Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Genome-Wide Association Study for Carcass Traits in an Experimental Nelore Cattle Population

Abstract

The purpose of this study was to identify genomic regions associated with carcass traits in an experimental Nelore cattle population. The studied data set contained 2,306 ultrasound records for longissimus muscle area (LMA), 1,832 for backfat thickness (BF), and 1,830 for rump fat thickness (RF). A high-density SNP panel (BovineHD BeadChip assay 700k, Illumina Inc., San Diego, CA) was used for genotyping. After genomic data quality control, 437,197 SNPs from 761 animals were available, of which 721 had phenotypes for LMA, 669 for BF, and 718 for RF. The SNP solutions were estimated using a single-step genomic BLUP approach (ssGWAS), which calculated the variance for windows of 50 consecutive SNPs and the regions that accounted for more than 0.5% of the additive genetic variance were used to search for candidate genes. The results indicated that 12, 18, and 15 different windows were associated to LMA, BF, and RF, respectively. Confirming the polygenic nature of the studied traits, 43, 65, and 53 genes were found in those associated windows, respectively for LMA, BF, and RF. Among the candidate genes, some of them, which already had their functions associated with the expression of energy metabolism, were found associated with fat deposition in this study. In addition, ALKBH3 and HSD17B12 genes, which are related in fibroblast death and metabolism of steroids, were found associated with LMA. The results presented here should help to better understand the genetic and physiologic mechanism regulating the muscle tissue deposition and subcutaneous fat cover expression of Zebu animals. The identification of candidate genes should contribute for Zebu breeding programs in order to consider carcass traits as selection criteria in their genetic evaluation.

Introduction

Beef cattle production in tropical and subtropical regions is predominantly based on Bos indicus (Zebu) breeds and their crosses with Bos taurus. Even though Zebu animals have adaptive advantages to tropical conditions over Taurine breeds, these animals still have lower reproductive efficiency, lower beef tenderness, higher age at slaughtering, lower proportion of fat in the carcass [1,2]. The challenge for livestock production system in (sub) tropical countries is producing beef to satisfy consumer demands for quantity and quality and at the same time.

The longissimus muscle area (LMA) is an efficient indicator of the carcass yield, carcass muscularity [3], carcass weight, fat and muscle traits in steers [4]. The subcutaneous fat covering over the longissimus muscle acts minimizing evaporative weight loss of the carcass in the cooler besides been an efficient carcass-finishing indicative [5]. Furthermore, subcutaneous fat thickness is related to beef quality through protecting the muscle from cold-shortening which occur after the slaughter in the carcass cooling process. The conventional refrigeration of carcasses after slaughter may result in tougher meat, thus, an adequate quality carcass must have enough fat thickness to guarantee its preservation and desirable quality for consumption [6]. In addition, the beef cattle market specifications require a fat thickness higher than the usually produced by Zebu breeds. However, the low propensity to subcutaneous fat deposition is a problem in Zebu breeds, which have lower proportion of subcutaneous and intramuscular fat percentage than Bos taurus breeds [7, 8].

Favourable genetic correlations between subcutaneous fat and reproductive traits reported by Caetano et al. [9] indicated that high subcutaneous fat deposition could denote early finishing and result in sexually more precocious animals. On the other hand, undesirable genetic correlation between subcutaneous fat with weight gain and residual feed intake have been reported [9, 10] and caused some concerns in the authors.

The use of genomic information has been a standard procedure for genetic evaluation in animal and plant breeding programs in order to improve quantitative traits. This procedure has been especially helpful for traits that are hard or expensive to measure and consequently not routinely recorded. Genome-wide association studies (GWAS) aim to identify regions in the genome that are associated with phenotypes and have been applied in many traits including carcass traits [11]. A common method used for GWAS analyses is based on testing one marker at time as fixed effect. However, this approach assumes absence of population structure, which is very unlikely to happen when it is working on real data. An alternative to traditional GWAS consists to integrate all genotypic and phenotypic information available (from genotyped and ungenotyped animals) in one-step procedure (single-step GWAS) that allows the use of any model, and all relationships simultaneously [12].

Most of genomic studies for carcass traits have been applied on Taurine breeds in temperate regions [13]. Moreover, there is a need to study those traits in Zebu breeds under tropical conditions to elucidate the genetic architecture of carcass traits in these breeds. The purpose of this study was to identify associations between chromosomal regions with carcass traits such as longissimus muscle area (LMA), subcutaneous rump fat thickness (RF), and subcutaneous backfat thickness (BF) in an experimental Nelore cattle population.

Material and Methods

Data

The analysed Nelore cattle data set was provided by the APTA Beef Cattle Center—Institute of Animal Science (IZ), Sertãozinho, SP, Brazil. This herd has three experimental lines: a selection line (NeS) which has been selected for yearling weight since 1978; the traditional line (NeT) which has been submitted to the same selection criterion as NeS but, eventually, receives animals from other herds; and a control line (NeC) selected for average yearling weight (stabilizing selection). Ethics committee of the School of Agricultural and Veterinarian Sciences (FCAV), Sao Paulo State University (UNESP) and APTA Beef Cattle Center—Institute of Animal Science approved this study.

The data set contained pedigree information on 9,529 animals, of which 2,306 had ultrasound records for longissimus muscle area (LMA), 1,832 for backfat thickness (BF), and 1,830 rump fat thickness (RF). The ultrasound animals were born from 1996 to 2013. Table 1 shows the descriptive statistics for all the studied traits.

thumbnail
Table 1. Descriptive statistics for longissimus muscle area (LMA), backfat thickness (BF) and rump fat thicknes (RF) traits.

https://doi.org/10.1371/journal.pone.0169860.t001

The herd management consisted on keeping the animals on pasture until seven months of age, when they were weaned. After this period, males and females (females born in 2004, 2005 and from 2008 to 2011) were submitted to a performance test on feedlot.

Serial ultrasound data were collected in vivo at 12 and 18 months of age for males and females, respectively. The difference among the age of phenotypic ultrasound measurement was mainly due the differed after weaning management applied for each sex. The animals were scanned by the PIE MEDICAL—Aquila equipment with probe of 7 inches and 3,5MHz, and then, the images were analysed using the Echo Image Viewer 1.0 (Pie Medical Equipament B.V., 1996) software. The probe was perpendicularly positioned between the 12th and 13th ribs on the left side to obtain the LMA and BF phenotypes. To obtain the RF phenotype, the probe was positioned at intersection of Gluteus Medius and Biceps Femoris muscles, which is located between ilium and ischium bones.

Genotyping Procedure

DNA was extracted from blood samples which were collected in 9 ml K2EDTA Vacutainer blood collection tubes (BD Diagnostics) by jugular venepuncture. These samples were inverted to mix and prevent clotting, immediately placed in isothermic boxes, and transferred to the laboratory. During sampling animals were handled by qualified professionals who observed the procedure of well-being and safety of all animals.

A high-density SNP panel (BovineHD BeadChip assay 700k, Illumina Inc., San Diego, CA) was used for genotyping. SNP markers with minor allele frequency (MAF) and call rate less than 2% and 98%, respectively, were excluded. Also, samples with a call rate less than 90% were not considered in analyses. After genomic data quality control, 437,197 SNPs from 761 animals were available, of which 721 had phenotypes for LMA, 669 for BF, and 718 for RF trait (Table 1).

(Co) Variance Component Estimation

The (co)variance components and genetic parameters were estimated by Bayesian inference, considering a linear single-trait animal model. Direct additive genetic and residual effects were included as random effects. Analyses were performed using GIBBS2f90 [14, 15]. The genetic and residual variance were obtained as follows: (1) where σ2a and σ2e are total additive genetic and residual variances, respectively, a is the vector of direct additive genetic effects, e is a vector of residual effects, and H is a matrix that combines pedigree and genomic relationships, and its inverse consists on the integration of additive and genomic relationship matrices, A and G, respectively [15]: (2) Where A-1 is the inverse of relationship matrix based on pedigree information, G-1 is the inverse of genomic relationship matrix, which was constructed and scaled as described by VanRaden [16], and A-122 is the inverse of pedigree-based relationship matrix for genotyped animals. The H matrix was built scaling G based on A22 considering that the average of diagonal of G is equal to average of the diagonal of A22 and average of off diagonal G is equal to average off diagonal A22.

The general model can be represented as follows: (3) where Y is the vector of phenotypic observations, X is an incidence matrix of phenotypes and fixed effects, b is the vector of fixed effects, that included contemporary groups (sex and year of birth), and age of ultrasound measurement as a covariable (linear and quadratic effects), Z is an incidence matrix that relates animals to phenotypes. The principal components analyses made based on genomic relationship matrix (G) revealed population sub-structure (Fig 1). Thus, the linear effect of the two principle components was considered in the model as covariable to correct for sub-structure of population as suggested by Price et al. [17].

thumbnail
Fig 1. Distribution of animals provided by principal component.

https://doi.org/10.1371/journal.pone.0169860.g001

Assumptions were: (4) with Ʃ = var(a) = Hσ2a and R = Iσ2a in the single-trait model, where Hσ2a is the additive genetic variance and Iσ2e the residual variance, H is described above and I is the appropriate identity matrix. An inverted chi-square distribution was used for the prior values of the direct and residual genetic variances.

The posterior conditional distributions of β, a, and e effects were sampled from a multivariate normal distribution. The analysis consisted of a single chain of 300,000 cycles with a "burn-in" of 100,000 cycles, taking a sample every 10 iterations. Thus, 20,000 samples were used to obtain the parameters. Chain convergence was assessed by visual inspection. The posterior estimates were obtained retrospectively using the POSTGIBBSF90 program [14].

Single Step Genome Wide Association Study (ssGWAS)

The analyses were performed using the single-step GWAS methodology [12] considering the same linear animal model used to estimate the (co) variance components before mentioned. The animal effects were decomposed in genotyped (ag) and ungenotyped (an) animals as describe by Wang et al. [12], with the animal effect of genotyped animal: (5) where Z was a matrix that related genotypes of each locus and u was a vector of marker effects. The variance of animal effects was assumed as: (6) where D was a diagonal matrix of weights for variances of markers (D = I for GBLUP), was the genetic additive variance captured by each SNP marker when the weighted relationship matrix (G*) was built with no weight.

Thus, the SNP effects were obtained following equation, as described by Wang et al. [12]: (7) where λ was defined by VanRaden et al. [18] as a normalizing constant, as described below (8) M was the number of SNPs and pi was the frequency of the second allele in the i-th SNP.

The following iterative process described by Wang et al. [12] was used considering D to estimate the SNP effects: 1. D = I; 2. To calculate GEBVs for all animals in data set using ssGBLUP; 3. To calculate the SNP effect: ȗ = λDZ’G*-1ȃg; 4. To calculate the variance of each SNP: di = ûi22pi (1—pi), where i is the i-th marker; 5. To normalize the values of SNPs to keep constant the additive genetic variance; 6. To calculate the G matrix; 7. Exit, or loop to step 2.

The effects of markers were obtained by 2 iterations from step 2 to 6 as showed by Wang et al. [12]. The percentage of genetic variance explained by i-th region was calculated as described by Wang et al. [19]: (9) where ai is genetic value of the i-th region that consists of 50 consecutive SNPs, σ2a is the total genetic variance, Zj is vector of SNP content of the j-th SNP for all individuals, and ûj is marker effect of the j-th within the i-th region.

Analyses were performed using BLUPf90 family software [14] modified to include genomic information [15]. The results were presented by the proportion of variance explained by each window of 50 SNPs with average of 280 Kb.

Search for Associated Genes

The chromosome segments that explained more than 0.5% of additive genetic variance were selected to explore and determine possible quantitative trait loci (QTL). The Map Viewer tool of bovine genome was used for identification of genes, available at "National Center for Biotechnology Information" (NCBI - http://www.ncbi.nlm.nih.gov) in UMD3.1 bovine genome assembly and Ensembl Genome Browser (http://www.ensembl.org/index.html). The classification of genes according to its biological function, identification of metabolic pathways and enrichment of genes was performed using the "Database for Annotation, Visualization and Integrated Discovery (DAVID) v. 6.7” (http://david.abcc.ncifcrf.gov/), GeneCards (http://www.genecards.org/), and UniProt (http://uniprot.org).

Results

The posterior distribution for heritabilities estimated in this study are shown in Table 2. The known genes located in the associated windows are presented per each studied trait in Tables 35. A total of 12, 18 and 15 different windows were found associated to LMA, RF, and BF traits, respectively. In addition, many uncharacterized genes (LOC) were found in those regions associated to all studied traits. However, as mentioned by Fragomeni et al. [20], GWAS results should be carefully interpreted avoiding to determine an association as a causative effect since many QTL have been described for different traits but just a few of them have been validated by others studies. General information about all results of ssGWAS for all the studied traits are described in S1S3 Files.

thumbnail
Table 2. Variance components, heritability estimates and their confidence interval for longissimus muscle area (LMA), backfat thickness (BF), and rump fat thickness (RF) carcass traits.

https://doi.org/10.1371/journal.pone.0169860.t002

thumbnail
Table 3. Identification of genes based on additive genetic variance explained by windows of 50 adjacent SNPs for longissimus muscle area.

https://doi.org/10.1371/journal.pone.0169860.t003

thumbnail
Table 4. Identification of genes based on additive genetic variance explained by windows of 50 SNPs for backfat thickness.

https://doi.org/10.1371/journal.pone.0169860.t004

thumbnail
Table 5. Identification of genes based on additive genetic variance explained by windows of 50 SNPs for rump fat thickness.

https://doi.org/10.1371/journal.pone.0169860.t005

A total of 28 known genes and 15 uncharacterized genes were found associated to LMA (Fig 2 and Table 3). The first cluster from enrichment analysis for LMA was related to processes that stop, prevent or reduce the rate of cell death by apoptotic process (S1 Table). In addition, the second cluster was related to DNA repair, mechanisms that minimize acute damage to the cell's overall integrity and response to DNA damage stimulus. Both processes are dependent on each other, because failures in DNA repair will increase the rate of cell death.

thumbnail
Fig 2. Additive genetic variance explained by windows of 50 adjacent SNPs distributed by chromosomes for longissimus muscle area.

https://doi.org/10.1371/journal.pone.0169860.g002

A total of 42 known genes and 23 uncharacterized genes (located on BTA1, BTA2, BTA7, BTA9, BTA10, BTA11, BTA14, BTA16, BTA21, and BTA29) were found in those windows associated to BF (Fig 3 and Table 4). The clusters obtained by the genes enrichment analysis for BF were related to ATP-binding, transcription regulation, and protein amino acid phosphorylation (S2 Table). Many genes that are related to lipids and fat expression were found in the windows associated with BF.

thumbnail
Fig 3. Additive genetic variance explained by windows of 50 adjacent SNPs distributed by chromosomes for backfat thickness in Nelore.

https://doi.org/10.1371/journal.pone.0169860.g003

A total of 32 known genes and 19 uncharacterized genes (located in BTA2, BTA5, BTA6, BTA9, BTA13, BTA14, BTA15, BTA19, and BTA20) were found in the windows associated to RF (Fig 4). Some of those regions have been reported to be associated with lipid production (Table 5). These findings confirm the polygenic nature of fat deposition. The clusters obtained by the genes enrichment analysis of RF are related to ATP binding, response to hormone stimulus and intracellular signalling cascade (S3 Table).

thumbnail
Fig 4. Additive genetic variance explained by windows of 50 adjacent SNPs distributed by chromosomes for rump fat thickness.

https://doi.org/10.1371/journal.pone.0169860.g004

A window located on BTA14 (at 24.87–25.11 Mb) was associated to both fat deposition traits, BF and RF, where the LYN, RPS20, MOS, PLAG1, and CHCHD7 genes are located.

Discussion

The confidence interval for posterior distribution of heritability was very accurate for all studied traits; the range was always within 4 points from the mean. This narrow posterior distribution indicates a good adjustment of the H matrix to the structure of analysed population. Heritability values showed that a great part of phenotypic variance was due to additive genetic effects. Thus, it is important to identify the genes involved in these traits expression, especially major effects genes.

Several known genes that have their functions previously reported associated to the expression of energy metabolism were confirmed associated with fat deposition in this study. In addition, genes such as the ALKBH3 and HSD17B12, which are related in fibroblast death and metabolism of steroids, were found associated with longissimus muscle area.

The ALKBH3 gene (found in a window that most explained additive genetic variance of LMA located in BTA15) codifies an intrinsic DNA repair protein that suppresses transcription associated DNA damage at highly expressed genes [21]. Even though this gene has not been associated for those traits in a cattle population, Nay et al. [22] reported that deleting it in mice made fibroblasts more susceptible to death. Still in this same associated window, the HSD17B12 gene is located. This gene is a member of the hydroxysteroid dehydrogenase superfamily, involved in the metabolism of steroids, retinoids, bile and fatty acids that apparently can be related in metabolic pathways involved in tumor [23]. Considering that tumor results from uncontrolled cells multiplication, it could be related to DNA repair.

The HTR2B gene, located on BTA2 (Table 4), was found in a window associated to BF. This gene has been previously associated to fat deposition in humans [24] and it is known that it encodes one of the several different receptors for 5-hydroxytryptamine (serotonin). The serotonin hormone acts as a neurotransmitter, a hormone, and a mitogen. Two candidate genes were found in the window that explained the greatest part of genetic variance (BTA29): ALDH3B1 and CHKA genes. The ALDH3B1 gene had its function related to oxidation of lipid-derived aldehydes generated in the human plasma membrane [25] and is associated to diabetes in humans [26]. The biological process of the CHKA gene is related to lipid biosynthesis and it has been reported differentially regulated in mice by high fatty acids levels [27]. Even though some genes found in the associated windows have not been directly associated to the expression of these traits in previous researches, it is important to discuss a bit about their functions.

The XKR4 gene (placed in a window associated with BF located on BTA14) was previously reported by Porto Neto et al. [27] to be associated with RF in a recent genome-wide association study in Belmont Red and Santa Gertrudis breeds. The authors found association between RF and three SNPs within XKR4 gene (P < 0.001), one of the SNPs alone explained 5.9% of the additive genetic variance. This gene has also been reported as a candidate gene for residual feed intake, average daily feed intake and average daily gain [28]. In addition, Bastion et al. [29] suggested the XKR4 gene participates in the regulation of prolactin secretion in cattle.

A window located on BTA14 (at 24.87–25.11 Mb) was associated to both fat deposition traits, BF and RF (Table 4 and Table 5). In this associated window were found the PLAG1, CHCHD7, MOS, RPS20, LYN, and RDHE2 (also known as SDR16C5) genes. There are evidences that these genes are directly associated with the expression of many traits in beef cattle. Independent studies have reported association between PLAG1 gene with fat, residual feed intake, carcass, performance and reproductive traits in beef cattle breeds including Nelore, Japanese Black and Korean cattle [3035]. The PENK gene, also located in a window on BTA14 (at 25.20–25.49 Mb), was previously reported associated with residual feed intake in a swine population [36]. The windows that these two genes, PENK and PLAG1, are located (at 24.87–25. 10 Mb and 25.20–25.49 Mb of BTA15) explained a great part of additive genetic variance of both fat deposition traits, BF and RF. Thus, once they are juxtaposed and with possible physical disequilibrium, it is hard to say whether these genes have independent function or the associated windows where they are located are in linkage disequilibrium. Karim et al. [37] reported that causative mutation in the PLAG1-CHCHD7 intergenic region affected the bovine stature in a European breed. Nishimura et al. [30] re-exanimated this region in a Japanese Black cattle founding three QTL for carcass weight and strongly suggested that it was identical to the causative variation reported by Karim et al. [37]. There are evidences that these genes have influence on human height [38, 39].

Even for similar traits such as BF and RF, which are highly phenotypic (0.61) and genetic (0.79 ± 0.05) correlated [40] differing in the site of fat deposition, they cannot be considered the same trait. Thus, differences in the genomic regions that influence these traits were found here, which might be due epigenetic effect that may regulate them. Li et al. [41] concluded that methylation differences between adipose depots are dependent on their location. The genes that act depositing lipids in several regions are not always the same, the methylation depends on the region where these genes are located.

Genes related with fat deposition were found in windows that explained a great part additive genetic variance of RF trait. The CAPN5 and MYO7A (located at 57.29–57.43 Mb of BTA15) genes are associated with blood cholesterol level in humans and with abdominal fat in chicken [42, 43]. In addition, the LUM gene (located at 20.99–21.12 Mb of BTA5) was related with omental fat deposition in humans [44]. The BAI3 gene, placed in a region of BTA9 acts on the fusion of myoblast and it is expressed on extracellular matrix [45] that participates in the formation of adipose tissue.

Some others genes had their expression identified in skeletal muscle and performance capability on adipose tissue, since that tissue are juxtaposed and have many genes with common acting. According to Wang et al. [46], the RPS6KB1 gene (located at 10.84–11.30 Mb of BTA19) is sub expressed in skeletal muscle of calves which were experimental fed with high levels of protein. In addition, the DCN gene (located at 20.99–21.12 Mb of BTA5) operates in intramyocellular space co-acting with myostatin gene [47] and it is in extracellular matrix with highly potential to contribute for fat deposition [48].

Considering that longissimus muscle area, backfat thickness and rump fat thickness traits are governed by many variants with small effects, the results in the present study confirm the polygenic effect of these traits. On the other hand, it is important to highlight that while several DNA regions with larger effects have been detected, some of these may not be due to variants but could be mark chromosome segments from important ancestors or just be sampling noise. In this study, 437,197 SNPs effects were predicted from about 700 data points (DGVs). With so few points, there is a statistical possibility that a limited number of SNP could provide good prediction as shown by cross-validation just by chance. Stam [49] pointed out that the number of independent chromosome segments due to small effective population size is small. Therefore, it is possible that GWAS as performed here may have a limited resolution and could be subjected to high sampling noise. However, our results are very consistent and many genomic regions with known genes that have their functions already described were confirmed associate to the studied traits. Besides the known genes, many uncharacterized genes were found in the associated windows. In this manner, information about the associations found in this study might be used in future studies with the intention of characterization of those genes. Such studies would be used to confirm the association between genes in those chromosomic regions and carcass traits.

The results presented here should help to better understand the genetic and physiologic mechanism regulating the muscle tissue deposition and subcutaneous fat cover expression of Zebu animals. The analysed data set contained information from proved animals of an experimental herd, which have progenies participating in many breeding programs in several regions of Brazil. Despite the polygenic nature of the studied traits, some genes found in the associated windows, such as ALKBH3 and HSD17B12 are more likely to be related to longissimus muscle area, as well as the PLAG1, CAPN5, MYO7A and XKR4 genes, which are possibly associated with fat deposition in Nelore cattle. Thus, the identification of these genes should contribute for future studies to validate them to use as candidate genes for Zebu breeding programs.

Supporting Information

S1 File. Results of ssGWAS for longissimus muscle area.

https://doi.org/10.1371/journal.pone.0169860.s001

(XLSX)

S2 File. Results of ssGWAS for back fat thickness.

https://doi.org/10.1371/journal.pone.0169860.s002

(XLSX)

S3 File. Results of ssGWAS for rump fat thickness.

https://doi.org/10.1371/journal.pone.0169860.s003

(XLSX)

S1 Table. Gene enrichment clustering for longissimus muscle area.

https://doi.org/10.1371/journal.pone.0169860.s004

(DOCX)

S2 Table. Gene enrichment clustering for backfat thickness.

https://doi.org/10.1371/journal.pone.0169860.s005

(DOC)

S3 Table. Gene enrichment clustering for rump fat thickness.

https://doi.org/10.1371/journal.pone.0169860.s006

(DOCX)

Acknowledgments

The costs associated to genotyping of animals was supported by the thematic project "Genomic tools for the genetic improvement of economically important traits directly in Nelore cattle" (FAPESP grant numbers 2009/16118-5 and 2013/01228-5).

The authors are thankful to the Center APTA Beef Cattle—Institute of Animal Science (IZ), Sertãozinho, SP, Brazil for providing the dataset used in this study.

Author Contributions

  1. Conceptualization: LGA RMOS.
  2. Data curation: MEZM RMOS.
  3. Formal analysis: RMOS BOF FB GMFC AAB DALL.
  4. Funding acquisition: LGA MEZM.
  5. Investigation: MEZM LGA TMC JNSGC.
  6. Methodology: RMOS FB LGA IM DALL BOF.
  7. Project administration: LGA MEZM.
  8. Resources: LGA.
  9. Software: IM.
  10. Supervision: LGA MEZM.
  11. Visualization: RMOS NBS.
  12. Writing – original draft: RMOS.
  13. Writing – review & editing: NBS FB IM DALL AAB.

References

  1. 1. O’Connor SF, Tatum JD, Wulf DM, Green RD, Smith GC. Genetic effects on beef tenderness in Bos indicus composite and Bos taurus cattle. J Anim Sci. 1997;75(7): 1822–1830. pmid:9222838
  2. 2. Alves DD, Mancio AB. Maciez da carne bovina–Uma revisão. Rev FZVA. 2007;14(1): 193–216.
  3. 3. Boggs DL, Merkel AR. Live animal carcass evaluation and selection manual. 3th ed. Kendall/Hunt Publishing Co., Dubuque, Iowa. 1990.
  4. 4. Bergen R, Miller SP, Wilton JW. Genetic correlations among indicator traits for carcass composition measured in yearling beef bulls and finished feedlot steers. Can. J. Anim. Sci.. 2005;85: 463–473.
  5. 5. Hedrick HB. Methods of estimating live animal and carcass composition. J Anim Sci.. 1983;57: 1316–1326.
  6. 6. Malheiros JM, Baldassini WA, Dias VAD, Silva JAIIV, Curi RA, Chardulo LAL. Chemical and sensory meat characteristics of Nellore cattle (Bos indicus) finished with different levels of backfat thickness in the longissimus thoracis muscle. B. Indústr. Anim. 2015;72(4): 341–348.
  7. 7. Crouse JD, Cundiff LV, Koch RM, Koohmaraie M, Seideman SC. Comparisons of Bos indicus and Bos taurus inheritance for carcass beef characteristics and meat palatability J. Ani. Sci. 1989;67: 2661–2668.
  8. 8. Moreira FB, Souza NE, Matsushita M, Prado IN, Nascimento WG. Evaluation of carcass characteristics and meat chemical composition of Bos indicus and Bos indicus x Bos taurus crossbred steers finished in pasture systems. Braz. Arch. Biol. Technol. 2003;46(4): 609–616.
  9. 9. Caetano SL, Savegnago RP, Boligon AA, Ramos SB, Chud TCS, Lôbo RB, et al. Estimates of genetic parameters for carcass, growth and reproductive traits in Nelore cattle. Livest Sci.. 2013;155: 1–7.
  10. 10. Santana MHA, Rossi P Jr, Almeida R, Cucco DC. Feed efficiency and its correlations with carcass traits measured by ultrasound in Nellore bulls. Livest Sci. 2012;145: 252–257.
  11. 11. Lu D, Sargolzaei M, Kelly M, Voort GV, Wang Z, Mandell I, et al. Genome-wide association analyses for carcass quality in crossbred beef cattle. BMC Genet. 2013;14: 80. pmid:24024930
  12. 12. Wang H, Misztal I, Aguilar I, Legarra A, Muir WM. Genome-wide association mapping including phenotypes from relatives without genotypes. Genet Res (Camb). 2012;94(2): 73–83.
  13. 13. Pryce JE, Wales WJ, Haas Y, Veerkamp RF, Hayes BJ. Genomic selection for feed efficiency in dairy cattle. Animal. 2014;8(1): 1–10. pmid:24128704
  14. 14. Misztal I, Tsuruta S, Strabel T, Auvray B, Druet T, Lee DH. BLUPF90 and related programs (BGF90). Proc. 7th World Congr. Genet. Appl. Livest. Prod., Montpellier, France. Communication No. 28. 2002.
  15. 15. Aguilar I, Misztal I, Johnson DL, Legarra A, Tsuruta S, Lawlor TJ. Hot topic: A unified approach to utilize phenotypic, full pedigree, and genomic information for genetic evaluation of Holstein final score. J Dairy Sci. 2010;93: 743–752. pmid:20105546
  16. 16. VanRaden PM. Efficient methods to compute genomic predictions. J Dairy Sci. 2008;91: 4414–4423. pmid:18946147
  17. 17. Price AL, Patterson NJ, Plenge RM, Weinblatt ME, Shadick NA, Reich D. Principal component analysis corrects for stratification in genome-wide association studies. Nat Genet. 2006;38(8): 904–909 pmid:16862161
  18. 18. VanRaden PM, Van Tassell CP, Wiggans GR, Sonstegard TS, Schnabel RD, Taylor JF, et al. Invited review: reliability of genomic predictions for North American Holstein bulls. J Dairy Sci. 2009;92: 16–24. pmid:19109259
  19. 19. Wang H, Misztal I, Aguilar I, Legarra A, Fernando RL, Vitezica Z, et al. Genome-wide association mapping including phenotypes from relatives without genotypes in a single-step (ssGWAS) for 6-week body weight in broiler chickens. Front Genet. 2014; 20:5:134.
  20. 20. Fragomeni BO, Misztal I, Lourenco DL, Aguilar I, Okimoto R, Muir WM. Changes in variance explained by top SNP windows over generations for three traits in broiler chicken. Front Genet. 2014;5: 332. pmid:25324857
  21. 21. Liefke R, Windhof-Jaidhauser IM, Gaedcke J, Salinas-Riester G, Wu F, Ghadimi M, et al. The oxidative demethylase ALKBH3 marks hyperactive gene promoters in human cancer cells. Genome Med. 2015;7(1): 66. pmid:26221185
  22. 22. Nay SL, Lee DH, Bates SE, O’Connor TR. Alkbh2 protects against lethality and mutation in primary mouse embryonic fibroblasts. DNA Repair (Amst). 2012;11(5): 502–510.
  23. 23. Visus C, Ito D, Dhir R, Szczepanski MJ, Chang YJ, Latimer JJ, et al. Identification of Hydroxysteroid (17β) dehydrogenase type 12 (HSD17B12) as a CD8+ T-cell-defined human tumor antigen of human carcinomas. Cancer Immunol. Immunother. 2011;60: 919–929. pmid:21409596
  24. 24. Sohle J, Machuy N, Smailbegovic E, Holtzmann U, Gronniger E, Wenck H, et al. Identification of New Genes Involved in Human Adipogenesis and Fat Storage. PLoS One. 2012;7: e31193. pmid:22384002
  25. 25. Kitamura T, Naganuma T, Abe K, Nakahara K, Ohno Y, Kihara A. Substrate specificity, plasma membrane localization, and lipid modification of the aldehyde dehydrogenase ALDH3B1. Biochim Biophys Acta. 2013;1831(8): 1395–1401. pmid:23721920
  26. 26. Jeff JM, Armstrong LL, Ritchie MD, Denny JC, Kho AN, Basford MA, et al. ADMIXTURE Mapping and subsequent fine-mapping suggests a biologically relevant and novel association on chromosome 11 for type 2 diabetes in African Americans. PLoS One. 2014;9: e86931. pmid:24595071
  27. 27. Porto Neto LR, Bunch RJ, Harrison BE, Barendse W. Variation in the XKR4 gene was significantly associated with subcutaneous rump fat thickness in indicine and composite cattle. Anim Genet. 2012;43(6): 785–789. pmid:22497494
  28. 28. Lindholm-Perry AK, Kuehn LA, Smith TP, Ferrell CL, Jenkins TG, Freetly HC, et al. A region on BTA14 that includes the positional candidate genes LYPLA1, XKR4 and TMEM68 is associated with feed intake and growth phenotypes in cattle. Anim Genet. 2012;43(2): 216–219. pmid:22404358
  29. 29. Bastin BC, Houser A, Bagley CP, Ely KM, Payton RR, Saxton AM, et al. A polymorphism in XKR4 is significantly associated with serum prolactin concentrations in beef cows grazing tall fescue. Anim Genet. 2014;45(3): 439–441. pmid:24666329
  30. 30. Nishimura S, Watanabe T, Mizoshita N, Tatsuda K, Fujita T, Watanabe N, et al. Genome-wide association study identified three major QTL for carcass weight including the PLAG1-CHCHD7 QTN for stature in Japanese Black cattle. BMC Genet. 2012;13: 40. pmid:22607022
  31. 31. Fortes MRS, Reverter A, Kelly M, Mcculloch R, Lehnert SA. Genome-wide association study for inhibin, luteinizing hormone, insulin-like growth factor 1, testicular size and semen traits in bovine species. Andrology. 2013;1(4): 644–650. pmid:23785023
  32. 32. Magalhães AFB, Carmargo GMF, Fernandes Junior GA, Gordo DGM, Tonussi RL, Costa RB, et al. Genome-wide association study of meat quality traits in Nellore cattle. PLoS One. 2016;11: e0157845. pmid:27359122
  33. 33. Sharma A, Dang CG, Kim KS, Kim JJ, Lee HK, Kim HC, et al. Validation of genetic polymorphisms on BTA14 associated with carcass trait in a commercial Hanwoo. Anim Genet. 2014;45(6): 863–867. pmid:25164077
  34. 34. Lee SH, Choi BH, Lim D, Gondro C, Cho YM, Dang CG, et al. Genome-Wide Association Study Identifies Major Loci for Carcass Weight on BTA14 in Hanwoo (Korean Cattle). PLoS One. 2013;8(10): e74677. pmid:24116007
  35. 35. Utsunomiya YT, Do Carmo AS, Carvalheiro R, Neves HH, Matos MC, Zavarez LB, et al. Genome-wide association study for birth weight in Nellore cattle points to previously described orthologous genes affecting human and bovine height. BMC Genet. 2013;14: 52. pmid:23758625
  36. 36. Jiao S, Maltecca C, Gray KA, Cassady JP. Feed intake, average daily gain, feed efficiency, and real-time ultrasound traits in Duroc pigs: II. Genome-wide association. J Anim Sci. 2014;92(2): 2846–2860
  37. 37. Karim L, Takeda H, Lin L, Druet T, Arias JA, Baurain D, et al. Variants modulating the expression of a chromosome domain encompassing PLAG1 influence bovine stature. Nat Genet. 2011;43(5): 405–413 pmid:21516082
  38. 38. Gudbjartsson DF, Walters GB, Thorleifsson G, Stefansson H, Halldorsson BV, Zusmanovich P, et al. Many sequence variants affecting diversity of adult human height. Nat Genet. 2008;40(5): 609–615. pmid:18391951
  39. 39. Lettre G, Jackson AU, Gieger C, Schumacher FR, Berndt SI, Sanna S, et al. Identification of ten loci associated with height highlights new biological pathways in human growth. Nat Genet. 2008;40(5): 584–591. pmid:18391950
  40. 40. Ceacero TM, Mercadante MEZ, Cyrillo JNdSG, Canesin RC, Bonilha SFM, de Albuquerque LG. (2016) Phenotypic and genetic correlations of feed efficiency traits with growth and carcass traits in Nellore cattle selected for postweaning weight. PLoS One. 2016;11(8).
  41. 41. Li M, Wu H, Wang T, Xia Y, Jin L, Jiang A, et al. Co-methylated genes in different adipose depots of pig are associated with metabolic, inflammatory and immune processes. Int J Biol Sci. 2012;8(6): 831–837. pmid:22719223
  42. 42. Sáez ME, Martínez-Larrad MT, Ramírez-Lorca R, González-Sánchez JL, Zabena C, Martinez-Calatrava MJ, et al. CALPAIN-5 gene variants are associated with diastolic blood pressure and cholesterol levels. BMC Med Genet. 2007;8: 1. pmid:17227582
  43. 43. Zhang H, Wang SZ, Wang ZP, Da Y, Wang N, Hu XX, et al. A genome-wide scan of selective sweeps in two broiler chicken lines divergently selected for abdominal fat content. BMC Genomics. 2012;13: 704. pmid:23241142
  44. 44. Oliva K, Barker G, Rice GE, Bailey MJ, Lappas M. 2D-DIGE to identify proteins associated with gestational diabetes in omental adipose tissue. J Endocrinol. 2013;218(2): 165–178. pmid:23709000
  45. 45. Hamound N, Trana V, Croteaua L, Kaniaa A, Côté J. G-protein coupled receptor BAI3 promotes myoblast fusion in vertebrates. Proc Natl Acad Sci U S A. 2014;111(10): 3745–3750. pmid:24567399
  46. 46. Wang P, Drackley JK, Stamey-Lanier JA, Keisler D, Loor JJ. Effects of level of nutrient intake and age on mammalian target of rapamycin, insulin, and insulin-like growth factor-1 gene network expression in skeletal muscle of young Holstein calves. J Dairy Sci. 2014;97: 383–391. pmid:24210480
  47. 47. Albrecht E, Liu X, Yang X, Zhao R, Jonas L, Maak S. Colocalization of myostatin and decorin in bovine skeletal muscle. Archiv Tierzucht. 2011; 54: 147–156.
  48. 48. Douglas T, Heinemann S, Bierbaum S, Scharnweber D, Worch H. Fibrillogenesis of collagen types I, II, and III with small leucine-rich proteoglycans decorin and biglycan. Biomacromolecules. 2006;7(8): 2388–2393 pmid:16903686
  49. 49. Stam P. The distribution of the fraction of the genome identical by descent in finite random mating populations. Genet Res. 1980; 35: 131–155.