Next Article in Journal
Welfare Implications for Tigers in Travelling Circuses
Previous Article in Journal
Epidemiological Insights into Feline Leukemia Virus Infections in an Urban Cat (Felis catus) Population from Brazil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing Genomic Prediction Accuracy for Body Conformation Traits in Korean Holstein Cattle

1
Department of Animal Science and Technology, College of Biotechnology and Natural Resources, Chung-Ang University, Anseong 17546, Republic of Korea
2
Korea Animal Improvement Association, Seoul 06668, Republic of Korea
3
Dairy Cattle Improvement Center of NH-Agree Business Group, National Agricultural Cooperative Federation, Goyang 10292, Republic of Korea
4
Animal Breeding and Genetics Division, National Institute of Animal Science, Rural Development Administration, Cheonan 31000, Republic of Korea
5
Department of Animal Biotechnology, Jeonbuk National University, Jeonju 54896, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 31 December 2023 / Revised: 18 March 2024 / Accepted: 28 March 2024 / Published: 29 March 2024
(This article belongs to the Section Animal Genetics and Genomics)

Abstract

:

Simple Summary

The Holstein breed is crucial in the dairy industry in Korea. In this study, we focused on the accuracy of genomic predictions for 12 of 24 significant body conformation traits examined in Korean Holstein cattle. Employing various statistical methods and levels of genetic data, we assessed how accurately these traits could be predicted. The prediction accuracy increased notably when both offspring and parental genetic information were considered. We identified 18 key genetic regions, offering valuable insights for future research to identify specific genes related to these traits. This study highlights the potential of advanced genetic tools to improve breeding strategies for Korean Holstein cattle, in particular for enhancing traits with significant economic and health impacts.

Abstract

The Holstein breed is the mainstay of dairy production in Korea. In this study, we evaluated the genomic prediction accuracy for body conformation traits in Korean Holstein cattle, using a range of π levels (0.75, 0.90, 0.99, and 0.995) in Bayesian methods (BayesB and BayesC). Focusing on 24 traits, we analyzed the impact of different π levels on prediction accuracy. We observed a general increase in accuracy at higher levels for specific traits, with variations depending on the Bayesian method applied. Notably, the highest accuracy was achieved for rear teat angle when using deregressed estimated breeding values including parent average as a response variable. We further demonstrated that incorporating parent average into deregressed estimated breeding values enhances genomic prediction accuracy, showcasing the effectiveness of the model in integrating both offspring and parental genetic information. Additionally, we identified 18 significant window regions through genome-wide association studies, which are crucial for future fine mapping and discovery of causal mutations. These findings provide valuable insights into the efficiency of genomic selection for body conformation traits in Korean Holstein cattle and highlight the potential for advancements in the prediction accuracy using larger datasets and more sophisticated genomic models.

1. Introduction

In Korea, the Holstein breed is the mainstay of milk and dairy production. The Korean Animal Improvement Association (KAIA) has developed advanced herd management strategies to maximize production efficiency and profitability, with a parallel focus on reducing methane emissions. This includes standardizing appearance and linear assessment to identify superior dairy cattle. Since the 1990s, assessing conformation traits, which are linked to body stature, longevity, and productivity, has been crucial in the global dairy cattle industry [1]. These traits not only influence cow productivity but also its value as a show animal, underscoring a comprehensive breeding strategy that emphasizes health, fertility, and performance within high-yield dairy systems.
Since its introduction by Meuwissen [2], genomic selection (GS), using thousands of single-nucleotide polymorphisms (SNPs), has revolutionized cattle breeding, enhancing the accuracy of estimated breeding value (EBV) in the dairy and beef cattle industries. This improvement has been attributed to the development of various GS models, validation methods, and SNP panels, all of which are aimed at refining genomic prediction accuracy. The effectiveness of GS is predominantly determined by the accuracy of the molecular breeding value (MBV), which is influenced by factors such as the size of the training population, marker density [3,4,5,6], clustering methods for cross-validation, and the relationship between reference and validation populations [7,8,9]. The selection of response variables, like EBV (estimated breeding value) or DEBV (deregressed-EBV), and the incorporation of a weighting factor or a residual polygenic component into the model [10,11] also play pivotal roles. Furthermore, the success of GS is contingent on the genetic architecture [12,13,14] and population structure [4,15].
To enhance the profitability of dairy cows, maximization of milk production, maintenance of good health status, and excellent reproductive capabilities are essential. Body conformation traits are associated with health and fertility, and improvements in these traits can reduce culling rates and facilitate easier calving and insemination, thereby directly impacting subsequent milk production [16]. It has been reported that cows requiring assistance during parturition produce 703 kg less milk, 24 kg less milk fat, and 21 kg less milk protein over a 305-day lactation period compared to cows that calve without assistance [17]. Body conformation traits in dairy cows are economically significant and correlate with calving ease, longevity, locomotion, and milk production traits. The pleiotropic associations between calving performance and rump traits support the known genetic correlation between characteristics of the pelvic area (e.g., rump traits) and calving ease and calf survival. A cow with a wide pin, long sloping rump, and slight slope from the pin bone to thurl is known to have easier calving [18,19]. Additionally, correlations have been observed between body conformation traits such as body condition score (BCS), stature (STA), udder depth (UDE), chest width (CWI), and milk production traits [20].
In this research, we focused on significantly enhancing the accuracy of genetic predictions for key conformation traits in modern Korean Holstein cattle. Our approach involved conducting a comprehensive genome-wide association study (GWAS) on 24 critical traits. Using five SNP panels, our objective was to pinpoint SNP markers, genes, and chromosomal regions linked to these traits across 29 autosomes, thereby enriching our understanding of their genetic underpinnings in Korean Holstein cattle. This study’s findings are crucial in assessing the effectiveness of genomic selection for body conformation traits in Korean Holsteins. Moreover, our results underscore the potential benefits of leveraging larger datasets and more advanced genomic models to markedly improve the precision of breeding strategies, particularly for traits that have significant economic and health relevance.

2. Materials and Methods

2.1. Phenotype and Genotype Data

We used phenotypic data for the following 24 traits: stature (STA), chest width (CWI), body depth (BDE), angularity (ANG), rump angle (RAN), rump width (RWI), rear leg set (RLS), foot angle (FAN), fore udder attachment (FUA), rear udder height (RUH), rear udder width (RUW), udder support (USU), udder depth (UDE), front teat placement (FTP), front teat length (FTL), overall conformation score (OCS), loin strength (LST), udder texture (UTX), rear teat placement (RTP), heel depth/foot height (HDE), bone quality (BQL), rear leg rear view (RLR), height at the front end (HHE), locomotion (LOC), and body condition score (BCS). Conformation scores were recorded according to the guidelines provided by KAIA. The estimated heritability of the 24 body conformation traits ranged from 0.031 to 0.334 (Table 1). Within the body traits, STA exhibited moderate heritability over 0.30, as did RAN within rump traits and UDE within udder traits.
Genomic data for a total of 11,660 cattle collected from three domestic institutions were compiled using the Illumina and Axiom platforms. This dataset serves as a foundation for analyzing genetic diversity and traits, utilizing the Illumina50Kv2 and Illumina50Kv3 (Illumina, Inc., San Diego, CA, USA), Axiom-BovMDv3, AxiomCustom300K platforms and KHOLs60Kv1 (Affymetrix, Inc., Santa Clara, CA, USA). To ensure the accuracy and reliability of the data, multiple stages of rigorous quality control procedures were applied to SNP markers and animal data.
During the quality control process, unmapped SNPs and SNPs located on sex chromosomes were initially excluded. Subsequently, SNPs with a call rate below 0.95, a minor allele frequency (MAF) below 0.01, and a Hardy–Weinberg equilibrium (HWE) p-value below 0.0001 were removed to enhance the quality of the data. Based on these criteria, the number of available SNP markers for each panel (KHOLs60Kv1, Illumina50Kv2 and v3, AxiomBovMDv3, AxiomCustom300K) was determined.
In the next step, genotypic data from various panels were imputed to the AxiomCustom300K level for integration, using the UMD 3.1 mapping information as a reference. This process is essential for maintaining consistency in genotypic data and increasing the accuracy of analyses. The SNP genotyping data for each panel underwent a separate imputation process using FImpute version 2.2 software [21], ensuring that all genetic information was accurately imputed back into their respective panels.
After the quality control and data imputation processes, the final number of animals analyzed per panel and the number of SNPs were determined. These genomic data form a crucial basis for analyzing genetic characteristics and diversity. The insights gained from these processes can be utilized in various research and applications, including genetic improvement, disease resistance, and productivity enhancement in cattle.

2.2. Model Definition

2.2.1. Deregressed Estimated Breeding Values (DEBVs) of the Response Variable in Genomic Analysis

Data were obtained from the national genetic evaluation of dairy cattle conducted by the National Institute of Animal Science (NIAS) of the Rural Development Administration (RDA), including variance components and estimated breeding values (EBVs). The variance components considering lactation group and HYS, provided by the National Institute of Animal Science in Korea, were applied to 25 linear-type traits. DEBVs were re-estimated using the estimated individual genetic values, reliabilities, and genetic variance components. DEBVs were re-estimated with the aim of using them as response variables for GS. Because the newly estimated response variables had varying reliability for each individual, a weighting factor was applied to account for this heterogeneous variance using Equation (1) [11]:
w i = ( 1 h 2 ) [ c + ( 1 r i 2 ) / r i 2 ] h 2
where r i 2 = reliability of DEBV response variables including and excluding the parent average (DEBVincPA and DEBVexcPA), h 2 = heritability of each trait, and c = proportion of genetic variation that cannot be explained by the markers. In this study, c was assumed to be 0.4 [22]. The resulting dataset, consisting of 11,095 reference animals with both genomic and phenotypic (DEBVs) information, was used for a GWAS and a genomic prediction model analysis using Bayesian methods.

2.2.2. Statistical Method for GS Model

In our research, the estimation of SNP marker effects for GWAS and genomic prediction models was conducted using an adapted Bayesian approach [2]. This method involved the application of modified BayesB [2] and BayesC [23] models, which were implemented through the specialized GenSel4R.8R software framework [24]. The approach was distinctive in that it employed a set of varied π values (0.75, 0.90, 0.99, and 0.995), each accompanied by specific weighting factors. Our version of the BayesB model was based on a mixed model concept, hypothesizing that a certain proportion π of SNP markers exhibit null effects, while the remainder display non-zero impacts. This method diverged from standard practices by utilizing a t-distribution as the foundational assumption for SNP marker effects and by factoring in locus-specific variances. The BayesC model is similar to the BayesB model, but it differs in that it assumes all markers contribute to genetic variation, with the effects of all markers sharing a common variance. BayesC additionally models the probability that the variance of marker effects is zero. For each individual trait under examination, we rigorously applied this tailored model to accurately estimate the effects of SNP markers, aligning with our modified equation framework.
y i = μ + j = 1 k Z i j u j δ j + e i
where y i = response variables (DEBVexcPA and DEBVincPA) for each trait, μ = overall mean, k = number of markers, Z i j = allelic state at the jth locus (−10, 0, and 10 in GenSel4R) in the ith individual, u j = random substitution effect for the jth marker, which follows a mixture distribution for this random substitution effect according to the indicator variable ( δ j ), a random 0/1 variable indicating the absence or presence of marker j in the model, with u j assumed to be normally distributed N ( 0 , σ u 2 ) when δ j = 1 , and otherwise, u j is assumed to be 0; and e i is a random residual effect assumed to be normally distributed N ( 0 , σ e 2 ).
To ascertain the posterior distributions of these parameters and their effects, we utilized Gibbs sampling in conjunction with a Markov Chain–Monte Carlo (MCMC) process encompassing 110,000 iterations in total. The initial 10,000 iterations were eliminated as part of the burn-in phase to ensure model stabilization. Post this phase, we meticulously gathered 10,000 samples, strategically selecting every tenth iteration to minimize the potential for autocorrelation, a technique referred to as ‘thinning’. These carefully curated samples were pivotal in calculating the posterior means for the effects and variances of the SNP markers.

2.2.3. Genomic Prediction Accuracy under Fivefold Cross-Validation

In this study, we used the K-means clustering method with 5-fold cross-validation to estimate genomic accuracy. K-means clustering divided the reference population into training and validation groups while minimizing the genetic relatedness between these two groups. For the K-means clustering method, we used pedigree data from 54,346 individuals related to the 11,095 individuals used in the genomic analysis to estimate a numerical relationship matrix (NRM). Estimates an NRM and transforms it into a distance matrix to effectively minimize the genetic relationships between groups. The clustering indicated that the statistics amax and aij within the groups were relatively distant from the statistics amax and aij between the groups (Table 2). However, notably, K-means clustering did not achieve perfect separation in one clustering group, possibly because of the limited connectivity between domestic cattle and imported individuals in the pedigree analysis.

3. Results and Discussion

3.1. GWAS of Each Trait

The significance of body conformation traits in dairy cows lies in their association with reproductive and milk production traits. In Chinese Holsteins, STA, BDE, and BQL exhibit strong positive genetic correlations with milk production traits, indicating that cows with greater STA, deeper BDE, and superior BQL tend to produce more milk enriched with higher fat and protein content and possess better somatic cell scores [25]. ANG showed a very high negative genetic correlation with BCS at −0.612, and milk yield and BCS exhibited a correlation of −0.386, suggesting that higher milk-producing cows tend to be leaner [26]. STA demonstrated significant genetic correlations with milk yield (0.34), milk fat percentage (0.30), milk protein percentage (0.32), somatic cell score (0.23), and 305-day milk yield (0.22). RAN was genetically correlated with calving interval at −0.16 [27]. Extensive research on the genetic correlation between body conformation traits and somatic cell scores revealed the strongest associations with FUA, FTP, and UDE [28].
In the GWAS results, we present regions that exhibit a proportion of genetic variance (GV) explained by 1.0 Mb of 1.0% or higher, along with informative SNPs and gene annotation outcomes (Table 3, Figure S1). Assuming an infinitesimal model, the expected proportion of genomic variance explained by each window was 0.04% ([1/2524] × 100). However, for cases with 1.0%, the association is 25 times higher, leading us to set the significance level at GV 1.0%. When a marker accounts for more than 1.0% of the genetic variation, it signals a strong genetic influence on the trait or disease in question. This level of variation is vital for achieving statistically significant results in GWAS, which requires analyzing large datasets. We used the BayesB method with a large value of π with DEBVincPA acting as a response variable for each trait. We opted for the BayesB method due to its superior handling of sparse genetic data and its ability to provide more accurate and flexible analysis of genetic contributions to traits, enhancing our research’s precision and insight. Among the 24 traits examined, 12 exhibited a GV of 1.0% or higher. Therefore, in the main text, we focused solely on traits with a GV exceeding 1.0%, while the remaining traits and research findings are provided in the supplementary materials (Table S1).
For ANG, we identified one informative SNP in a specific region, and gene annotation revealed presence of the GC gene. GC, responsible for encoding vitamin D-binding protein (DBP), has consistently emerged as the leading candidate gene for clinical mastitis (CM) resistance QTL [29,30]. Furthermore, the application of vitamin D as a therapeutic measure in lactating cows fed CM has been shown to reduce inflammation, implying a potential association between vitamin D and inflammation [31,32,33]. For the BCS trait, we identified eight informative SNPs within a specific region, and annotation revealed the presence of ANKAR, MSTN, ASNSD1, and OSGEPL1. Among these, the MSTN, OSGEPL1, ANKAR, and ASNSD1 genes have been reported to show a strong association with direct calving difficulty in cattle [34] and have been identified as candidate genes for milk yield and composition in dual-purpose Belgian Blue cows [35]. Additionally, the ANKAR, OSGEPL1, and ASNSD1 genes are adjacent to markers showing significant effects on the lactose percentage trait [36]. The MSTN gene is known to regulate the proliferation and differentiation of skeletal muscle cells, and various gene mutations are associated with skeletal muscle mass in several mammalian species [37,38]. Additionally, muscle weakness has been observed in mice with a knockout of the ASNSD1 gene [39]. For BQL, we identified seven informative SNPs within three distinct regions. Gene annotation of these regions revealed the presence of GC, ENSBTAG00000046149, GDPD5, ENSBTAG00000000628, MEIS2, and TMCO5A. Notably, the silencing of GDPD5 has been shown to decrease cell proliferation, migration, and invasion in breast cancer [40]. Additionally, it is considered one of the hub genes influencing milk yield traits in buffaloes [41]. Furthermore, MEIS2 was associated with carcass traits in Chinese Simmental beef cattle [42]. For CWI, we identified 10 informative SNPs within three distinct regions. Gene annotation of these regions revealed the presence of DECR1, CCT8L2, XRCC2, KMT2C, GALNTL5, and CTNNA2. However, no study has associated these genes with conformation traits. Nevertheless, CTNNA2 has been reported as a candidate gene for milking speed in North American Holstein cattle [43]. For the FTL trait, we identified one informative SNP within a specific region. Gene annotation of this region revealed the presence of the FAM49A gene. The FAM49A gene is associated with milk production in German Holstein cattle [44] and also had been reported as one of the candidate genes for milking speed in French Holstein cattle [45]. This gene was additionally associated with rear udder height in Holstein cattle [46]. For FTP, we identified 12 informative SNPs within two distinct regions. Gene annotation of these regions revealed the presence of SHOX2, ENSBTAG00000047546, and ZNF608. The SHOX2 gene is strongly associated with milk traits and growth and development in dairy goats [47] and is reported to be involved in height and chondrogenesis [48,49]. For HHE, we identified two informative SNPs within a specific region. Gene annotation of this region revealed the presence of ANKH and TRIO. The ANKH gene plays a crucial role in familial and severe forms of sporadic chondrocalcinosis [50]. Furthermore, ANKH encodes an inorganic pyrophosphate transport regulator that prevents deposition of calcium and phosphate in bones [51]. For RAN, we identified one informative SNP within one region. The FCHSD1 gene has been identified in this region; however, no study has associated it with body conformation traits. For RTP, we identified five informative SNPs within two regions. The annotation results for these regions revealed the presence of DOCK2 and CA10. The DOCK2 gene is associated with immune processes, including lymphocyte trafficking [52]. For RUW, we identified three informative SNPs within one region. The annotation results for this region revealed the presence of the GC and GPC5 genes. Overexpression of GPC5 is also known to inhibit cell proliferation [53]. For RWI, we identified four informative SNPs within the regions. The annotation results for these regions revealed IKBKB, HELZ, CEP112, and CACNG5. The HELZ gene is known to promote cell proliferation [54]. For STA, we identified three informative SNPs within a single region. Annotation results for this region revealed that MACROD2 was located near the markers. The MACROD2 gene is a candidate gene affecting RUH in Holstein cattle [46]. It is also associated with reduced antral follicle count (AFC) and anti-Müllerian hormone (AMH) in humans [55]. Furthermore, MACROD2 is associated with survival in dairy cattle [56], resistance to bovine tuberculosis [57], and mediation of cell growth and proliferation in humans [58].

3.2. Estimation of Genomic Prediction Accuracy

In this study, we compared genomic accuracy using various π levels (0.75, 0.90, 0.99, 0.995) and two Bayesian methods (BayesB and BayesC). Accuracies were estimated for 12 traits: ANG, BCS, BQL, CWI, FTL, FTP, HHE, RAN, RTP, RUW, RWI, and STA, at π levels of 0.75, 0.90, 0.99, and 0.995 (Table 4 and Table 5). The remaining traits and research findings are provided in the supplementary materials (Tables S2 and S3). We observed a trend where accuracy increased with increasing π levels for certain traits (ANG, BCS, BQL CWI, FTL, FTP, and RTP using DEBVexcPA; RWI using DEBVexcPA) while it decreased for others (HHE and RTP using DEBVincPA;, RUW and RWI using DEBVincPA, STA). The highest accuracy was noted for the RAN trait when DEBVincPA was used as the response variable, ranging between 0.441 and 0.447. Overall, the accuracy improved when using DEBVincPA compared with that using DEBVexcPA for all traits. Similarly, in the BayesC method, accuracy changed with different π levels. We categorized these changes as negligible (ANG, CWI, HHE, and RAN), increasing with increasing π levels (BCS, BQL, FTL, FTP, RTP, and RWI), and decreasing with increasing π levels (RUW and STA). For RAN using DEBVincPA as the response variable, the highest accuracy ranged from 0.445 to 0.451, which was slightly higher than that observed with BayesB.
Although there were observable discrepancies, the distinctions in genomic precision among varied π values did not present significant statistical variance for the traits under study. Notably, it is recognized that BayesC exhibits a heightened sensitivity to the actual distribution of marker effects compared to other Bayesian methodologies [29], suggesting that significant accuracy differences could be expected at various π levels. However, in the present study, we did not find any such differences. The accuracy estimated by BayesB was higher for some traits (BCS, FTL, and HHE using DEBVincPA; RTP using DEBVexcPA, STA), whereas BayesC estimated higher accuracy for others (ANG, BQL, CWI, FTP, and HHE using DEBVexcPA; RAN and RTP using DEBVincPA); in some cases (RUW and RWI), no differences were observed.
BayesB, assuming varied marker variance values, tends to outperform BayesC, which operates under the assumption of uniform marker variance values. Notably, BayesB shows superior effectiveness when an estimated π value is applied. Moreover, a range of factors markedly impacts the precision of genomic predictions [59]. Among these, the choice of response variable is paramount, influenced by the data’s structure and distinct attributes. Earlier research advocated for the use of estimated breeding values (EBV) as a response variable, considering the limited scope of applicable information [14,60]. EBV is often favored over DEBV when its reliability is consistent across all genotyped animals. Omitting PA from EBV is advisable to avoid redundancy and to moderate individual EBV towards the PA [11]. On the other hand, accounting for PA after deregression helps in acknowledging variances in PA among genotyped animals, such as disparities arising from inter-family differences [61].
In our study, we selected DEBV as the response variable and noted enhanced genomic precision when PA was considered. The issue of double counting seemed less critical as our model included both progeny and parental genotypes, reducing genetic linkages among genotyped animals from different family lines. Importantly, we found that body conformation and udder-related traits were associated with milk production, while traits like pelvis width were related to reproductive and calving traits. Therefore, applying a genomic selection model could lead to an increase in genetic improvement. Although heritability may not be high and genomic accuracy might not be as elevated as other milk yield-related economic traits, utilizing genomic information can still boost genetic improvement. This inclusion of PA into DEBV was particularly beneficial for the genomic selection of body conformation traits in Korean Holstein cattle. Our study serves as a pioneering effort in predicting genomic accuracy for these traits, providing a crucial reference for future research. Given the paucity of similar studies, further research involving larger groups of animals is imperative to corroborate these findings in Korean Holstein cattle.

4. Conclusions

This study marks a pivotal advancement in genomic prediction accuracy for body conformation traits in Korean Holstein cattle, enriched by key findings from GWAS and an analysis of various π levels (0.75, 0.90, 0.99, and 0.995) along with Bayesian methods (BayesB and BayesC). Our GWAS identified several novel genetic markers that were significantly associated with body conformation traits. These markers are crucial for understanding the genetic underpinnings of these traits and provide a solid foundation for genomic predictions. Notably, specific loci were found to have a profound influence on traits, such as RUW and BCS, offering new insights into cattle genetics. This study also underscores the higher predictive accuracy of DEBVincPA over DEBVexcPA for body conformation traits. This suggests that including PA in DEBV considerably enhances prediction accuracy, which is an important consideration for breeding programs. Furthermore, our investigation revealed that the choice between BayesB and BayesC should be based on the traits and genetic markers identified by the GWAS. This approach helps tailor genomic prediction strategies for specific traits, thereby optimizing the accuracy and effectiveness of selection in breeding programs.
In conclusion, the integration of GWAS findings with genomic prediction methods provides a more nuanced understanding of the genetic factors that influence body conformation traits in Korean Holstein cattle. These insights are not only pivotal for refining cattle breeding strategies but also add valuable knowledge to the field of animal genetics and GS.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/ani14071052/s1, Table S1: Informative SNPs in the significant 1 Mb windows associated with body conformation traits in Korean Holstein dairy cattle from GWAS using SNP genetic markers from the AxiomCustom300K genotyping platform; Table S2: Comparisons of the genomic prediction accuracy with standard error among genotyping platforms including AxiomCustom300K according to BayesB method with four different π values in each response variable for traits with GV less than 1.0%; Table S3: Comparisons of the genomic prediction accuracy with standard error among genotyping platforms including AxiomCustom300K according to BayesC method with four different π values in each response variable for traits with GV less than 1.0%; Figure S1: Manhattan plots of genome-wide association analysis (GWAS) based on the Bayesian C (BayesC) method for each trait. Angularity (ANG), body condition score (BCS), bone quality (BQL), chest width (CWI), front teat length (FTL), front teat placement (FTP), fore udder attachment (FUA), height at front end (HHE), rump angle (RAN), rear teat placement (RTP), rear udder width (RUW), rump width (RWI), stature (STA).

Author Contributions

Conceptualization, J.L. (Jungjae Lee), C.D. and J.P.; methodology, J.L. (Jungjae Lee), Y.K., J.L. (Jaegu Lee) and J.P.; software, J.L. (Jungjae Lee), S.L., H.S., S.K. and J.P.; validation, S.P., S.Y., J.S. (Jihyun Son) and J.S. (Jiseob Shin); formal analysis, J.L. (Jungjae Lee) and J.P.; investigation, H.M., S.P., J.K., S.Y., J.S. (Jiseob Shin) and C.P.; resources, Y.K., S.L. and C.D.; data curation, J.K., H.S. and S.K.; writing—original draft preparation, J.L. (Jungjae Lee) and J.P.; writing—review and editing, J.L. (Jungjae Lee), Y.K. and C.D.; visualization, J.L. (Jungjae Lee) and J.P.; supervision, H.M., Y.K. and C.D.; project administration, J.L. (Jungjae Lee), H.S., J.L. (Jaegu Lee) and C.D.; funding acquisition, J.L. (Jungjae Lee), H.M., Y.K., S.L. and C.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (2021R1I1A1A01047693).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

Authors Jiseob Shin, Seokhyun Lee and Hyungjun Song ware employed by the company Dairy Cattle Improvement Center of NH-Agree Business Group. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Wu, X.; Fang, M.; Liu, L.; Wang, S.; Liu, J.; Ding, X.; Zhang, S.; Zhang, Q.; Zhang, Y.; Qiao, L. Genome Wide Association Studies for Body Conformation Traits in the Chinese Holstein Cattle Population. BMC Genom. 2013, 14, 897. [Google Scholar] [CrossRef] [PubMed]
  2. Meuwissen, T.H.; Hayes, B.J.; Goddard, M.; ME1461589 Goddard. Prediction of Total Genetic Value Using Genome-Wide Dense Marker Maps. Genetics 2001, 157, 1819–1829. [Google Scholar] [CrossRef] [PubMed]
  3. Daetwyler, H.; Schenkel, F.; Sargolzaei, M.; Robinson, J. Andrew B Robinson. A Genome Scan to Detect Quantitative Trait Loci for Economically Important Traits in Holstein Cattle Using Two Methods and a Dense Single Nucleotide Polymorphism Map. J. Dairy Sci. 2008, 91, 3225–3236. [Google Scholar] [CrossRef] [PubMed]
  4. Goddard, M. Genomic Selection: Prediction of Accuracy and Maximisation of Long Term Response. Genetica 2009, 136, 245–257. [Google Scholar] [CrossRef] [PubMed]
  5. Su, G.; Brøndum, R.; Ma, P.; Guldbrandtsen, B.; Aamand, G.; Lund, M. Comparison of Genomic Predictions Using Medium-Density (∼54,000) and High-Density (∼777,000) Single Nucleotide Polymorphism Marker Panels in Nordic Holstein and Red Dairy Cattle Populations. J. Dairy Sci. 2012, 95, 4657–4665. [Google Scholar] [CrossRef] [PubMed]
  6. Habier, D.; Fernando, R.L.; Garrick, D.J. Genomic Blup Decoded: A Look into the Black Box of Genomic Prediction. Genetics 2013, 194, 597–607. [Google Scholar] [CrossRef] [PubMed]
  7. Habier, D.; Fernando, R.L.; Dekkers, J.C.M.; JCM2219482 Dekkers. The Impact of Genetic Relationship Information on Genome-Assisted Breeding Values. Genetics 2007, 177, 2389–2397. [Google Scholar] [CrossRef] [PubMed]
  8. Saatchi, M.; McClure, M.C.; McKay, S.D.; Rolf, M.M.; Kim, J.; Decker, J.E.; Taxis, T.M.; Chapple, R.H.; Ramey, H.R.; Northcutt, S.L.; et al. Accuracies of Genomic Breeding Values in American Angus Beef Cattle Using K-Means Clustering for Cross-Validation. Genet. Sel. Evol. 2011, 43, 40. [Google Scholar] [CrossRef]
  9. Habier, D.; Tetens, J.; Seefried, F.-R.; Lichtner, P.; Thaller, G. The Impact of Genetic Relationship Information on Genomic Breeding Values in German Holstein Cattle. Genet. Sel. Evol. 2010, 42, 5. [Google Scholar] [CrossRef]
  10. Calus, M.; Veerkamp, R. Accuracy of Breeding Values When Using and Ignoring the Polygenic Effect in Genomic Breeding Value Estimation with a Marker Density of One Snp Per Cm. J. Anim. Breed. Genet. 2007, 124, 362–368. [Google Scholar] [CrossRef]
  11. Garrick, D.J.; Taylor, J.F.; Fernando, R.L. Deregressing Estimated Breeding Values and Weighting Information for Genomic Regression Analyses. Genet. Sel. Evol. 2009, 41, 55. [Google Scholar] [CrossRef]
  12. Hayes, B.J.; Pryce, J.; Chamberlain, A.J.; Bowman, P.J.; Goddard, M.E. Genetic Architecture of Complex Traits and Accuracy of Genomic Prediction: Coat Colour, Milk-Fat Percentage, and Type in Holstein Cattle as Contrasting Model Traits. PLoS Genet. 2010, 6, e1001139. [Google Scholar] [CrossRef] [PubMed]
  13. Gunia, M.; Saintilan, R.; Venot, E.; Hozé, C.; Fouilloux, M.N.; Phocas, F. Genomic Prediction in French Charolais Beef Cattle Using High-Density Single Nucleotide Polymorphism Markers. J. Anim. Sci. 2014, 92, 3258–3269. [Google Scholar] [CrossRef] [PubMed]
  14. Gao, H.; Lund, M.S.; Zhang, Y.; Su, G. Accuracy of Genomic Prediction Using Different Models and Response Variables in the N Ordic R Ed Cattle Population. J. Anim. Breed. Genet. 2013, 130, 333–340. [Google Scholar] [CrossRef] [PubMed]
  15. Goddard, M.E.; Hayes, B.J. Mapping Genes for Complex Traits in Domestic Animals and Their Use in Breeding Programmes. Nat. Rev. Genet. 2009, 10, 381–391. [Google Scholar] [CrossRef] [PubMed]
  16. Abo-Ismail, M.K.; Brito, L.F.; Miller, S.P.; Sargolzaei, M.; Grossi, D.A.; Moore, S.S.; Plastow, G.; Stothard, P.; Nayeri, S. Genome-Wide Association Studies and Genomic Prediction of Breeding Values for Calving Performance and Body Conformation Traits in Holstein Cattle. Genet. Sel. Evol. 2017, 49, 82. [Google Scholar] [CrossRef] [PubMed]
  17. Dematawena, C.; Berger, P. Effect of Dystocia on Yield, Fertility, and Cow Losses and an Economic Evaluation of Dystocia Scores for Holsteins. J. Dairy Sci. 1997, 80, 754–761. [Google Scholar] [CrossRef] [PubMed]
  18. Ali, T.; Burnside, E.; Schaeffer, L. Relationship between External Body Measurements and Calving Difficulties in Canadian Holstein-Friesian Cattle. J. Dairy Sci. 1984, 67, 3034–3044. [Google Scholar] [CrossRef] [PubMed]
  19. Sawa, A.; Bogucki, M.; Krężel-Czopek, S.; Neja, W. Association between Rump Score and Course of Parturition in Cows. Arch. Anim. Breed. 2013, 56, 816–822. [Google Scholar] [CrossRef]
  20. Schmidtmann, C.; Segelke, D.; Bennewitz, J.; Tetens, J.; Thaller, G. Genetic Analysis of Production Traits and Body Size Measurements and Their Relationships with Metabolic Diseases in German Holstein Cattle. J. Dairy Sci. 2023, 106, 421–438. [Google Scholar] [CrossRef]
  21. Sargolzaei, M.; Chesnais, J.P.; Schenkel, F.S. A New Approach for Efficient Genotype Imputation Using Information from Relatives. BMC Genom. 2014, 15, 478. [Google Scholar] [CrossRef] [PubMed]
  22. Saatchi, M.; Schnabel, R.D.; Rolf, M.M.; Taylor, J.F.; Garrick, D.J. Accuracy of Direct Genomic Breeding Values for Nationally Evaluated Traits in Us Limousin and Simmental Beef Cattle. Genet. Sel. Evol. 2012, 44, 38. [Google Scholar] [CrossRef] [PubMed]
  23. Kizilkaya, K.; Fernando, R.L.; Garrick, D.J. Genomic Prediction of Simulated Multibreed and Purebred Performance Using Observed Fifty Thousand Single Nucleotide Polymorphism Genotypes. J. Anim. Sci. 2010, 88, 544–551. [Google Scholar] [CrossRef] [PubMed]
  24. Garrick, D.J.; Fernando, R.L. Implementing a Qtl Detection Study (Gwas) Using Genomic Prediction Methodology. Genome-Wide Assoc. Stud. Genom. Predict. 2013, 1019, 275–298. [Google Scholar]
  25. Xue, X.; Hu, H.; Zhang, J.; Ma, Y.; Han, L.; Hao, F.; Jiang, Y.; Ma, Y. Estimation of Genetic Parameters for Conformation Traits and Milk Production Traits in Chinese Holsteins. Animals 2022, 13, 100. [Google Scholar] [CrossRef] [PubMed]
  26. Battagin, M.; Sartori, C.; Biffani, S.; Penasa, M.; Cassandro, M. Genetic Parameters for Body Condition Score, Locomotion, Angularity, and Production Traits in Italian Holstein Cattle. J. Dairy Sci. 2013, 96, 5344–5351. [Google Scholar] [CrossRef] [PubMed]
  27. Wall, E.; White, I.; Coffey, M.; Brotherstone, S. The Relationship between Fertility, Rump Angle, and Selected Type Information in Holstein-Friesian Cows. J. Dairy Sci. 2005, 88, 1521–1528. [Google Scholar] [CrossRef] [PubMed]
  28. Xu, L.; Luo, H.; Zhang, X.; Lu, H.; Zhang, M.; Ge, J.; Zhang, T.; Yan, M.; Tan, X.; Huang, X.; et al. Factor Analysis of Genetic Parameters for Body Conformation Traits in Dual-Purpose Simmental Cattle. Animals 2022, 12, 2433. [Google Scholar] [CrossRef]
  29. Olsen, H.G.; Knutsen, T.M.; Lewandowska-Sabat, A.M.; Grove, H.; Nome, T.; Svendsen, M.; Arnyasi, M.; Sodeland, M.; Sundsaasen, K.K.; Dahl, S.R.; et al. Fine Mapping of a Qtl on Bovine Chromosome 6 Using Imputed Full Sequence Data Suggests a Key Role for the Group-Specific Component (Gc) Gene in Clinical Mastitis and Milk Production. Genet. Sel. Evol. 2016, 48, 79. [Google Scholar] [CrossRef]
  30. Cai, Z.; Guldbrandtsen, B.; Lund, M.S.; Sahana, G. Prioritizing Candidate Genes Post-Gwas Using Multiple Sources of Data for Mastitis Resistance in Dairy Cattle. BMC Genom. 2018, 19, 656. [Google Scholar] [CrossRef]
  31. Poindexter, M.B.; Kweh, M.F.; Zimpel, R.; Zuniga, J.; Lopera, C.; Zenobi, M.G.; Jiang, Y.; Engstrom, M.; Celi, P.; Santos, J.E.P.; et al. Feeding Supplemental 25-Hydroxyvitamin D3 Increases Serum Mineral Concentrations and Alters Mammary Immunity of Lactating Dairy Cows. J. Dairy Sci. 2020, 103, 805–822. [Google Scholar] [CrossRef] [PubMed]
  32. Merriman, K.E.; Powell, J.L.; Santos, J.E.; Nelson, C.D. Intramammary 25-Hydroxyvitamin D3 Treatment Modulates Innate Immune Responses to Endotoxin-Induced Mastitis. J. Dairy Sci. 2018, 101, 7593–7607. [Google Scholar] [CrossRef] [PubMed]
  33. Lippolis, J.D.; Reinhardt, T.A.; Sacco, R.A.; Nonnecke, B.J.; Nelson, C.D. Treatment of an Intramammary Bacterial Infection with 25-Hydroxyvitamin D3. PLoS ONE 2011, 6, e25479. [Google Scholar] [CrossRef] [PubMed]
  34. Purfield, D.C.; Bradley, D.G.; Evans, R.D.; Kearney, F.J.; Berry, D.P. Genome-Wide Association Study for Calving Performance Using High-Density Genotypes in Dairy and Beef Cattle. Genet. Sel. Evol. 2015, 47, 47. [Google Scholar] [CrossRef] [PubMed]
  35. Atashi, H.; Wilmot, H.; Vanderick, S.; Hubin, X.; Gengler, N. Genome-Wide Association Study for Milk Production Traits in Dual-Purpose Belgian Blue Cows. Livest. Sci. 2022, 256, 104831. [Google Scholar] [CrossRef]
  36. Costa, A.; Schwarzenbacher, H.; Mészáros, G.; Fuerst-Waltl, B.; Fuerst, C.; Sölkner, J.; Penasa, M. On the Genomic Regions Associated with Milk Lactose in Fleckvieh Cattle. J. Dairy Sci. 2019, 102, 10088–10099. [Google Scholar] [CrossRef]
  37. Kambadur, R.; Sharma, M.; Smith, T.P.; Bass, J.J. Mutations in Myostatin (Gdf8) in Double-Muscled Belgian Blue and Piedmontese Cattle. Genome Res. 1997, 7, 910–915. [Google Scholar] [CrossRef] [PubMed]
  38. Mosher, D.S.; Quignon, P.; Bustamante, C.D.; Sutter, N.B.; Mellersh, C.S.; Parker, H.G.; Ostrander, E.A. A Mutation in the Myostatin Gene Increases Muscle Mass and Enhances Racing Performance in Heterozygote Dogs. PLoS Genet. 2007, 3, e79. [Google Scholar] [CrossRef]
  39. Vogel, P.; Ding, Z.M.; Read, R.; DaCosta, C.M.; Hansard, M.; Small, D.L.; Gui-lan, Y.; Hansen, G.; Brommage, R.; Powell, D.R. Progressive Degenerative Myopathy and Myosteatosis in Asnsd1-Deficient Mice. Vet. Pathol. 2020, 57, 723–735. [Google Scholar] [CrossRef] [PubMed]
  40. Cao, M.D.; Cheng, M.; Rizwan, A.; Jiang, L.; Krishnamachary, B.; Bhujwalla, Z.M.; Bhujwalla, Z.M.; Bathen, T.F.; Glund, K. Targeting Choline Phospholipid Metabolism: Gdpd5 and Gdpd6 Silencing Decrease Breast Cancer Cell Proliferation, Migration, and Invasion. NMR Biomed. 2016, 29, 1098–1107. [Google Scholar] [CrossRef]
  41. Deng, T.; Liang, A.; Liang, S.; Ma, X.; Lu, X.; Duan, A.; Pang, C.; Hua, G.; Liu, S.; Campanile, G.; et al. Integrative Analysis of Transcriptome and Gwas Data to Identify the Hub Genes Associated with Milk Yield Trait in Buffalo. Front. Genet. 2019, 10, 36. [Google Scholar] [CrossRef]
  42. Raza, S.H.A.; Khan, S.; Amjadi, M.; Abdelnour, S.A.; Ohran, H.; Alanazi, K.M.; El-Hack, M.E.A.; Taha, A.E.; Khan, R.; Gong, C.; et al. Genome-Wide Association Studies Reveal Novel Loci Associated with Carcass and Body Measures in Beef Cattle. Arch. Biochem. Biophys. 2020, 694, 108543. [Google Scholar] [CrossRef]
  43. Chen, S.-Y.; Oliveira, H.R.; Schenkel, F.S.; Pedrosa, V.B.; Melka, M.G.; Brito, L.F. Using Imputed Whole-Genome Sequence Variants to Uncover Candidate Mutations and Genes Affecting Milking Speed and Temperament in Holstein Cattle. J. Dairy Sci. 2020, 103, 10383–10398. [Google Scholar] [CrossRef]
  44. Halli, K.; Vanvanhossou, S.F.; Bohlouli, M.; König, S.; Yin, T. Identification of Candidate Genes on the Basis of Snp by Time-Lagged Heat Stress Interactions for Milk Production Traits in German Holstein Cattle. PLoS ONE 2021, 16, e0258216. [Google Scholar] [CrossRef]
  45. Marete, A.; Sahana, G.; Fritz, S.; Lefebvre, R.; Barbat, A.; Lund, M.S.; Guldbrandtsen, B.; Boichard, D. Genome-Wide Association Study for Milking Speed in French Holstein Cows. J. Dairy Sci. 2018, 101, 6205–6219. [Google Scholar] [CrossRef]
  46. Gonzalez, M.; Villa, R.; Villa, C.; Gonzalez, V.; Montano, M.; Medina, G.; Mahadevan, P. Inspection of Real and Imputed Genotypes Reveled 76 Snps Associated to Rear Udder Height in Holstein Cattle. J. Adv. Vet. Anim. Res. 2020, 7, 234. [Google Scholar] [CrossRef]
  47. Zhang, B.; Chang, L.; Lan, X.; Asif, N.; Guan, F.; Fu, D.; Li, B.; Yan, C.; Zhang, H.; Zhang, X.; et al. Genome-Wide Definition of Selective Sweeps Reveals Molecular Evidence of Trait-Driven Domestication among Elite Goat (Capra Species) Breeds for the Production of Dairy, Cashmere, and Meat. GigaScience 2018, 7, giy105. [Google Scholar] [CrossRef]
  48. Sanna, S.; Jackson, A.U.; Nagaraja, R.; Willer, C.J.; Chen, W.M.; Bonnycastle, L.L.; Shen, H.; Timpson, N.; Lettre, G.; Usala, G.; et al. Common Variants in the Gdf5-Uqcc Region Are Associated with Variation in Human Height. Nat. Genet. 2008, 40, 198–203. [Google Scholar] [CrossRef]
  49. Cobb, J.; Dierich, A.; Huss-Garcia, Y.; Duboule, D. A Mouse Model for Human Short-Stature Syndromes Identifies Shox2 as an Upstream Regulator of Runx2 During Long-Bone Development. Proc. Natl. Acad. Sci. USA 2006, 103, 4511–4515. [Google Scholar] [CrossRef]
  50. Kumar, J.P.; Kulkarni, R.; Kulkarni, R.N. A Study on the Anomalies Associated with the Human Sterna in South Indian Population. J. Anat. Soc. India. 2015, 64, S34. [Google Scholar] [CrossRef]
  51. Sanchez, M.P.; Rocha, D.; Charles, M.; Boussaha, M.; Hozé, C.; Brochard, M.; Grosperrin, P. Sequence-Based Gwas and Post-Gwas Analyses Reveal a Key Role of Slc37a1, Ankh, and Regulatory Regions on Bovine Milk Mineral Content. Sci. Rep. 2021, 11, 7537. [Google Scholar] [CrossRef]
  52. Dobbs, K.; Domínguez Conde, C.; Zhang, S.Y.; Parolini, S.; Audry, M.; Chou, J.; Haapaniemi, E.; Keles, S.; Bilic, I.; Okada, S.; et al. Inherited Dock2 Deficiency in Patients with Early-Onset Invasive Infections. N. Engl. J. Med. 2015, 372, 2409–2422. [Google Scholar] [CrossRef]
  53. Yuan, S.; Yu, Z.; Liu, Q.; Zhang, M.; Xiang, Y.; Wu, N.; Wu, L.; Hu, Z.; Xu, B.; Cai, T.; et al. Gpc5, a Novel Epigenetically Silenced Tumor Suppressor, Inhibits Tumor Growth by Suppressing Wnt/Β-Catenin Signaling in Lung Adenocarcinoma. Oncogene 2016, 35, 6120–6131. [Google Scholar] [CrossRef]
  54. Hasgall, P.A.; Hoogewijs, D.; Faza, M.B.; Panse, V.G.; Wenger, R.H.; Camenisch, G. The Putative Rna Helicase Helz Promotes Cell Proliferation, Translation Initiation and Ribosomal Protein S6 Phosphorylation. PLoS ONE 2011, 6, e22107. [Google Scholar] [CrossRef]
  55. Schuh-Huerta, S.M.; Johnson, N.A.; Rosen, M.P.; Sternfeld, B.; Cedars, M.I.; Reijo Pera, R.A. Genetic Markers of Ovarian Follicle Number and Menopause in Women of Multiple Ethnicities. Hum. Genet. 2012, 131, 1709–1724. [Google Scholar] [CrossRef]
  56. Shabalina, T.; Yin, T.; König, S. Survival Analyses in Holstein Cows Considering Direct Disease Diagnoses and Specific Snp Marker Effects. J. Dairy Sci. 2020, 103, 8257–8273. [Google Scholar] [CrossRef]
  57. González-Ruiz, S.; Strillacci, M.G.; Durán-Aguilar, M.; Cantó-Alarcón, G.J.; Herrera-Rodríguez, S.E.; Bagnato, A.; Guzmán, L.F.; Milián-Suazo, F.; Román-Ponce, S.I. Genome-Wide Association Study in Mexican Holstein Cattle Reveals Novel Quantitative Trait Loci Regions and Confirms Mapped Loci for Resistance to Bovine Tuberculosis. Animals 2019, 9, 636. [Google Scholar] [CrossRef]
  58. Mohseni, M.; Cidado, J.; Croessmann, S.; Cravero, K.; Cimino-Mathews, A.; Wong, H.Y.; Scharpf, R.; Zabransky, D.J.; Abukhdeir, A.M.; Garay, J.P.; et al. Macrod2 Overexpression Mediates Estrogen Independent Growth and Tamoxifen Resistance in Breast Cancers. Proc. Natl. Acad. Sci. USA 2014, 111, 17606–17611. [Google Scholar] [CrossRef]
  59. Fernando, R.L.; Garrick, D. Bayesian Methods Applied to Gwas. Methods Mol Biol. 2013, 1019, 237–274. [Google Scholar]
  60. Guo, G.; Lund, M.; Zhang, Y.; Su, G. Comparison between Genomic Predictions Using Daughter Yield Deviation and Conventional Estimated Breeding Value as Response Variables. J. Anim. Breed. Genet. 2010, 127, 423–432. [Google Scholar] [CrossRef]
  61. Lee, J.; Su, H.; Fernando, R.L.; Garrick, D.J.; Taylor, J. Characterization of the F94l Double Muscling Mutation in Pure-and Crossbred Limousin Animals. Anim. Ind. Rep. 2015, 661, 19. [Google Scholar]
Table 1. Description, variance components, and genetic parameters of body conformation traits in Korean Holsteins.
Table 1. Description, variance components, and genetic parameters of body conformation traits in Korean Holsteins.
TraitAbbreviation σ a 2 * σ e 2 * h 2 *
Body Trait
AngularityANG0.1060.7770.120
Body condition scoreBCS0.1090.4740.187
Body depthBDE0.2180.6050.265
Chest widthCWI0.1240.6710.156
Height at front endHHE0.0270.3020.082
LocomotionLOC0.0270.8440.031
Overall conformation scoreOCS1.0275.5990.155
StatureSTA0.4110.8730.320
Rump Trait
Loin strengthLST0.0750.6830.099
Rump angleRAN0.5271.1900.307
Rump widthRWI0.1640.8060.169
Feet and leg traits
Bone qualityBQL0.0640.5290.108
Foot angleFAN0.0731.0170.067
Heel depth/foot heightHDE0.0310.5130.057
Rear leg rear viewRLR0.0921.0730.079
Rear leg setRLS0.1080.8230.116
Udder traits
Front teat lengthFTL0.3281.2190.212
Front teat placementFTP0.1940.9340.172
Fore udder attachmentFUA0.1781.1700.132
Rear teat placementRTP0.0770.7690.091
Rear udder heightRUH0.2521.2390.169
Rear udder widthRUW0.1140.9320.109
Udder depthUDE0.4180.8330.334
Udder supportUSU0.1150.9600.107
Udder textureUTX0.0770.7690.091
*, additive genetic variance, residual variance, heritability.
Table 2. Comparison of relationships among animals within and across clusters in a 5-fold cross-validation.
Table 2. Comparison of relationships among animals within and across clusters in a 5-fold cross-validation.
ClusterNo. of AnimalsinBreC 1amax_within 2amax_between 3aij_within 4aij_between 5
115850.0490.5040.4010.1820.095
217150.0180.3620.3810.0450.064
318400.0410.4490.4560.0980.092
420390.0520.5160.4260.1650.100
539160.0490.5380.4300.1590.092
Avg.0.0420.4740.4190.1300.089
1, the average of inbreeding coefficients within cluster; 2, the average of amax (the maximum of relationships [aij] for each animal) values within cluster; 3, the average of amax values between the clustered (training and validation) groups; 4, the average of aij (relationships) values within cluster; 5, the average of aij values between clustered groups.
Table 3. Informative SNPs in the significant 1 Mb windows associated with each trait in Korean Holstein dairy cattle from GWAS using SNP genetic markers from the AxiomCustom300K genotyping platform.
Table 3. Informative SNPs in the significant 1 Mb windows associated with each trait in Korean Holstein dairy cattle from GWAS using SNP genetic markers from the AxiomCustom300K genotyping platform.
Trait 1BTA
_MB 2
nSNPsGV
(%) 3
Informative SNPModel
_Freq 4
VariantGene Annotation 5
ANG6_88181.52AX-1067319670.026intergenicGC (152,138)
BCS2_61741.39AX-4279024380.024intronANKAR
AX-4197933080.018downstream geneMSTN (4600)
AX-4197830710.0745_prime_UTRASNSD1
AX-3721089270.021splice regionOSGEPL1
AX-3208877550.055missenseASNSD1
AX-3208815170.019downstream geneMSTN (471)
AX-3105032260.050missenseASNSD1
AX-1170900490.031intergenicASNSD1 (29,106)
BQL6_88181.72AX-1067319670.022intergenicGC (152,138)
15_55281.31AX-1151084550.011intergenicENSBTAG00000046149 (6310)
AX-1067514110.602intronGDPD5
AX-1067348280.026intergenicENSBTAG00000000628 (18,676)
10_33731.16AX-4294602210.150intergenicMEIS2 (276,619)
AX-4288726780.016intergenicMEIS2 (236,727)
AX-4281464060.039intergenicTMCO5A (34,691)
CWI14_76531.85AX-1695156190.015synonymousDECR1
AX-1151073890.020missenseDECR1
4_115661.16AX-4296986510.323intergenicCCT8L2 (65,651)
AX-4294117020.012intergenicXRCC2 (35,174)
AX-4292522900.024intronKMT2C
AX-4292008770.397intronKMT2C
AX-4283778470.021intronGALNTL5
11_52411.00AX-4291876270.612intergenicCTNNA2 (1,899,780)
AX-4280776810.012intergenicCTNNA2 (1,912,867)
AX-1816354430.036intergenicCTNNA2 (2,016,344)
FTL11_82441.60AX-4297508720.890intergenicFAM49A (120,784)
FTP1_110421.43AX-1851144270.107intergenicSHOX2 (136,580)
AX-1067262810.017intergenicSHOX2 (13,662)
7_30821.13AX-4295488870.017intergenicENSBTAG00000047546 (303,880)
AX-4280512780.272intergenicZNF608 (329,604)
AX-4280255130.017intronZNF608
AX-1851218780.015intergenicENSBTAG00000047546 (270,407)
AX-1816360850.024intergenicENSBTAG00000047546 (194,374)
AX-1243758410.168intergenicZNF608 (53,173)
AX-1243749440.019intergenicENSBTAG00000047546 (174,952)
AX-1151071300.034intergenicZNF608 (85,253)
AX-1067580930.020intergenicENSBTAG00000047546 (208,718)
AX-1067315710.037intergenicZNF608 (13,676)
HHE20_58852.25AX-4295745570.017intergenicANKH (89,616)
AX-1067500020.733intronTRIO
RAN7_54411.04AX-4294643520.642intronFCHSD1
RTP6_88771.53AX-4280943980.691intronDOCK2
19_1531.42AX-4298320350.030intronCA10
AX-4288910260.018intronCA10
AX-3209103170.547intronCA10
AX-3105327330.062intronCA10
RUW6_88182.41AX-1067319670.140intergenicGC (152,138)
AX-1851113440.021intronGPC5
AX-1067218080.152intronGPC5
RWI27_36401.17AX-4290285920.726intronIKBKB
19_63461.12AX-4288059940.047intronHELZ
AX-3209135750.047intronCEP112
AX-1151169400.068intergenicCACNG5 (19,442)
AX-1067443480.021intergenicHELZ (25,925)
AX-1067209920.045intronHELZ
STA13_9931.07AX-4289764080.841IntergenicMACROD2 (340,366)
AX-4287131250.028IntergenicMACROD2 (299,032)
AX-3105511370.019intergenicMACROD2 (546,675)
1, angularity (ANG), body condition score (BCS), bone quality (BQL), chest width (CWI), front teat length (FTL), front teat placement (FTP), height at front end (HHE), rump angle (RAN), rear teat placement (RTP), rear udder width (RUW), rump width (RWI), stature (STA); 2, Bos taurus chromosome; 3, percentage of additive genetic variance explained by SNP markers within each 1 Mb window region; 4, proportion of Markov Chain–Monte Carlo iterations that included the corresponding SNP marker; 5, gene symbols when intronic or gene symbols (distance) adjacent to the marker were intergenic.
Table 4. Comparisons of genomic prediction accuracy with standard error among genotyping platforms including AxiomCustom300K according to BayesB method with four different π values in each response variable.
Table 4. Comparisons of genomic prediction accuracy with standard error among genotyping platforms including AxiomCustom300K according to BayesB method with four different π values in each response variable.
Traits 1resVars 2BayesB (with π)
0.750.90.990.995
ANGDEBVexcPA0.038 (0.032)0.040 (0.032)0.052 (0.031)0.057 (0.032)
DEBVincPA0.126 (0.031)0.128 (0.031)0.139 (0.031)0.144 (0.031)
BCSDEBVexcPA0.135 (0.033)0.136 (0.033)0.145 (0.033)0.152 (0.033)
DEBVincPA0.218 (0.032)0.220 (0.032)0.232 (0.032)0.237 (0.032)
BQLDEBVexcPA0.104 (0.034)0.105 (0.034)0.113 (0.034)0.118 (0.034)
DEBVincPA0.177 (0.033)0.180 (0.033)0.193 (0.033)0.195 (0.033)
CWIDEBVexcPA0.077 (0.032)0.077 (0.032)0.091 (0.032)0.100 (0.032)
DEBVincPA0.156 (0.032)0.154 (0.032)0.159 (0.031)0.159 (0.032)
FTLDEBVexcPA0.241 (0.031)0.245 (0.031)0.269 (0.030)0.276 (0.030)
DEBVincPA0.328 (0.029)0.333 (0.029)0.353 (0.029)0.356 (0.029)
FTPDEBVexcPA0.210 (0.031)0.214 (0.031)0.236 (0.031)0.243 (0.030)
DEBVincPA0.293 (0.030)0.297 (0.029)0.309 (0.029)0.309 (0.029)
HHEDEBVexcPA0.112 (0.035)0.113 (0.035)0.120 (0.035)0.124 (0.035)
DEBVincPA0.186 (0.034)0.186 (0.034)0.186 (0.034)0.184 (0.034)
RANDEBVexcPA0.362 (0.031)0.366 (0.031)0.377 (0.030)0.373 (0.030)
DEBVincPA0.445 (0.029)0.447 (0.029)0.447 (0.029)0.441 (0.029)
RTPDEBVexcPA0.247 (0.031)0.248 (0.031)0.254 (0.031)0.254 (0.031)
DEBVincPA0.281 (0.030)0.283 (0.030)0.290 (0.030)0.288 (0.030)
RUWDEBVexcPA0.109 (0.032)0.108 (0.032)0.097 (0.032)0.090 (0.032)
DEBVincPA0.182 (0.031)0.181 (0.031)0.174 (0.031)0.170 (0.031)
RWIDEBVexcPA0.277 (0.031)0.278 (0.031)0.285 (0.030)0.287 (0.030)
DEBVincPA0.339 (0.030)0.341 (0.030)0.343 (0.030)0.340 (0.030)
STADEBVexcPA0.312 (0.036)0.267 (0.031)0.260 (0.031)0.246 (0.031)
DEBVincPA0.350 (0.030)0.350 (0.030)0.340 (0.030)0.330 (0.030)
1, angularity (ANG), body condition score (BCS), bone quality (BQL), chest width (CWI), front teat length (FTL), front teat placement (FTP), height at front end (HHE), rump angle (RAN), rear teat placement (RTP), rear udder width (RUW), rump width (RWI), stature (STA); 2, DEBVexcPA = deregressed-EBV excluding parent average; DEBVincPA = deregressed-EBV including parent average.
Table 5. Comparisons of genomic prediction accuracy with standard error among genotyping platforms, including AxiomCustom300K according to the BayesC method, with four different π values in each response variable.
Table 5. Comparisons of genomic prediction accuracy with standard error among genotyping platforms, including AxiomCustom300K according to the BayesC method, with four different π values in each response variable.
Traits 1resVars 2BayesC (with π)
0.750.90.990.995
ANGDEBVexcPA0.102 (0.031)0.101 (0.031)0.099 (0.031)0.098 (0.031)
DEBVincPA0.158 (0.031)0.158 (0.031)0.157 (0.031)0.158 (0.031)
BCSDEBVexcPA0.137 (0.033)0.138 (0.033)0.140 (0.033)0.144 (0.033)
DEBVincPA0.218 (0.032)0.218 (0.032)0.224 (0.032)0.230 (0.032)
BQLDEBVexcPA0.137 (0.033)0.137 (0.033)0.138 (0.033)0.141 (0.033)
DEBVincPA0.176 (0.033)0.177 (0.033)0.188 (0.033)0.196 (0.033)
CWIDEBVexcPA0.093 (0.032)0.093 (0.032)0.101 (0.032)0.107 (0.032)
DEBVincPA0.170 (0.031)0.172 (0.031)0.173 (0.031)0.171 (0.031)
FTLDEBVexcPA0.237 (0.031)0.239 (0.031)0.262 (0.030)0.272 (0.030)
DEBVincPA0.325 (0.029)0.328 (0.029)0.350 (0.029)0.356 (0.029)
FTPDEBVexcPA0.249 (0.030)0.249 (0.030)0.254 (0.030)0.260 (0.030)
DEBVincPA0.301 (0.029)0.301 (0.029)0.310 (0.029)0.313 (0.029)
HHEDEBVexcPA0.123 (0.035)0.123 (0.035)0.125 (0.035)0.127 (0.035)
DEBVincPA0.170 (0.035)0.170 (0.035)0.169 (0.035)0.170 (0.035)
RANDEBVexcPA0.372 (0.030)0.372 (0.030)0.376 (0.030)0.374 (0.030)
DEBVincPA0.448 (0.029)0.448 (0.029)0.451 (0.029)0.445 (0.029)
RTPDEBVexcPA0.232 (0.031)0.232 (0.031)0.240 (0.031)0.245 (0.031)
DEBVincPA0.280 (0.030)0.280 (0.030)0.288 (0.030)0.290 (0.030)
RUWDEBVexcPA0.109 (0.032)0.108 (0.032)0.103 (0.032)0.097 (0.032)
DEBVincPA0.182 (0.031)0.182 (0.031)0.178 (0.031)0.176 (0.031)
RWIDEBVexcPA0.275 (0.031)0.276 (0.031)0.282 (0.030)0.285 (0.030)
DEBVincPA0.338 (0.030)0.339 (0.030)0.346 (0.030)0.346 (0.030)
STADEBVexcPA0.248 (0.031)0.247 (0.031)0.241 (0.031)0.235 (0.031)
DEBVincPA0.342 (0.030)0.342 (0.030)0.336 (0.030)0.329 (0.030)
1, angularity (ANG), body condition score (BCS), bone quality (BQL), chest width (CWI), front teat length (FTL), front teat placement (FTP), height at front end (HHE), rump angle (RAN), rear teat placement (RTP), rear udder width (RUW), rump width (RWI), stature (STA); 2, DEBVexcPA = deregressed-EBV excluding parent average; DEBVincPA = deregressed-EBV including parent average.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lee, J.; Mun, H.; Koo, Y.; Park, S.; Kim, J.; Yu, S.; Shin, J.; Lee, J.; Son, J.; Park, C.; et al. Enhancing Genomic Prediction Accuracy for Body Conformation Traits in Korean Holstein Cattle. Animals 2024, 14, 1052. https://0-doi-org.brum.beds.ac.uk/10.3390/ani14071052

AMA Style

Lee J, Mun H, Koo Y, Park S, Kim J, Yu S, Shin J, Lee J, Son J, Park C, et al. Enhancing Genomic Prediction Accuracy for Body Conformation Traits in Korean Holstein Cattle. Animals. 2024; 14(7):1052. https://0-doi-org.brum.beds.ac.uk/10.3390/ani14071052

Chicago/Turabian Style

Lee, Jungjae, Hyosik Mun, Yangmo Koo, Sangchul Park, Junsoo Kim, Seongpil Yu, Jiseob Shin, Jaegu Lee, Jihyun Son, Chanhyuk Park, and et al. 2024. "Enhancing Genomic Prediction Accuracy for Body Conformation Traits in Korean Holstein Cattle" Animals 14, no. 7: 1052. https://0-doi-org.brum.beds.ac.uk/10.3390/ani14071052

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