material and Methods

2.1. Ethics Statement

All procedures were conducted in accordance with the guidelines of China’s Council on Animal Care, and the study protocol was approved by the Institute of Animal Science of the Chinese Academy of Agricultural Science, Beijing, China (approval number: RNL09/07). Analyses associated with animals adhered to the rules of China’s Council on Animal Welfare.

2.2. Animal Resources and Phenotype Data

The Simmental beef cattle population examined in this study comprised 1346 cattle born between 2009 and 2015 in Ulgai, Xilingol League, Inner Mongolia, China. Following weaning, the cattle were transferred to the Jinweifuren Farm feedlot in Beijing. All calves were raised under standard conditions and fed a total mixed ration based on the nutritional requirements of beef cattle (NRC, 2018). At 16–18 months of age, the animals were slaughtered. Immediately after bloodletting, the target carcasses were refrigerated at 4°C until trait measurement (about 9 h after slaughter). The carcasses were hanged, and the carcass depth and hind leg circumference were measured in centimeters using a tape (Figure 1). Trait means, standard deviations, and ranges are shown in Table 1.

2.3. Genotype Examination and Quality Control

Illumina Bovine HD SNP Beadchip (770 k) and genotype examinations were performed using Illumina Genome Studio (Illumina Inc., San Diego, CA, USA). For quality control, individuals with high Pi-Hat values (representing sample duplication) were excluded, and animals and SNPs were excluded if the SNP call rate <90%, minor allele frequency (MAF) <5%, deviation from Hardy–Weinberg equilibrium detected (p < 10−6), or >10% genotype data missing. These criteria were assessed using PLINK software (ver. 1.07) [29]. After the exclusion of 94 cattle, 671,204 autosomal SNPs from 1252 cattle were analyzed.

2.4. Resequencing

Using genomic relationship data and Pi-Hat values, we chose 44 unrelated cattle for resequencing. The TIANamp Blood DNA kit (Tiangen Biotech Company Limited, Beijing, China) was used to extract genomic DNA. DNA concentrations and purity were quantified using the NanoPhotometer N50 device (Implen GmbH, Munich, Germany). DNA (1.5 μg) with an A260/280-nm absorbance ratio of 1.8–2 and A260/230-nm absorbance ratio of 2.0–2.2 was splintered using the Ultrasonicator S2 (Covaris, Woburn, MA, USA). Sequencing libraries were prepared using the Truseq Nano DNA HT sample preparation kit (Illumina Inc.) according to the manufacturer’s instructions of manufacturers. Each sample was given an index code for sequence recognition. The DNA samples were splintered into ≤350–base pair fragments by sonication, and DNA fragments were chosen for end polishing, addition of A-tailing, and sequencing with further PCR amplification using a full-length adapter. Sequencing libraries were prepared using the Illumina HiSeq 2500 system (Illumina Inc.). We obtained 9,621,765,847 reads and subjected them to quality control; low-quality reads with >10% unrecognized bases, >10% mismatches, or >50% poor-quality bases were removed. To facilitate down-stream analysis, SAM files were converted to BAM files. Then, we performed variant calling using samtools-0.1.19 mpileup [30]. Variants were excluded if an overall quality (QUAL) score of <20, a mapping quality (MQ) score of <30, and a read depth of <10. Moreover, we used the following proximity filters: if a variant was within 3 bp of another variant, the variant with the lower QUAL score was also excluded. Only bi-allele SNPs were retained for subsequent analyses. The average sequencing depth (representing the number of sequences for each base) of the samples was about 20×.

2.5. Variant Imputation

NSG-derived variants with MAFs > 0.05 (21,043,398 sequence variants from 44 animals) were subjected to imputation using BEAGLE (ver. 4.1) [31] with the default settings, based on algorithms determined by population data to infer genotypes for animals with missing information and haplotypes. We labeled genotypes for imputed sequence variants as 0 (homozygotes), 1 (heterozygotes), and 2 (alternative homozygotes). In total, 12,468,401 SNPs were obtained for chromosomes 1–29 from the RNA sequencing data with the use of the standard of imputation quality >0.1 [32].

2.6. Statistical Analysis

We employed a general mixed linear model based on the formula y = 1 μ + X b + m j b j + Z u + e , where y is a vector phenotypic value and μ represents the population mean. We applied two fixed-effect variables for single marker regression; b reflects noise associated with fixed effects (weight, sex, fattening days, and birth year) and bj reflects SNP effects. mj is the vector for the ith marker and u reflects the polygenic effect assumed with N (0, σ2K), where K corresponds to the kinship matrix and σ2 is the additive genetic variance. Although all SNPs on autosomal chromosomes were qualified for inclusion in the analysis, SNPs on chromosomes with mj were excluded. X and Z are the incidence matrices connecting the phenotypic values to fixed and polygenic effects, respectively. e represents random residual effects assumed with     V ( e ) = I σ e 2 , where I is an identity matrix and σ e 2 is the residual variance.
Associated SNPs were identified using GenABEL (ver. 1.8-0) [33] in R, with the application of the significance criterion of Bonferroni-corrected p < 0.05 in accordance with division by the number of SNPs. For carcass depth, dependable QTLs and putative candidate genes were identified with a stringent significance threshold of 7.45 × 10−8. To obtain p values for the SNPs, t values were calculated. Considering the fact that Bonferroni correction outcomes are too stringent with low statistical power [34], we determined the threshold values for GWAS by adopting false discovery rate (FDR). The FDR was defined as 0.01, and the threshold p value was computed based on the formula P = F D R × n / m , where n is the number of p < 0.01, and m represents the total number of SNPs [35].
To detect candidate genes of interest, we used the UCSC genome browser (http://www.genome.ucsc.edu, Bos_taurus_UMD_3.1, accessed on 2 October 2020).
NGS dataset analysis is complex due to the enormous number of markers, which can make conventional p-value computation for SNPs difficult [23,24]. We coped with this problem by dividing chromosomes into the following segments: segment 1 consisted of chromosomes 1–10 with 5,830,727 SNPs, segment 2 consisted of chromosomes 11–20 with 4,063,690 SNPs, and segment 3 consisted of chromosomes 21–29 with 2,573,984 SNPs. Dependable QTLs and putative candidate genes were detected using the significance thresholds of p = 8.58 × 10−9 for segment 1, p = 1.24 × 10−7 for segment 2, and p = 1.95 × 10−7 for segment 3, based on the number of markers in each segment.

2.7. Heritability Estimation

The heritabilities for carcass depth and hind leg circumference were estimated using GCTA v1.93.1 software in accordance with Bovine HD SNP array and imputed NGS. The genomic heritability of each trait was calculated using   h 2 = σ α   2 /   σ α 2 + σ e 2 .

3. Results

3.1. SNPs Associated with Carcass Depth

The estimated heritability for carcass depth was high (0.41 ± 0.06). With beadchip analysis, we identified several SNPs associated significantly with carcass depth (Table 2), including BovineHD0200025561 (p = 1.49 × 10−14) in candidate gene CLK1 on chromosome 2 and BovineHD0200025553 (p = 3.4 × 10−13) and BovineHD0200025550 (p = 8.3 × 10−13) in candidate gene BZW1. Exploration of the region that is about 10 kb upstream of the marker BovineHD0200025562 (p = 1.49 × 10−14) led to the identification of candidate gene PPIL3. The marker BovineHD0500023139 (p = 2.17 × 10−8) was found in the gene CCDC91 on chromosome 5. In addition, the candidate genes DIP2B and LARP4 were identified about 3 kb upstream and 24 kb downstream of the marker BovineHD0500008673 (p = 1.3 × 10−13), respectively. On chromosome 17, the gene c17h22orf31 was identified about 14 kb downstream of the SNP BovineHD1700020542 (p = 3.47 × 10−8). We also identified two significantly associated SNPs on chromosomes 10 (BovineHD1000031388; p = 1.17 × 10−8) and 26 (BovineHD2600013096; p = 5.58 × 10−8), but no associated candidate gene. Figure 2 shows the Manhattan plot for putative QTLs related to carcass depth.

3.2. SNPS Associated with Hind Leg Circumference

The estimated heritability for hind leg circumference was high (0.52 ± 0.12). Of the 9,621,765,847 raw reads acquired, 9,584,920,309 reads were examined after the application of quality control measures. In segment 1, we identified one SNP (rs91327124; p = 1.23 × 10−9) in BTA5 related to hind leg circumference, but no candidate gene in the vicinity of this SNP. In segment 2, several associated SNPs were identified (Table 3). The candidate gene ZBTB16 was identified about 50 kb downstream of the most strongly associated SNP in this segment (rs25168454; p = 1.31 × 10−10), NXPE4 was identified about 4.7 kb downstream of the SNP rs25439719 (p = 4.14 × 10−8), and the candidate gene RBM7 was found about 20 kb upstream of the SNP rs25237349 (p = 1.12 × 10−7).
In segment 3, we found many putative QTLs on chromosomes 23, 26, and 29 (Table 4). Exploration around flanking SNPs in BTA26 and BTA29 revealed no related candidate genes. The only associated SNP in BTA23 (rs37207517; p = 8.61 × 10−8) was located in the gene CDKAL1, which might be a crucial gene for Simmental beef cattle. About 47 kb upstream of this gene, the candidate gene E2F3 was identified. Figure 3 shows the Manhattan plots for putative QTLs in segments 1–3 related to hind leg circumference.

4. Discussion

GWASs have been conducted for many animals (e.g., [36]) and plant (e.g., [37]) species, contributing to a better understanding of molecular mechanisms underlying economically important traits in farm animals, including beef and dairy cattle [15,23,24,38,39,40,41,42,43,44]. GWASs are a powerful approach to the detection of genome regions that carry causative genes [45,46,47]. We anticipate that NGS technology will aid scientists’ detection of currently unknown genetic variants [48]. In this study, NGS analysis yielded no results related to carcass depth, but the use of an SNP array did yield such results; the opposite was true for hind leg circumference.
Beadchip analysis revealed that the CLK1 gene harbors the SNP BovineHD0200025561 in BTA2, associated most strongly with carcass depth. This gene codes for a protein kinase with binary specificity that phosphorylates tyrosine and serine/threonine residues [49]. Its N-terminal includes a regulatory portion containing receptors for EGF and other growth elements [50]. EGF plays a key role in cellular development by promoting cell proliferation and differentiation. CLK1 has been reported to play a role in cancer due to the associations among CLK1, ASF/SF2, and SRp20 [49] and in muscular dystrophy [51]. The CLK1 inhibitor, TG693, is a splicing modulator with therapeutic potential for patients with Duchenne muscular dystrophy [51]. In pigs, high expression levels of the full-length splice form of CLK1 were observed in the kidney, small intestine, and longissimus dorsi tissues [52]. In a rat model, CLK gene expression showed downregulation related to muscle fat during lactation [51]. In a gene ontology analysis, Lee et al. [53] detected a negative association between CLK1 and the intermuscular fat content in Hanwoo cattle. In another study of Hanwoo cattle, CLK1 was associated significantly with marbling (i.e., intramuscular fat) [54]. However, information on the effects of CLK1 on meat quality and its relationship to cellular (especially muscle and bone) development in Simmental beef cattle is lacking. We hypothesize that CLK1 plays a role related to meat quality and likely muscle cell development in Simmental beef cattle, but more research is needed to validate the relationship of this candidate gene to these traits.
In BTA2, we identified the candidate gene BZW1, which harbors two SNPs associated significantly with carcass depth (BovineHD0200025553 and BovineHD0200025550). This gene has been reported to be a proliferation regulator that promotes the progression of many aggressive cancers, such as salivary mucoepidermoid carcinoma [55]. BZW1 plays an important role in cell cycle processes and transcriptionally manages the histone H4 gene in the G1/S phase. Chiou et al. [56] reported that its downregulation can repress lung adenocarcinoma metastasis. The roles of BZW1 in tumor growth [57] and Alzheimer’s disease progression [58] have also been well demonstrated. However, very few studies have examined the function of this gene in farm animals. BZW1 has been associated with bovine endometriosis [59], and Hoelker et al. [59] showed that its downregulation could influence the dynamic progression of embryos from cattle with subclinical endometritis. We hypothesize that BZW1 is associated with carcass depth in Simmental beef cattle. More research is warranted to validate this association. We also identified the gene PPIL3 about 10 kb upstream of the SNP BovineHD0200025562. PPIL3 has been associated with estrogen receptor-negative breast cancer [60], but information on its function in farm animals is lacking. Thus, our identification of the potential association of PPIL3 with carcass depth in Simmental cattle is novel. Further research is needed to corroborate this association.
In BTA5, we identified the candidate gene CCDC91, which harbors the SNP BovineHD0500023139, associated significantly with carcass depth. This gene is expressed in numerous cancer cell lines. CCDC91 has been identified as a candidate gene associated with production traits (fat percentage) in dairy cattle [61], but few studies have examined this gene in farm animals. We hypothesize that this gene is associated with carcass depth in Simmental beef cattle. Moreover, we identified the candidate gene DIP2B about 3 kb downstream of the SNP BovineHD0500008673. In cattle, this gene has been associated with fatty acid content [62], an important trait affecting meat quality. DIP2B has also been associated with lipid metabolism, skeletal system development, and insulin signaling [62]. Information on its associations with traits in Simmental beef cattle is lacking; we hypothesize that it is related to carcass depth. We also identified the candidate gene LARP4 about 24 kb upstream of the SNP BovineHD0500008673. This gene may be able to suppress cancer cell migration and invasion [63]. Egiz et al. [64] reported that LARP4 repressed the motility and metastatic development of ovarian cancer cells. This gene has also been shown to control mRNA solidity and translation [65]. Zhou et al. [66] identified several candidate genes, including LARP4B, that influence milk protein composition traits in Holstein cattle. We hypothesize that LARP4 is associated with carcass depth in Simmental beef cattle.
In the NGS analysis, we found that the SNP rs25168454 in BTA15 was associated most strongly with hind leg circumference. The candidate gene ZBTB16 was located about 50 kb downstream of this SNP in segment 2. This conserved gene, broadly expressed in normal tissues such as cardiac and skeletal muscle [67,68] and also recognized as PLZF and Zfp145, acts as a crucial transcriptional suppressor or activator [69] that is required for various biological and developmental processes, including cellular apoptosis [70], hind limb construction [71], spermatogenesis [72], and anti-tumorigenesis activities [73]. It is also involved in processes such as metabolism regulation [74] and mesenchymal stem cell (MSC) differentiation [75], which is important during bone regeneration. In addition, ZBTB16 controls homeostasis and the fate of hematopoietic stem cells [76]. Epigenomics and transcriptomics research has revealed its role in adipogenesis, as it controls fat accumulation [77,78], and it has been found to control thermogenics in brown fat and muscle [79]. Wei et al. [80] reported that the overexpression of ZBTB16 enhances white adipogenesis and brown-like adipocyte formation in intramuscular preadipocytes in cattle. The bovine hind leg contains important muscles and bones, and its circumference is a good criterion for the assessment of traits related to bone and muscle development. Thus, we hypothesize that ZBTB16 plays a role in bone and muscle development in Simmental beef cattle, but more research is needed to validate this hypothesis. In BTA15, about 4.7 kb downstream of SNP rs25439719, we found the candidate gene NXPE4. This gene has been recognized as a potentially promising prognostic biomarker for colorectal cancer [81]. Sarghale et al. [82] reported that NXPE4 was a candidate gene located near SNPs related most closely to methane emission and fat traits in Holstein cattle. The candidate gene RBM7 was also identified in BTA15, almost 20 kb upstream of the SNP rs25237349. The literature contains little information about this gene. A GWAS revealed that it was a candidate gene associated with milk production in Iranian buffalo [83]. We hypothesize that NXPE4 and RBM7 are associated with hind leg circumference in Simmental beef cattle.
In segment 3, we identified the very promising candidate gene CDKAL1, which harbors the only SNP associated with hind leg circumference in BTA23 (rs37207517). Several GWASs have revealed the role of CDKAL1 in type 2 diabetes [84,85,86]. In addition, CDKAL1 encodes the CDK5 regulatory subunit–associated protein 1-like 1. This protein results from the translation of CDKAL1 and is homologous to cyclin-dependent kinase 5 regulatory subunit–associated protein, a neuronal protein that can hamper CDK5 activation [87]. CDK5 can act as a myogenic protein kinase; it is required for myoblast differentiation and thus plays a key role in muscle myogenesis [88]. CDKAL1 has also been associated with mitochondrial morphology and adipogenesis [89]. Thus, we hypothesize that it is associated with muscle development and hind leg circumference in Simmental beef cattle. We identified E2F3, another important candidate gene, about 47 kb upstream of CDKAL1. As a promoter of the muscle cell cycle, E2F is related to myogenesis [88]. Importantly, E2F3 can induce myogenic differentiation [90]. Ma et al. [91] reported that microRNA-432 could hamper myogenesis by targeting E2F3 in muscle cells. E2F3 is associated with various other processes and plays an important role in the progression of a variety of cancers, such as bladder, breast, and prostate cancers [92,93,94]. Information about E2F3 in Simmental beef cattle is lacking; we hypothesize that this gene is related to myogenesis in this breed.

5. Conclusions

Using the Illumina Bovine HD Beadchip and NGS dataset analyses, we identified many putative QTLs on several chromosomes that were associated with carcass traits in 1252 Simmental beef cattle. Several candidate genes related to carcass depth and hind leg circumference were identified. Among them, CDKAL1 and E2F3 may be associated with myogenesis and muscle development. Although the data were obtained by analyzing the genome of the Simmental breed, they can also be extended and validated in other cattle breeds. The identification of such genes helps breeders to understand and improve traits such as carcass depth and hind leg circumference.

Author Contributions

Statistical analyses, writing—original draft preparation and Writing—review and editing, F.B.; validation and supervision, J.L.; conceptualization and project administration, M.M.; conceptualization and writing—review and editing, L.X. and L.Z.; validation and writing—review and editing, J.J. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Agricultural Science and Technology Innovation Program in the Chinese Academy of Agricultural Sciences (ASTIP-IAS03 and CAAS-ZDRW202102), China Agriculture Research System of MOF and MARA and the National Beef Cattle Industrial Technology System (CARS-37). The project was also partly supported by the Science and Technology Project of Inner Mongolia Autonomous Region (2020GG0210 and KJXM2020002-01) for the data analysis and interpretation of the study.

Institutional Review Board Statement

All procedures were conducted in accordance with the guidelines of China’s Council on Animal Care, and the study protocol was approved by the Institute of An-imal Science of the Chinese Academy of Agricultural Science, Beijing, China (approval number: RNL09/07). Analyses associated with animals adhered to the rules of China’s Council on Animal Welfare.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data underlying this study have been uploaded to Dryad. The raw genotype data are accessible using the following doi: 10.5061/dryad.4qc06.

Conflicts of Interest

The authors declare that they have no competing interest.

Abbreviations

EGF: epidermal growth factor; GWAS: genome-wide association study; MAF: minor allele frequency; NGS: next-generation sequencing; QTL: quantitative trait locus; SNP: single-nucleotide polymorphism; MSC: mesenchymal stem cells.

References

  1. Seré, C.; Steinfeld, H. World Livestock Production Systems; Animal Production and Health Paper 127; FAO: Rome, Italy, 1996; Available online: http://www.fao.org/3/a-w0027e.pdf (accessed on 7 July 2021).
  2. Harris, D.L. Breeding for Efficiency in Livestock Production: Defining the Economic Objectives. J. Anim. Sci. 1970, 30, 860–865. [Google Scholar] [CrossRef] [Green Version]
  3. Yuan, Z.; Li, J.; Li, J.; Gao, X.; Gao, H.; Xu, S. Effects of DGAT1 gene on meat and carcass fatness quality in Chinese commercial cattle. Mol. Biol. Rep. 2012, 40, 1947–1954. [Google Scholar] [CrossRef] [PubMed]
  4. Albertí, P.; Panea, B.; Sañudo, C.; Olleta, J.L.; Ripoll, G.; Ertbjerg, P.; Christensen, S.; Gigli, S.; Failla, S.; Concetti, J.F.; et al. Live weight, body size and carcass characteristics of young bulls of fifteen European breeds. Livest. Sci. 2008, 114, 19–30. [Google Scholar] [CrossRef]
  5. Chambaz, A.; Scheeder, M.R.; Kreuzer, M.; Dufey, P.A. Meat quality of Angus, Simmental, Charolais and Limousin steers compared at the same intramuscular fat content. Meat Sci. 2003, 63, 491–500. [Google Scholar] [CrossRef]
  6. Cuvelier, C.; Cabaraux, J.F.; Dufrasne, I.; Clinquart, A.; Hocquette, J.F.; Istasse, L.; Hornick, J.L. Performance, slaughter characteristics and meat quality of young bulls from Belgian Blue, Limousin and Aberdeen Angus breeds fattened with a sugar-beet pulp or a cereal-based diet. Anim. Sci. 2006, 82, 125–132. [Google Scholar] [CrossRef]
  7. Vieira, C.; Cerdeño, A.; Serrano, E.; Lavín, P.; Mantecón, A.R. Breed and ageing extent on carcass and meat quality of beef from adult steers (oxen). Livest. Sci. 2007, 107, 62–69. [Google Scholar] [CrossRef]
  8. May, M.L.; Dikeman, M.E.; Schalles, R. Longissimus Muscle Histological Characteristics of Simmental × Angus, Hereford × Angus and Limousin × Angus Crossbred Steers as Related to Carcass Composition and Meat Palatability Traits. J. Anim. Sci. 1977, 44, 571–580. [Google Scholar] [CrossRef]
  9. Ozlütürk, A.; Tüzemen, N.; Yanar, M.; Esenbuga, N.; Dursun, E. Fattening performance, carcass traits and meat quality characteristics of calves sired by Charolais, Simmental and Eastern Anatolian Red sires mated to Eastern Anatolian Red dams. Meat Sci. 2004, 67, 463–470. [Google Scholar] [CrossRef]
  10. Sami, A.S.; Augustini, C.; Schwarz, F.J. Effects of feeding intensity and time on feed on performance, carcass characteristics and meat quality of Simmental bulls. Meat Sci. 2004, 67, 195–201. [Google Scholar] [CrossRef]
  11. Kebede, A.; Komlosi, I. Evaluation of genetic parameters and growth traits of Hungarian Simmental cattle breed. Livest. Res. 2015, 27, Article #172. Available online: http://www.lrrd.org/lrrd27/9/dami27172.html (accessed on 8 July 2021).
  12. Tabor, H.K.; Risch, N.J.; Myers, R.M. Candidate-gene approaches for studying complex genetic traits: Practical considerations. Nat. Rev. Genet. 2002, 3, 391–397. [Google Scholar] [CrossRef] [PubMed]
  13. Zhu, M.; Zhao, S. Candidate Gene Identification Approach: Progress and Challenges. Int. J. Biol. Sci. 2007, 3, 420–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Xia, J.; Qi, X.; Wu, Y.; Zhu, B.; Xu, L.; Zhang, L.; Gao, X.; Chen, Y.; Li, J.; Gao, H. Genome-wide association study identifies loci and candidate genes for meat quality traits in Simmental beef cattle. Mamm. Genome 2016, 27, 246–255. [Google Scholar] [CrossRef] [PubMed]
  15. Chang, T.; Xia, J.; Xu, L.; Wang, X.; Zhu, B.; Zhang, L.; Gao, X.; Chen, Y.; Li, J.; Gao, H. A genome-wide association study suggests several novel candidate genes for carcass traits in Chinese Simmental beef cattle. Anim Genet 2018, 49, 312–316. [Google Scholar] [CrossRef]
  16. Du, L.; Duan, X.; An, B.; Chang, T.; Liang, M.; Xu, L.; Zhang, L.; Li, J.; E, G.; Gao, H. Genome-Wide Association Study Based on Random Regression Model Reveals Candidate Genes Associated with Longitudinal Data in Chinese Simmental Beef Cattle. Animals 2021, 11, 2524. [Google Scholar] [CrossRef]
  17. Hirschhorn, J.N.; Daly, M.J. Genome-wide association studies for common diseases and complex traits. Nat. Rev. Genet. 2005, 6, 95–108. [Google Scholar] [CrossRef]
  18. Coon, K.D.; Myers, A.J.; Craig, D.W.; Webster, J.A.; Pearson, J.V.; Lince, D.H.; Zismann, V.L.; Beach, T.G.; Leung, D.; Bryden, L.; et al. A high-density whole-genome association study reveals that APOE is the major susceptibility gene for sporadic late-onset Alzheimer’s disease. J. Clin. Psychiatry 2007, 68, 613–618. [Google Scholar] [CrossRef]
  19. Gudmundsson, J.; Sulem, P.; Gudbjartsson, D.F.; Blondal, T.; Gylfason, A.; Agnarsson, B.A.; Benediktsdottir, K.R.; Magnusdottir, D.N.; Orlygsdottir, G.; Jakobsdottir, M.; et al. Genome-wide association and replication studies identify four variants associated with prostate cancer susceptibility. Nat. Genet. 2009, 41, 1122–1126. [Google Scholar] [CrossRef] [Green Version]
  20. Visscher, P.M.; Wray, N.R.; Zhang, Q.; Sklar, P.; McCarthy, M.I.; Brown, M.A.; Yang, J. 10 years of GWAS discovery: Biology, function, and Translation. Am. J. Hum. Genet. 2017, 101, 5–22. [Google Scholar] [CrossRef] [Green Version]
  21. Sharma, A.; Cho, Y.; Choi, B.H.; Chai, H.H.; Park, J.E.; Lim, D. Limited representation of OMIA causative mutations for cattle in SNP databases. Anim. Genet. 2017, 48, 369–370. [Google Scholar] [CrossRef]
  22. Shin, D.H.; Lee, H.J.; Cho, S.; Kim, H.J.; Hwang, J.Y.; Lee, C.K.; Jeong, J.; Yoon, D.; Kim, H. Deleted copy number variation of Hanwoo and Holstein using next generation sequencing at the population level. BMC Genom. 2014, 15, 240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Bordbar, F.; Jensen, J.; Zhu, B.; Wang, Z.; Xu, L.; Chang, T.; Xu, L.; Du, M.; Zhang, L.; Gao, H.; et al. Identification of muscle-specific candidate genes in Simmental beef cattle using imputed next generation sequencing. PLoS ONE 2019, 14, e0223671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bordbar, F.; Jensen, J.; Du, M.; Abied, A.; Guo, W.; Xu, L.; Gao, H.; Zhang, L.; Li, J. Identification and validation of a novel candidate gene regulating net meat weight in Simmental beef cattle based on imputed next-generation sequencing. Cell Prolif. 2020, 53, e12870. [Google Scholar] [CrossRef] [PubMed]
  25. Choi, J.W.; Chung, W.H.; Lee, K.T.; Cho, E.S.; Lee, S.W.; Choi, B.H.; Lee, S.H.; Lim, W.; Lim, D.; Lee, Y.G.; et al. Whole-genome resequencing analyses of five pig breeds, including Korean wild and native, and three European origin breeds. DNA Res. 2015, 22, 259–267. [Google Scholar] [CrossRef] [Green Version]
  26. Piedrafita, J.; Quintanilla, R.; Sanudo, C.; Olleta, J.L.; Campo, M.M.; Panea, B.; Renand, G.; Turin, F.; Jabet, S.; Osoroe, K.; et al. Carcass quality of 10 beef cattle breeds of the Southwest of Europe in their typical production systems. Livest. Prod. Sci. 2003, 82, 1–13. [Google Scholar] [CrossRef] [Green Version]
  27. Han, R.; Zan, L.; Yang, D.; Hao, R. [SNPs detection of IGF2 gene and its relationship with carcass and meat quality traits in Qinchuan cattle]. Yi Chuan 2008, 30, 1579–1584. [Google Scholar] [CrossRef]
  28. Mohammadabadi, M.; Bordbar, F.; Jensen, J.; Du, M.; Guo, W. Key Genes Regulating Skeletal Muscle Development and Growth in Farm Animals. Animals 2021, 11, 835. [Google Scholar] [CrossRef]
  29. Purcell, S.; Neale, B.; Todd-Brown, K.; Thomas, L.; Ferreira, M.A.R.; Bender, D.; Maller, J.; Sklar, P.; Bakker, P.; Daly, M.J.; et al. PLINK: A tool set for whole-genome association and population-based linkage analyses. Am. J. Hum. Genet. 2007, 81, 559–575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Li, H. A statistical framework for SNP calling, mutation discovery, association mapping and population genetical parameter estimation from sequencing data. Bioinformatics 2011, 27, 2987–2993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Browning, B.L.; Browning, S.R. Genotype imputation with millions of reference samples. Am. J. Hum. Genet. 2016, 98, 116–126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Fuchsberger, C.; Abecasis, G.R.; Hinds, D.A. Minimac2: Faster genotype imputation. Bioinformatics 2014, 31, 782–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Aulchenko, Y.S.; Ripke, S.; Isaacs, A.; Duijn, C. GenABEL: An R library for genome-wide association analysis. Bioinformatics 2007, 23, 1294–1296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Nakagawa, S. A farewell to Bonferroni: The problems of low statistical power and publication bias. Behav. Ecol. 2004, 15, 1044–1045. [Google Scholar] [CrossRef]
  35. Benjamini, Y.; Hochberg, Y. Controlling the false discovery rate: A practical and powerful approach to multiple testing. J. R. Stat. Soc. Ser. B 1995, 57, 289–300. [Google Scholar] [CrossRef]
  36. Takasuga, A.; Watanabe, T.; Mizoguchi, Y.; Hirano, T.; Ihara, N.; Takano, A.; Yokouchi, K.; Fujikawa, A.; Chiba, K.; Kobayashi, N.; et al. Identification of bovine QTL for growth and carcass traits in Japanese Black cattle by replication and identical-by-descent mapping. Mamm. Genome 2017, 18, 125–136. [Google Scholar] [CrossRef] [PubMed]
  37. Li, F.; Chen, B.; Xu, K.; Wu, J.; Song, W.; Bancroft, I.; Harper, A.L.; Trick, M.; Liu, G.; Gao, G.; et al. Genome-wide association study dissects the genetic architecture of seed weight and seed quality in rapeseed (Brassica napus L.). DNA Res. 2014, 21, 355–367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Cole, J.B.; Wiggans, G.R.; Ma, L.; Sonstegard, T.S.; Lawlor, T.J.; Crooker, B.A.; Van Tassell, C.P.; Yang, J.; Wang, S.; Lakshmi, K.; et al. Genome-wide association analysis of thirty one production, health, reproduction and body conformation traits in contemporary U.S. Holstein cows. BMC Genom. 2011, 12, 408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Bolormaa, S.; Hayes, B.J.; Savin, K.; Hawken, R.; Barendse, W.; Arthur, P.F.; Herd, R.M.; Goddard, M.E. Genome-wide association studies for feedlot and growth traits in cattle. J. Anim. Sci. 2011, 89, 1684–1697. [Google Scholar] [CrossRef] [Green Version]
  40. Bolormaa, S.; Pryce, J.E.; Reverter, A.; Zhang, Y.; Barendse, W.; Kemper, K.; Tier, B.; Savin, K.; Hayes, B.; Goddard, M. A multi-trait, meta-analysis for detecting pleiotropic polymorphisms for stature, fatness and reproduction in beef cattle. PLoS Genet. 2014, 10, e1004198. [Google Scholar] [CrossRef] [Green Version]
  41. Wu, X.; Fang, M.; Liu, L.; Wang, S.; Liu, J.; Ding, X.; Zhang, S.; Zhang, Q.; Zhang, Y.; Qiao, L.; et al. Genome wide association studies for body conformation traits in the Chinese Holstein cattle population. BMC Genom. 2013, 14, 897. [Google Scholar] [CrossRef] [Green Version]
  42. Abo-Ismail, M.K.; Vander Voort, G.; Squires, J.J.; Swanson, K.C.; Mandell, I.B.; Liao, X.; Stothard, P.; Moore, S.; Plastow, G.; Miller, S. Single nucleotide polymorphisms for feed efficiency and performance in crossbred beef cattle. BMC Genet. 2014, 15, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Feugang, J.M.; Kaya, A.; Page, G.P.; Chen, L.; Mehta, T.; Hirani, K.; Nazareth, L.; Topper, E.; Gibbs, R.; Memili, E. Two-stage genome-wide association study identifies integrin beta 5 as having potential role in bull fertility. BMC Genom. 2009, 10, 176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Duijvesteijn, N.; Knol, E.F.; Merks, J.W.; Crooijmans, R.P.; Groenen, M.A.; Bovenhuis, H.; Harlizius, B. A genome-wide association study on androstenone levels in pigs reveals a cluster of candidate genes on chromosome 6. BMC Genet. 2010, 11, 42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Barendse, W.; Reverter, A.; Bunch, R.J.; Harrison, B.E.; Barris, W.; Thomas, M.B. A validated whole-genome association study of efficient food conversion in cattle. Genetics 2007, 176, 1893–1905. [Google Scholar] [CrossRef] [Green Version]
  46. Kolbehdari, D.; Wang, Z.; Grant, J.R.; Murdoch, B.; Prasad, A.; Xiu, Z.; Marques, E.; Stothard, P.; Moore, S. A whole genome scan to map QTL for milk production traits and somatic cell score in Canadian Holstein bulls. J. Ani. Breed. Genet. 2009, 126, 216–227. [Google Scholar] [CrossRef]
  47. Settles, M.; Zanella, R.; McKay, S.D.; Schnabel, R.D.; Taylor, J.F.; Whitlock, R.; Schukken, Y.; Van Kessel, J.S.; Smith, J.M.; Neibergs, H. A Whole Genome Association analysis identifies loci associated with Mycobacterium avium subsp. paratuberculosis infection status in US Holstein Cattle. Anim. Genet. 2009, 40, 655–662. [Google Scholar] [CrossRef]
  48. Reis-Filho, J.S. Next-generation sequencing. Breast Cancer Res. 2009, 11, S12. [Google Scholar] [CrossRef] [Green Version]
  49. Doherty, S.; Maurer, J. CLK1 and Its Effects on Skin Stem Cell Differentiation. Bachelor’s Thesis, Worcester Polytechnic Institute, Worcester, MA, USA, 29 April 2010. [Google Scholar]
  50. Menegay, H.J.; Myers, M.P.; Moeslein, F.M.; Landreth, G.E. Biochemical Characterization and Localization of the Dual Specificity Kinase CLK1. J. Cell Sci. 2000, 113, 3241–3253. [Google Scholar] [CrossRef]
  51. Sako, Y.; Ninomiya, K.; Okuno, Y.; Toyomoto, M.; Nishida, A.; Koike, Y.; Ohe, K.; Kii, I.; Yoshida, S.; Hashimoto, N.; et al. Development of an orally available inhibitor of CLK1 for skipping a mutated dystrophin exon in Duchenne muscular dystrophy. Sci. Rep. 2017, 7, 46126. [Google Scholar] [CrossRef] [Green Version]
  52. Xie, S.H.; Li, J.Q.; Chen, Y.S.; Gao, P.; Zhang, H.; Li, Z.Y. Molecular Cloning, Expression, and Chromosomal Mapping of the Porcine CDC-2-Like Kinase 1 (CLK1) Gene. Biochem. Genet. 2009, 47, 266–275. [Google Scholar] [CrossRef]
  53. Lee, S.; Gondro, G.; van der Werf, J.; Kim, N.; Lim, D.; Park, E.; Oh, S.; Gibson, J.P.; Thompson, J.M. Use of a bovine genome array to identify new biological pathways for beef marbling in Hanwoo (Korean Cattle). BMC Genom. 2010, 11, 623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Lim, D.; Lee, S.; Kim, N.; Cho, Y.; Chai, H.; Seong, H.; Kim, H. Gene Co-expression Analysis to Characterize Genes Related to Marbling Trait in Hanwoo (Korean) Cattle. Asian-Aust. J. Anim. Sci. 2013, 26, 19–29. [Google Scholar] [CrossRef]
  55. Li, S.; Chai, Z.; Li, Y.; Liu, D.; Bai, Z.; Li, Y.; Li, Y.; Situ, Z. BZW1, a novel proliferation regulator that promotes growth of salivary muocepodermoid carcinoma. Cancer Lett. 2009, 284, 86–94. [Google Scholar] [CrossRef] [PubMed]
  56. Chiou, J.; Chang, Y.; Jan, Y.; Tsai, H.; Yang, C.; Huang, M.; Yu, Y.; Hsiao, M. Overexpression of BZW1 is an independent poor prognosis marker and its down-regulation suppresses lung adenocarcinoma metastasis. Sci. Rep. 2019, 9, 14624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Shi, Z.; Xiao, C.; Lin, T.; Wu, J.; Li, K. BZW1 promotes cell proliferation in prostate cancer by regulating TGF-β1/Smad pathway. Cell Cycle 2021, 20, 894–902. [Google Scholar] [CrossRef] [PubMed]
  58. Wang, W.; Mandel, J.; Bouaziz, B.; Commenges, D.; Nabirotchkine, S.; Chumakov, I.; Cohen, D.; Guedj, M.; the Alzheimer’s Disease Neuroimaging Initiative. A Multi-Marker Genetic Association Test Based on the Rasch Model Applied to Alzheimer’s Disease. PLoS ONE 2015, 10, e0138223. [Google Scholar] [CrossRef] [PubMed]
  59. Hoelker, M.; Salilew-Wondim, D.; Drillich, M.; Christine, G.B.; Ghanem, N.; Goetze, L.; Tesfaye, D.; Schellander, K.; Heuwieser, W. Transcriptional response of the bovine endometrium and embryo to endometrial polymorphonuclear neutrophil infiltration as an indicator of subclinical inflammation of the uterine environment. Reprod. Fertil. Dev. 2012, 24, 778–793. [Google Scholar] [CrossRef]
  60. Sarin, K.Y.; Lin, Y.; Daneshjou, R.; Ziyatdinov, A.; Thorleifsson, G.; Rubin, A.; Pardo, L.M.; Wu, W.; Khavari, P.A.; Uitterlinden, A.; et al. Genome-wide meta-analysis identifies eight new susceptibility loci for cutaneous squamous cell carcinoma. Nat. Commun. 2020, 11, 820. [Google Scholar] [CrossRef]
  61. Nayeri, S.; Stothard, P. Tissues, Metabolic Pathways and Genes of Key Importance in Lactating Dairy Cattle. Springer Sci. Rev. 2016, 4, 49–77. [Google Scholar] [CrossRef] [Green Version]
  62. Oliveira, P.S.N.; Cesar, A.S.M.; Oliveira, G.B.; Tizioto, P.C.; Poleti, M.D.; Diniz, W.J.S.; Lima, A.O.D.; Reecy, J.M.; Coutinho, L.L.; Regitano, L.C.A. miRNAs related to fatty acids composition in Nellore cattle. J. Anim. Sci. 2016, 94, 164. [Google Scholar] [CrossRef] [Green Version]
  63. Seetharaman, S.; Flemyng, E.; Shen, J.; Conte, M.R.; Ridley, A.J. The RNA-binding protein LARP4 regulates cancer cell migration and invasion. Cytoskeleton 2016, 73, 680–690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Egiz, M.; Usui, T.; Ishibashi, M.; Zhang, X.; Shigeta, S.; Toyoshima, M.; Kitatani, K.; Yaegashi, N. La-Related Protein 4 as a Suppressor for Motility of Ovarian Cancer Cells. Tohoku J. Exp. Med. 2019, 247, 59–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Blagden, S.P.; Gatt, M.K.; Archambault, V.; Lada, K.; Ichihara, K.; Lilley, K.S.; Inoue, Y.H.; Glover, D.M. Drosophila Larp associates with poly(A)-binding protein and is required for male fertility and syncytial embryo development. Dev. Biol. 2009, 334, 186–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Zhou, C.; Li, C.; Cai, W.; Liu, S.; Yin, H.; Shi, S.; Zhang, Q.; Zhang, S. Genome-Wide Association Study for Milk Protein Composition Traits in a Chinese Holstein Population Using a Single-Step Approach. Front. genet. 2019, 10, 72. [Google Scholar] [CrossRef] [Green Version]
  67. Suliman, B.A.; Xu, D.; Williams, B.R. The promyelocytic leukemia zinc finger protein: Two decades of molecular oncology. Front. Oncol. 2012, 2, 74. [Google Scholar] [CrossRef] [Green Version]
  68. Uhlen, M.; Bjorling, E.; Agaton, C.; Szigyarto, C.A.; Amini, B.; Andersen, E.; Angelidou, P.; Asplund, A.; Asplund, C.; Berglund, L.; et al. A human protein atlas for normal and cancer tissues based on antibody proteomics. Mol. Cell Proteomic. 2005, 4, 1920–1932. [Google Scholar] [CrossRef] [Green Version]
  69. Kolesnichenko, M.; Vogt, P.K. Understanding PLZF: Two transcriptional targets, REDD1 and smooth muscle alpha-actin, define new questions in growth control, senescence, self-renewal and tumor suppression. Cell Cycle 2011, 10, 771–775. [Google Scholar] [CrossRef] [Green Version]
  70. Cheung, M.; Pei, J.; Pei, Y.; Jhanwar, S.C.; Pass, H.I.; Testa, J.R. The promyelocytic leukemia zinc-finger gene, PLZF, is frequently downregulated in malignant mesothelioma cells and contributes to cell survival. Oncogene 2010, 29, 1633–1640. [Google Scholar] [CrossRef] [Green Version]
  71. Barna, M.; Hawe, N.; Niswander, L.; Pandolfi, P.P. Plzf regulates limb and axial skeletal patterning. Nat. Genet. 2000, 25, 166–172. [Google Scholar] [CrossRef]
  72. Costoya, J.A.; Hobbs, R.M.; Barna, M.; Cattoretti, G.; Manova, K.; Sukhwani, M.; Orwig, K.E.; Wolgemuth, D.J.; Pandolfi, P.P. Essential role of Plzf in maintenance of spermatogonial stem cells. Nat. Genet. 2004, 36, 653–659. [Google Scholar] [CrossRef] [Green Version]
  73. Felicetti, F.; Bottero, L.; Felli, N.; Mattia, G.; Labbaye, C.; Alvino, E.; Peschle, C.; Colombo, M.P.; Care, A. Role of PLZF in melanoma progression. Oncogene 2004, 23, 4567–4576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Chen, S.; Qian, J.; Shi, X.; Gao, T.; Liang, T.; Liu, C. Control of hepatic gluconeogenesis by the promyelocytic leukemia zinc finger protein. Mol. Endocrinol. 2014, 28, 1987–1998. [Google Scholar] [CrossRef] [PubMed]
  75. Marofi, F.; Vahedi, G.; Solali, S.; Alivand, M.; Salarinasab, S.; Heydarabad, M.Z.; Farshdousti Hagh, M. Gene expression of TWIST1 and ZBTB16 is regulated by methylation modifications during the osteoblastic differentiation of mesenchymal stem cells. J. Cell Physiol. 2019, 234, 6230–6234. [Google Scholar] [CrossRef] [PubMed]
  76. Vincent-Fabert, C.; Platet, N.; Vandevelde, A.; Poplineau, M.; Koubi, M.; Finetti, P.; Tiberi, G.; Imbert, A.M.; Bertucci, F.; Duprez, E. PLZF mutation alters mouse hematopoietic stem cell function and cell cycle progression. Blood 2016, 127, 1881–1885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Mikkelsen, T.S.; Xu, Z.; Zhang, X.; Wang, L.; Gimble, J.M.; Lander, E.S.; Rosen, E.D. Comparative epigenomic analysis of murine and human adipogenesis. Cell 2010, 143, 156–169. [Google Scholar] [CrossRef] [Green Version]
  78. Ambele, M.A.; Dessels, C.; Durandt, C.; Pepper, M.S. Genome-wide analysis of gene expression during adipogenesis in human adipose-derived stromal cells reveals novel patterns of gene expression during adipocyte differentiation. Stem Cell Res. 2016, 16, 725–734. [Google Scholar] [CrossRef] [Green Version]
  79. Plaisier, C.L.; Bennett, B.J.; He, A.; Guan, B.; Lusis, A.J.; Reue, K.; Vergnes, L. Zbtb16 has a role in brown adipocyte bioenergetics. Nutr. Diabetes 2012, 2, e46. [Google Scholar] [CrossRef] [Green Version]
  80. Wei, S.; Zhang, M.; Zheng, Y.; Yan, P. ZBTB16 Overexpression Enhances White Adipogenesis and Induces Brown-Like Adipocyte Formation of Bovine White Intramuscular Preadipocytes. Cell Physiol. Biochem. 2018, 48, 2528–2538. [Google Scholar] [CrossRef]
  81. Liu, Y.; Hu, Y.; Zeng, Y.; Li, Z.; Zhang, H.; Deng, J.; Wang, G. Neurexophilin and PC-esterase domain family member 4 (NXPE4) and prostate androgen-regulated mucin-like protein 1 (PARM1) as prognostic biomarkers for colorectal cancer. J. Cell Biochem. 2019, 120, 18041–18052. [Google Scholar] [CrossRef]
  82. Sarghale, A.J.; Shahrebabak, M.M.; Shahrebabak, H.M.; Javaremi, A.N.; Saatchi, M.; Khansefid, M.; Miar, Y. Genome-wide association studies for methane emission and ruminal volatile fatty acids using Holstein cattle sequence data. BMC Genet. 2020, 21, 129. [Google Scholar]
  83. Mokhber, M. Genome-wide association study for milk production in Iranian buffalo. In Proceedings of the Conference on organic vs. Conventional agriculture, Ardabil, Iran, 16–17 August 2017. [Google Scholar]
  84. Steinthorsdottir, V.; Thorleifsson, G.; Reynisdottir, I.; Benediktsson, R.; Jonsdottir, T.; Walters, G.B.; Styrkarsdottir, U.; Gretarsdottir, S.; Emilsson, V.; Ghosh, S.; et al. A variant in CDKAL1 influences insulin response and risk of type 2 diabetes. Nat. Genet. 2007, 39, 770–775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Zeggini, E.; Weedon, M.N.; Lindgren, C.M.; Frayling, T.M.; Elliott, K.S.; Lango, H.; Timpson, N.J.; Perry, J.R.B.; Rayner, N.W.; Freathy, R.M.; et al. Replication of genome-wide association signals in UK samples reveals risk loci for type 2 diabetes. Science 2007, 316, 1336–1341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Scott, L.J.; Mohlke, K.L.; Bonnycastle, L.L.; Willer, C.J.; Li, Y.; Duren, W.L.; Erdos, M.R.; Stringham, H.M.; Chines, P.S.; Jackson, A.U.; et al. A genome-wide association study of type 2 diabetes in Finns detects multiple susceptibility variants. Science 2007, 316, 1341–1345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Lew, J.; Huang, Q.Q.; Qi, Z.; Winkfein, R.J.; Aebersold, R.; Hunt, T.; Wang, J.H. A brain-specific activator of cyclin-dependent kinase 5. Nature 1994, 371, 423–426. [Google Scholar] [CrossRef] [PubMed]
  88. Knight, J.D.R.; Kothary, R. The myogenic kinome: Protein kinases critical to mammalian skeletal myogenesis. Skelet. Muscle 2011, 1, 29. [Google Scholar] [CrossRef] [Green Version]
  89. Palmer, C.J.; Bruckner, R.J.; Paulo, J.A.; Kazak, L.; Long, J.Z.; Mina, A.I.; Deng, Z.; LeClair, K.B.; Hall, J.A.; Hong, S.; et al. Cdkal1, a type 2 diabetes susceptibility gene, regulates mitochondrial function in adipose tissue. Mol. Metab. 2017, 6, 1212–1225. [Google Scholar] [CrossRef]
  90. Asp, P.; Acosta-Alvear, D.; Tsikitis, M.; van Oevelen, C.; Dynlacht, B.D. E2f3b plays an essential role in myogenic differentiation through isoform-specific gene regulation. Gen. Dev. 2009, 23, 37–53. [Google Scholar] [CrossRef] [Green Version]
  91. Ma, M.; Wang, X.; Chen, X.; Cai, R.; Chen, F.; Dong, W.; Yang, G.; Pang, W. MicroRNA-432 targeting E2F3 and P55PIK inhibits myogenesis through PI3K/AKT/mTOR signaling pathway. RNA Biol. 2017, 14, 347–360. [Google Scholar] [CrossRef] [Green Version]
  92. Foster, C.S.; Falconer, A.; Dodson, A.R.; Norman, A.R.; Dennis, N.; Fletcher, A.; Southgate, C.; Dowe, A.; Dearnaley, D.; Jhavar, S.; et al. Transcription factor E2F3 overexpressed in prostate cancer independently predicts clinical outcome. Oncogene 2004, 23, 5871–5879. [Google Scholar] [CrossRef] [Green Version]
  93. Feber, A.; Clark, J.; Goodwin, G.; Dodson, A.R.; Smith, P.H.; Fletcher, A.; Edwards, S.; Flohr, P.; Falconer, A.; Roe, T.; et al. Amplification and overexpression of E2F3 in human bladder cancer. Oncogene 2004, 23, 1627–1630. [Google Scholar] [CrossRef] [Green Version]
  94. Vimala, K.; Sundarraj, S.; Sujitha, M.V.; Kannan, S. Curtailing overexpression of E2F3 in breast cancer using siRNA (E2F3)-based gene silencing. Arch. Med. Res. 2012, 43, 415–422. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Measurement of hind leg circumference and carcass depth.
Figure 1. Measurement of hind leg circumference and carcass depth.
Animals 12 01103 g001
Figure 2. Manhattan plot of −log10 (p) for carcass depth. The horizontal line represents the threshold.
Figure 2. Manhattan plot of −log10 (p) for carcass depth. The horizontal line represents the threshold.
Animals 12 01103 g002
Figure 3. Manhattan plots of −log10 (p) for hind leg circumference in segments 1–3. The horizontal lines represent the thresholds.
Figure 3. Manhattan plots of −log10 (p) for hind leg circumference in segments 1–3. The horizontal lines represent the thresholds.
Animals 12 01103 g003
Table 1. Mean carcass depth and hind leg circumference.
Table 1. Mean carcass depth and hind leg circumference.
TraitMean (cm)Standard Deviation (cm)Maximum (cm)Minimum (cm)
Carcass depth
Circumference of hind leg
63.88
68.01
4.62
16.58
82
110
13
22
Table 2. SNPs associated with carcass depth, determined by GWAS analysis.
Table 2. SNPs associated with carcass depth, determined by GWAS analysis.
SNPChromosomePositionp-Value
BovineHD0200025561
BovineHD0200025562
BovineHD0200025553
BovineHD0200025550
BovineHD0500008673
BovineHD0500023139
BovineHD1000031388
BovineHD1700020542
BovineHD2600013096
2
2
2
2
5
5
10
17
26
89923045
89926675
89896059
89887085
29582709
81891942
44213755
70557785
46267236
1.49 × 10−14
1.49 × 10−14
3.4 × 10−13
8.3 × 10−13
1.3 × 10−13
2.17 × 10−8
1.17 × 10−8
3.47 × 10−8
5.58 × 10−8
Table 3. SNPs in segment 2 associated with hind leg circumference, as determined by NGS dataset analysis.
Table 3. SNPs in segment 2 associated with hind leg circumference, as determined by NGS dataset analysis.
SNPChromosomePositionp-Value
rs84436800
rs84450386
rs84441028
rs84429941
rs84428499
rs25168454
rs25439719
rs25457101
rs25237349
rs25224780
12
12
12
12
12
15
15
15
15
15
84436800
84450386
84441028
84429941
84428499
25168454
25439719
25457101
25237349
25224780
2.17 × 10−8
3.45 × 10−8
6.93 × 10−8
8.13 × 10−8
8.24 × 10−8
1.31 × 10−10
4.14 × 10−8
4.14 × 10−8
1.12 × 10−7
1.22 × 10−7
Table 4. SNPs in segment 3 associated with hind leg circumference, as determined by NGS dataset analysis.
Table 4. SNPs in segment 3 associated with hind leg circumference, as determined by NGS dataset analysis.
SNPChromosomePositionp-Value
rs37207517
rs12211246
rs12211363
rs12211411
rs34314488
rs34314780
rs34329034
rs34330397
rs34417763
rs34423556
rs34376791
rs34381008
rs34381454
rs34381516
rs34416372
23
26
26
26
29
29
29
29
29
29
29
29
29
29
29
37207517
12211246
12211363
12211411
34314488
34314780
34329034
34330397
34417763
34423556
34376791
34381008
34381454
34381516
34416372
8.61 × 10−8
1.54 × 10−7
1.54 × 10−7
1.54 × 10−7
6.32 × 10−8
6.32 × 10−8
6.32 × 10−8
6.32 × 10−8
9.3 × 10−8
9.3 × 10−8
1.27 × 10−7
1.27 × 10−7
1.27 × 10−7
1.27 × 10−7
1.5 × 10−7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bordbar, F.; Mohammadabadi, M.; Jensen, J.; Xu, L.; Li, J.; Zhang, L. Identification of Candidate Genes Regulating Carcass Depth and Hind Leg Circumference in Simmental Beef Cattle Using Illumina Bovine Beadchip and Next-Generation Sequencing Analyses. Animals 2022, 12, 1103. https://doi.org/10.3390/ani12091103

AMA Style

Bordbar F, Mohammadabadi M, Jensen J, Xu L, Li J, Zhang L. Identification of Candidate Genes Regulating Carcass Depth and Hind Leg Circumference in Simmental Beef Cattle Using Illumina Bovine Beadchip and Next-Generation Sequencing Analyses. Animals. 2022; 12(9):1103. https://doi.org/10.3390/ani12091103

Chicago/Turabian Style

Bordbar, Farhad, Mohammadreza Mohammadabadi, Just Jensen, Lingyang Xu, Junya Li, and Lupei Zhang. 2022. "Identification of Candidate Genes Regulating Carcass Depth and Hind Leg Circumference in Simmental Beef Cattle Using Illumina Bovine Beadchip and Next-Generation Sequencing Analyses" Animals 12, no. 9: 1103. https://doi.org/10.3390/ani12091103

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

Article Metrics

Back to TopTop