Next Article in Journal
Identifying the Best Herbicides for Weed Control in Chicory (Cichorium intybus)
Previous Article in Journal
Bacillus aryabhattai Mitigates the Effects of Salt and Water Stress on the Agronomic Performance of Maize under an Agroecological System
Previous Article in Special Issue
DIO1 Gene Polymorphism Is Associated with Thyroid Profiles and Reproductive Performance in Dairy Cows
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Overview of SNPs Associated with Trans Fat Content in Cow’s Milk

1
Department of Biotechnology and Food Products, Faculty of Biotechnology and Food Engineering, Federal State Budgetary Educational Institution of Higher Education Urals State Agricultural University, 620075 Yekaterinburg, Russia
2
L.K. Ernst Federal Research Center for Animal Husbandry, 142132 Podolsk, Russia
*
Author to whom correspondence should be addressed.
Agriculture 2023, 13(6), 1151; https://doi.org/10.3390/agriculture13061151
Submission received: 21 March 2023 / Revised: 10 May 2023 / Accepted: 27 May 2023 / Published: 30 May 2023
(This article belongs to the Special Issue Molecular Markers in Farm Animal Breeding and Genome Analysis)

Abstract

:
Lipids consumed with milk derivatives are one of the main parts of the human diet. Trans fatty acids in milk are causing a debate about their impact on the incidence of cardiovascular disease, pathological abnormalities, and cancer. The fatty acid profile of milk is influenced by a large number of different factors, one of which is genetic. The development of genetic studies, including Genome-Wide Association Studies (GWAS), may help define genomic regions associated with fatty acid content in milk, including trans fatty acids. This article provides an overview of international studies on the identification of genomic regions and SNPs associated with the trans fatty acids in cow’s milk. The results are based on research of cattle such as Norwegian Red cattle, Holstein, Jersey, and Brown Swiss. The presented review shows that 68 SNPs were localized on chromosomes 1, 2, 4–6, 8–10, 12, 14–20, 22–25, and 27–29. Further research in this direction will provide new information that will serve as an impetus for the creation of modern breeding technologies and increase the performance of the manufacture of high-quality dairy products. The search for genetic markers associated with the content of TFA in milk is a promising direction in agricultural science and will allow more complete breeding work with cattle.

1. Introduction

In recent years, the problem of rationing the maintenance of non-saturated fatty acids trans-isomers (TFA) in animal origin products has remained relevant all over the world, while Russian legislation practically does not regulate the content of TFA, which is hazardous to human health.
TFAs are spatial isomers of naturally occurring unsaturated fatty acids (FA) found in animal fats. They are formed in small quantities in the rumen of ruminants during the digestion of plant materials rich in unsaturated fatty acids, and they enter animal fats, including milk fat, and then transfer into milk processing products with a high fat content, such as sour cream and butter. The proportion of cow’s milk, cheese, and butter TFAs ranges from 3.2% to 6.2% of the total amount of fatty acids. Goat and sheep dairy products contain from 2.7% to 7.1% of these compounds [1,2,3].
Feeding lipids in the rumen undergo two important microbial transformations: lipolysis and biohydrogenation (decomposition of FA double-carbon bonds under the action of metabolic hydrogen). As a result of the rapid and intense biohydrogenation of double bonds of unsaturated fatty acids released during synthesis in the rumen, saturated fatty acids are formed. The initial stage in this process is an isomerization reaction catalyzed by the isomerase, that converts the double bond from the cis- to the trans-isomer.
Milk fat contains TFAs, represented mostly by C18:1 isomers. The most common among them is vaccenic acid (18:1t11), which accounts for more than 60% of the total TFAs. Vaccenic acid can be transformed to rumenic acid by the action of the stearol coenzyme A-desaturase. Isomers C16:1 and C18:2 make up about 15% of the total TFAs. Milk obtained in summer contains twice as many trans fatty acids as winter milk [4,5].
The effects of TFAs on the body and human health are varied and depend on many factors, such as the origin, structure, composition, and type. Despite extensive discussion of the adverse health effects of industrial TFAs, the impact of natural TFAs on health is still doubtful [6].
According to a number of foreign researchers, TFA can block the mechanisms of gene expression associated with lipid compounds at the level of metabolism [7,8,9].
An anticancer effect of TFAs has been established, which showed an inhibition of tumor development and stopped the formation of metastases due to impaired signaling, suppression of DNA synthesis, and the process of programmed death of tumor cells. In addition, there is an increased synthesis of cytokines and immunoglobulins A (IgA) and M (IgM), and the production of inflammation signals is also reduced. TFAs contribute to a reduction in the proportion of fat reserves while maintaining muscle mass through their influence on hormones that regulate lipid metabolism [10,11,12,13].
Milk TFAs are natural and are able to provide many health benefits, including boosting immune function, reducing excess fat, and completely removing or slowing down the process of neoplasm of tumors of various types of cancer [14,15]. This information warrants further research as it suggests that naturally occurring TFAs are metabolized very differently from industrially produced ones [15].
The composition of fatty acids in cow’s milk depends on many factors, but the quality of the diet and [16] habitat play the greatest role. However, data obtained on lactating dairy cows showed that even in herds in which all cows are kept and fed similarly, there is a three-fold distinction in the concentration of CLA in milk fat [17]. Differences in milk fatty acid content between various breeds have been proven by different groups of scientists. Maurice-Van Eijndhoven et al. [18], when comparing four breeds bred in the Netherlands, revealed a clear and significant difference in the concentration of short- and medium-chain acids and linoleic acid (C18:2c9,t11), and in the percentage of fat in milk. Carroll et al. [14] and Ramalho et al. [19] observed similar differences between breeds in milk fat composition and concentration. The advent of technologies for relatively inexpensive whole-genome genotyping of cows makes it possible to study genetic changes in cattle in order to study the genetic architecture of trans fatty acids in cow’s milk.
This article presents a review of studies looking for genomic regions and single-nucleotide polymorphisms (SNPs) associated with trans fatty acids in cow’s milk, collecting information on their effects on various traits in cattle.

2. SNPs Associated with Trans Fat Content in Cow’s Milk

The Cattle Quantitative Trait Locus (QTL) Database (Cattle QTLdb) contains data on four large studies on the search for SNPs related to the composition and content of fatty acids in cow’s milk, which were performed on Norwegian Red cattle [20], Holstein [21,22], Jersey [23], and Brown Swiss cattle [23,24,25], Table 1. Such SNPs were localized on chromosomes 1, 2, 4–6, 8–10, 12, 14–20, 22–25, and 27–29.
Olsen et al. [20] performed a GWAS of the fat composition of cow’s milk, including C18:1t9, C18:1t10, C18:1t11, and FA percentage, with trans bonds in Norwegian red cows according to FTIR data. They found that rs29019625, located on BTA1, was strongly associated with C18:1t11, while other identified SNPs were less expressed. Ibeagha-Awemu et al. [21] found an association between FADS1 and FADS2 gene polymorphisms in Holstein cattle with the levels of trans fatty acids in milk: rs210169303 with C18:2t10 and C18:2t6 (p ≤ 0.05), rs42187261 with C18:2t6 (p ≤ 0.05), and rs109772589 with C14:1m (p ≤ 0.05). An association of the KK genotype for DGAT1_K232A with the amount of C18:1t11 (−0.082, p ≤ 0.05) was found in the Danish Jersey and Holstein breeds [22]. Strillacci et al. [23] identified SNPs rs42176310, rs110847444, rs41575963, rs42357017, rs41578757, and rs109178989 as significantly associated with C18:1t11 content in the Brown Swiss cattle population. Pegolo et al. [24,25], in 2 experiments on Brown Swiss cattle, first studied the associations of 38 different SNPs belonging to 30 genes with 57 fatty acid traits in milk. They revealed a relationship between rs110757796 and C16:1t9 content (effect = 0.006), between rs41255713, rs110757796, and rs109578101 (effect = 0.003, 0.004, and −0.005, respectively), between rs43705173, rs110937773, and C16:1t6–8 (effect = −0.006 and 0.005, respectively), between rs110445169 and C18:1t9 (effect = −0.005), and rs43703011 and rs109136815 were associated with C18:1t10 (effect = −0.014 and 0.11, respectively). An association was found between rs43705173, rs43706906, and rs43710288 and the content of C18:1t16 (effect = −0.008, −0.006, and −0.006, respectively). Additionally, three SNPs, rs29004170, rs43703013, and rs110684599, were associated with rumenic acid content (C18:2c9,t11) (effect = −0.020, −0.022, and 0.036, respectively). SNP rs29004170 was associated with C18:2t11,c15 (effect = −0.008), and rs211032652 with C18:2t9,t12 (effect = −0.014) content. Two SNPs showed an association with C18:3c9,t11,c15: rs41624917 and rs137182814 (effect = 0.003 and −0.043, respectively). In their next experiment, they performed a GWAS with 65 milk fatty acid traits in Brown Swiss cattle and found a total of 19 SNPs significantly associated with TFA content in cow’s milk: C18:1t6–8, C16:1t9, C18:1t11, C18:1t16, C18:1t4, C18:1t10, C18:1t4, C18:1t11, C18:1t16, C18:2c9,t11, C18:2t11,c15, and C18:3c9,t11,c15.

3. Candidate Genes for Trans Fat Content in Cow’s Milk

3.1. Candidate Genes with Predicted Functions

Studies to identify an association of the trans fatty acid content in cow’s milk with SNP were based on two approaches. In the first approach, an analysis of the polymorphism of genes involved in the synthesis of fatty acids, proteins, and in the functioning of the immune system, or located in genomic regions associated with quality and technological traits of milk, was performed [21,22,24]. In one case, 5 potentially functional SNPs (1 synonymous, 1 non-synonymous, 1 splicing site, and 2 3′UTR mutations) and 2 randomly selected intron mutations in the FADS1 and FADS2 genes were genotyped in 450 samples, and an association analysis was performed, including the content of fatty acids in milk [21]. In the second case, the effect of the DGAT_K232A mutation on the content of fatty acids in the milk of Holstein and Jersey cows was evaluated [22]. In the third case, the association of 51 SNPs, selected using functional and positional approaches, with 47 fatty acids, 9 fatty acid groups, and 5 Δ9-desaturation indices, was studied in milk samples from Brown Swiss cows [24].
STAT1 (signal transducer and activator of transcription 1-alpha/beta) is a member of the transcription factor family of signal transducers and transcription activators. STAT1 is involved in the upregulation of genes by interferon signals. Significant differences were found between estimates of the breeding value of milk productivity traits in cattle with different genotypes for the STAT1 gene [26,27].
Leptin (LEP) occupies the most important role in body weight control and energy balance. The association of leptin gene polymorphism with the increase in milk yield and the quality of milk during three lactations of Holstein cows was revealed. Animals with the TT genotype for the LEP gene (c.73T > C) showed the best result for three lactations during the experiments [28]. Polymorphism of the LEP-Sau3AI was associated with the somatic cell count (SCC) (p  ≤ 0.01), electrical conductivity (EC) (p  ≤  0.01), and pH (p ≤ 0.05) in Holstein cow milk. The BB genotype led to a higher SCC, EC, and pH compared to other genotypes [29]. Maletić et al. [30] studied the total somatic cell count (SCC), chemical composition of milk (fat, protein, lactose, total solids, and percent of solids-not-fat), and the evaluation of freezing point depression (FPD) in milk of Busha cows with different genotypes of LEP-Sau3AI. Animals with the AA genotype had an average SNF content (8.74%, p = 0.021) in milk, significantly lower compared to those with genotype AB (9.28%), while cows with genotype AA had significantly higher average FPD values (−0.54 °C, p = 0.004) than those with the AB genotype (−0.58 °C). The genotype of the LEP gene was significant for five individual saturated and unsaturated FAs and the branched-chain fatty acids (BCFA) group [31]. Haruna et al. [32] investigated the three variations (A3, B3, and C3) in Holstein Friesian × Jersey (HF × J) dairy cows in New Zealand and studied their effects on the milk fatty acid composition. The A3 variant was associated with a reduced level of C15:1, C18:1t11, C18:2t9,c12, C22:0, and C24:0, but with elevated levels of C12:1 and C13:0 iso (p ≤ 0.05). The B3 variant was associated with decreased levels of C6:0, C8:0, C11:0, C13:0, and C20:0, but with elevated C17:0 iso and C24:0 levels (p ≤ 0.05). The C3 variant was associated with reduced C17:0 iso levels but elevated C20:0 levels (p ≤ 0.05). The A3B3 genotype was found to be associated with reduced levels of C22:0 and C24:0 but elevated C8:0, C10:0, C11:0, C13:0, C15:0, and grouped medium-chain fatty acid (MCFA) levels (p ≤ 0.05). Genotype A3C3 was associated with decreased levels of C10:0, C11:0, C13:0, and aggregated MCFA (p ≤ 0.05).
The functions of the genes used in the first approach have been somewhat studied in detail. The surface properties of casein micelles are highly dependent on beta-casein (CSN2), and rs43703011 is associated with A2 milk [33,34]. The statistical analysis showed that polymorphism of the CSN2 gene had a significant effect on the protein content in the milk of the Slovak Holstein cattle. The percentage of fat in milk in cows with the AA genotype is increased compared to the A2A2 genotype. Amalfitano et al. [35] identified the influence of the CSN2 genotypes on the cattle milk protein profile of the Brown Swiss cows.
Diacylglycerol O-acyltransferase 1 (DGAT1) catalyzes triacylglycerol synthesis by using diacylglycerol and fatty acyl CoA as substrates [36]. It is involved in the esterification of exogenous fatty acids to glycerol in the liver, plays an important role in the synthesis of fat for storage, and is expressed in the female mammary glands, where it produces milk fat [36,37]. Polymorphism at position rs109326954 leads to p.A232K substitution, which is associated with protein and fat content in milk [38]. Elzaki et al. [39] found a significant effect of the DNA marker rs109234250 (DGAT1_K232A) on milk yield (p = 7.6 × 10−20), fat yield (p = 2.2 × 10−17), protein yield (p = 2.0 × 10−19), and lactose yield (p = 4.0 × 10−18) in crossbred Butana × Holstein cattle. The breeding value for the amounts of milk (in kg) of animals with the AA genotype was significantly (p ≤ 0.0001) higher than in animals with the KK genotype [40]. The polymorphism of the DGAT1 gene in position p.A232K influenced the fatty acid composition: milk from AA cows had a more favorable fatty acid composition due to a lower total saturated fatty acids content and higher levels of oleic acid and total unsaturated fatty acids, a higher ratio of the saturated to unsaturated acids, and a higher atherogenic index [41].
AGPAT6 plays an important role in the process of synthesis of triglycerides (TG) in mammals. For further use in cattle breeding, the AGPAT6 gene is one of the potential candidates due to its ability to regulate the synthesis of milk fat. The SNP g.36,175,805C > T had a significant (p ≤ 0.05) influence on the EBV of fat percentage (EBV-FP) in Karan Fries Breeding Bulls. The KF bulls with the TT genotype had a comparatively lower EBV-FP than the bulls with the CC and CT genotypes. The substitution of the C allele with the T allele led to a decline of 0.0045% in the EBV-FP [42]. Wavenumbers studied by using Fourier transform infrared milk spectra and GWAS in Danish Jersey cows were associated with the AGPAT6 gene, which is involved in fatty acid synthesis in milk [43]. The most significant and favorable associations were observed between rs110454169 and rs109913786 polymorphisms located in the AGPAT6 gene, for fat yield (0.033 kg/day), fat percentage (0.093), and rennet coagulation time (−0.462 min) [44].
FABP4 is responsible for the lipid transport protein in adipocytes, which links long-chain fatty acids and retinoic acid. The gene delivers long-chain fatty acids and retinoic acid to related receptors in the nucleus. Viale et al. [44] revealed a trend (p ≤ 0.10) of the effect of the rs110757796 polymorphism in the FABP4 gene on milk yield, protein yield, and casein yield. After adjusting for the effect of the p.K232A amino acid substitution in diacylglycerol-O-acyltransferase 1 (DGAT1), which is associated with altered levels of many milk fatty acid components, the effect of FABP4 c.328A/G on milk FA levels was generally small. After analyses of five genotypes, AB cows produced more medium–long-chain fatty acids than CC cows (p ≤ 0.05) and more C14:0 acids than AC and AA cows (p ≤ 0.05). Cows with the AC genotype had more C24:0 fatty acids (p ≤ 0.05) than BC cows. Cows with the CC genotype produced more long-chain fatty acids than cows with the BC genotype (p ≤ 0.05). AA and AC cows produced less C22:0 FAs (p ≤ 0.01) than BC cows [45].
Cell survival depends on many factors, such as division, differentiation, and migration. Fibroblast growth factor 2 (FGF2) plays an important role in these functions. Genetic variants of FGF2 have been linked to the productive life, with significant dominance effects and overall dominance, milk fat content, and somatic cell scores [46]. Li et al. [47] have shown that the SNP rs210169303 was linked to the highest 305-day milk yield. Brzáková et al. [48] assessed the effect of SNP11646 in the FGF2 gene on the regressive evidence of the breeding value of sires in terms of reproductive qualities and milk production. The difference in milk production in animals with different genotypes was negligible. Milk production in sires with the AA genotype showed the lowest DRP value. The reproductive qualities of the same bulls were highly rated both in terms of the direct genetic effect (male fertility) and the maternal genetic effect (daughter conception). According to the conception rate of daughters, in some cases, the differences reached the threshold of significance.
The nucleotide-binding oligomerization domain-containing protein 2 (NOD2), also known as Caspase Recruitment Domain 15 (CARD15), pattern recognition receptor (PRR) detects bacterial peptidoglycan fragments and other danger signals and plays an important role in gastrointestinal immunity. It is a cytosolic protein capable of initiating inflammation. The CARD15 SNPs c.3020A > T and c.4500A > C were associated with EBVs for health and production traits in Canadian Holsteins. The SNP c.3020A > T was also associated with EBVs for SCS (p = 0.0097). Hap22 (TC) was associated with increased milk (p ≤ 0.0001) and protein (p = 0.0007) yield EBVs compared to the most frequent haplotype Hap12 (AC). The hap21 (TA) was significantly associated with elevated SCS EBVs (p = 0.0120) [49]. Wang et al. [50] showed that transitions (A→T) at position 114 bp were associated with the somatic cell score (p ≤ 0.01). The G→A at position 1594 bp plays a critical role in increasing 305-day milk yields in Chinese Holstein and Chinese Simmental breeds.
Signal transducer and activator of transcription 5A (STAT5A) performs a dual function: signal transduction and transcription activation, and regulates the expression of milk proteins during lactation. Significant differences between the genotypes of the STAT5A_MslI polymorphism were revealed: cows with the TT genotype produced a milk with a higher content of fat and protein compared to cows with the TC genotype [51]. Significant relationships were found between STAT5A_AvaI genotypes and milk electrical conductivity (p  =  0.007). The greatest EC values were observed in STAT5A-AvaI-TT-genotyped animals [52]. Association testing of SNP12195 (exon 8) and SNP14217 (intron 9) showed that allele G of SNP12195 was associated with a decrease in protein and fat percentages [53].
CC-motif chemokine ligand 2 (CCL2) is a small chemokine that belongs to the CC-type chemokine family and has the ability of chemoattractant activity to recruit monocytes to sites of inflammation. CCL2 induces proliferation of MAC-T cells, a bovine mammary epithelial cell line, and enhances cell cycle progression by increasing the expression of cyclin D1 [54]. Allele C increased the yield of milk and protein. After replacing the allele, the milk yield increased by 248 kg, and the protein yield by 7.4 kg. Several significant associations were identified: CCL2 c.-95T > C with udder depth (p = 0.008), and CCL2 c.1364A > G with milk yield (p = 0.03) and protein yield (p = 0.01) [55].
The growth hormone receptor (GHR) is involved in the regulation of postnatal growth of the body. Polymorphism in the GHR gene showed an association with milk yield traits and composition in Turkish Holstein, Turkish Jersey, and Chinese Holstein cattle [56,57]. The p.Phe279Tyr mutation in the GHR was associated with the protein percentage in the Chinese dairy population (p = 0.0014) [58]. A strong association of the F279Y polymorphism with milk yield, fat, protein, and casein content was confirmed in a population of 1370 dairy cows. The influence of the 279Y allele on the increase in lactose content was shown. Substitution effects of the Tyr allele in the GHR amounted to 320 kg of milk (305 days), 0.02 kg of casein per day, and 0.07 kg lactose yields per day. The Tyr allele was associated with a lower somatic cell score (SCS) (p ≤ 0.05) [59].
Prolactin (PRL) stimulates lactation by affecting the mammary gland. A significant relationship was shown between promoter genotypes (−1043A > G and −402A > G) and milk production characteristics in Chinese Holsteins [60]. A significant difference was also found between different genotypes of the PRL gene at position A103G in the average percentage of fat (p ≤ 0.05) [61]. A meta-analysis of various published studies of the relationship between PRL_Rsa I polymorphism and milk production showed that the overall effect of the gene on milk production is 0.533, and cows with the genotype AA have higher productivity than cows with the BB genotype (p ≤ 0.01), however this applies to animals of the non-Holstein breed [62]. According to the results of the analysis, in cattle with the AB genotype compared to BB (SMD = 0.289, 95% CI 0.005, 0.573), a statistically significantly higher protein yield was revealed compared to other genotypes [63].
PLCE1 encodes the phospholipase enzyme, which catalyzes the hydrolysis of phosphatidylinositol-4,5-bisphosphate to form two second messengers: inositol-1,4,5-triphosphate (IP3) and diacylglycerol (DAG). These second messengers regulate various processes that affect cell growth, differentiation, and gene expression. A GWAS for 22 milk fatty acids in 784 Chinese Holstein cows showed associations of the ARS-BFGL-NGS-110475 in gene PLCE1 with monounsaturated and polyunsaturated fatty acid traits (MUFA and PUFA) and indices of fatty acid traits [64], BovineHD2600004009 with milk yield by multibreed genome-wide association [65], rs42816577 with subcutaneous fat deposition traits in Holstein cattle [66], and rs41624917 with fat percentage [44].
Fatty acid desaturase 1 (FADS1) is involved in lipid metabolism and polyunsaturated fatty acid biosynthesis. Beak et al. [67] investigated the rs42187261 polymorphism and found its association with low-concentration C20:4 n-6 (p = 0.044) in Hanwoo beef. FADS1 is a potential genetic marker for indices of fatty acid traits [64].
Acyl-CoA 6-desaturase (FADS2) is involved in the biosynthesis of highly unsaturated fatty acids (HUFA). It can desaturate (11E)-octadecenoate (trans-vaccenoate, a metabolite in the biohydrogenation pathway of LA and the predominant trans fatty acid in cow milk) at carbon 6, generating (6Z,11E)-octadecadienoate. FADS2 is interesting as a candidate gene for selection to increase milk production and improve resistance against mastitis [47]. The SNP FADS2_c.1571G > A is a potential genetic marker in the breeding of cattle to elevate beneficial fatty acid content in milk [68].

3.2. Candidate Genes and Genomic Regions Identified by GWAS

It is interesting to consider the functions and participation in various biological processes of genes in which mutations were identified by the results of GWAS. GWAS was used as the second approach [20,23,25].
Of the 49 mutations identified by GWAS [20,23,25], 25 were localized in intergenic regions, and most of the mutations within the genes were localized in the intron part (Figure 1). Information about genes in which SNPs have been identified by GWAS, the proteins they encode, and their functions according to UNIPROT is provided in Table S1. SNPs were localized within or in close proximity to 22 genes. In Figure 2, we demonstrate protein–protein interactions based on genes in which SNPs were identified by using GWAS.
The network built using 22 identified genes showed a two-way relationship between CLS37A1 and CSF2RB (Figure 2a). With additional genes, the number of connections increased (Figure 2b).
We performed an analysis of the occurrence of SNPs in the literature data and present information about the traits with which they were associated.
The SNP rs29019625 (BTA1, close to SLC37A1) was associated with phosphorus content [69], with the infrared wavenumber WN414 [70]. SLC37A1 is responsible for the mineral composition of milk from cows, and such data were obtained through GWAS and post-GWAS analyses [71]. SLC37A1 is located in a region from 144.38 to 145.13 Mbp, simultaneously associated with milk yield and the somatic cell score [72]. It was identified as a candidate gene by single-trait analysis for three growth stages (6, 12, and 18 months after birth) in Simmental beef cattle [73].
The TBC1D23 gene, within which rs43232419 was localized, is included in the gene contents of cattle CNV regions [74].
SNP rs110614098 was located in the intron of the gene EPHB2, which was identified as a candidate gene for the maternal effect on calving traits in cattle [75,76].
CSF2RB is of interest as a candidate gene as it is responsible for increased expression in the mammary gland of cows [77,78,79]. A study of SNPs in the CSF2RB gene showed a correlation among the − log10 p-values for milk yield QTL and co-located eQTL for CSF2RB [79].
GWAS studies by Olsen et al. [20] identified four SNPs near the CSN3 gene. Marker rs29024681 located on BTA6 was significantly associated with the protein percent (p ≤ 0.0003) [80]. During the meta-analysis, the following data were obtained: lactation yield and fat percentage showed a statistical relationship between CSN3 genotypes and these traits in additive (p ≤ 0.05) and dominant (p ≤ 0.01) genetic models [81]. Alim et al. [82] found an association of SNPs g.10944A > G, g.12703G > T, g.10985G > A, g.10993T > A, g.10888T > C, and g.10924C > A with traits of milk production. Significant effects of the CSN3 polymorphism to milk infrared spectra were found in 5 regions, where the wavenumbers were from: 1238 to 1292 cm−1, 1431 to 1477 cm−1, 1504 to 1573 cm−1, 2371 to 2607 cm−1, and 3682 to 5008 cm−1. The largest −log10(P) of 19.2 was found for wavenumber 3717 cm−1 [83]. SNP rs29024684 in the CSN3 gene was significantly associated with κ-CN (p = 505,443 × 10−59) [84]. This SNP was the highest SNP reliably identified for fat, solids, and energy in processed milk [85]; for protein percentage, rennet clotting time for samples coagulating within 45 min after enzyme addition, rennet clotting time for samples reaching 20 mm hardness within 45 min after enzyme addition, serum density 30 min after enzyme addition, rennet clotting time calculated from the clot compaction equation, potential asymptotic clot density, maximum clot density, clot compaction rate constant, and time to reach maximum clot density [86]; for α-LA [87], and for αS1-CN, α-LA, and milk protein content [88].
The SNP rs41653769 is located in the KDR gene. A strong selection signal was identified close to the KDR gene coding the coat color in the beef cattle [89] and in Sahiwal cattle [90], and KDR is also a commonly differentiated candidate gene associated with tropical adaptation in Ethiopian cattle populations [91].
Ha et al. [92] report that the BTB-01594395 mutation in the CEP162 gene was associated with Bovine viral diarrhea virus (BVDV) in Korean Holstein cattle.
The SNP rs41611219 is an intron variant of the ME1 gene. This SNP is associated with udder traits in Montbéliarde, Normande, and Holstein cows [93]. SNPs in the ME1 gene are associated with meat quality traits in Chinese Red cattle [94], and with meat and carcass quality traits in commercial Angus cattle [95].
Guanine nucleotide-binding protein G(I)/G(S)/G(O) subunit gamma-2 (GNG2), within which the SNP rs41568929 was identified [20], was associated with the fertility traits of Bos indicus [96], with retained placenta (RETP) in Canadian Holstein dairy cows [97], with the regulation of metabolic processes of hormones involved in food intake in the Holstein breed [98], and with 305-day milk yield (MY), 305-day fat yield (FY), and age at first calving (AFC) in the Thai multibreed dairy population [99].
Leal-Gutiérrez et al. [100] studied the meat quality traits from a multibreed Angus-Brahman population and found a signal for WBSF (Warner-Bratzler Shear Force) and tenderness in the ZC3H12C region.
The SNP rs29012314, located within the TMPRSS13 gene identified in Norwegian Red cattle [20], is the start SNP in large-effect QTL associated with maternal weaning weight in Red Angus cattle [101].
The SNP rs41578757 [23] is an intron mutation in the DISP1 gene. Dispatch homologue protein 1 (DISP1) has a function in hedgehog (Hh) signaling, regulates the release and extracellular accumulation of cholesterol-modified hedgehog proteins, and is, therefore, required for efficient production of the Hh signal, and it interacts with SCUBE2 to increase SHH secretion. Suggestively associated with growth in Reggiana breed haplotype was identified on BTA16:26.20–26.35 Mb (p = 1.40 × 10−7), in the DISP1 gene region [102]. DISP1 is within the window associated with the somatic cell score for Thai dairy cattle [103]. DISP1 is a candidate gene associated with the climatic covariable [104].
The RIMBP2 gene was included in the top 10 upregulated DEGs (differentially expressed genes) in Longissimus dorsi in Wagyu and Chinese Red Steppe cattle [105].
Soares [106] identified, for SCK1.1 (subclinical ketosis in the first lactation), a location on the chromosome 17 window, explaining the largest proportion of genetic variability. Genes involved in the regulation of gene expression were found in this region. Zinc finger proteins were mostly present in this region of the genome, including ZNF891.
The SNP rs41660449 is located in the intron of the gene NCOR2. NCOR2 is a possible functional candidate gene for bilateral convergent strabismus with exophthalmos [107], with candidate variants for perosomus elumbis [108]. NCOR2 is one of the overlapped genomic regions identified in at least two approaches in Valdostana Red Pied, Valdostana Black Pied, and Valdostana Chestnut populations [109].
Near the TRIM37 gene, rs110006082 is associated with lactation persistency [110]. Yu et al. [111] identified differentially expressed genes between preadipocytes and adipocytes and reported TRIM37 as a downregulated gene in differentiated adipocytes.
The SNP rs41589759 is localized within the gene JARID2. JARID2 is a gene located within the QTL which is associated with the development of the hind quarter [112], and it is located in the window from the GWAS explaining >1% of the genetic variation of the performance traits of angus cattle in high-altitude regions (elevation at 2170 m) [113]. Nyman et al. [114] point to JARID2 as a possible candidate gene for cow fertility traits.
The SNP rs29024014 in the NOL4 gene is associated with the content of C18:1t9 in milk [20]. It was shown that SNP rs109278135 near NOL4 was significantly associated with the MFP trait in Chinese Holstein [115], and with residual feed intake (RFI) in dairy cattle [116].
The SNP rs41567529, associated with the content of trans fatty acids in milk [20], is localized in the gene CUX1 (Homeobox protein cut-like 1). The role of CUX1 in hairline phenotypes makes it a strong adaptive candidate when animals are exposed to heat, cold, or toxic ergot alkaloids as a result of fescue stress [117]. This is a candidate gene identified in envGWAS multivariate analysis using continuous environmental attributes as dependent variables for Red Angus [118].
TENM3 is a candidate gene within the most significant QTL, which is associated with height or stature [119].
The SNP rs3423094014, associated with C18:2c9,t11 content in Brown Swiss [25], is localized in the DISC1 gene. Fonseca et al. [120] identified DISC1 as a prioritized candidate gene mapped within and close to (within a 200 kb interval) the haplotype associated with stillbirth events.
The gene CCDC15 (Coiled-coil domain-containing 15), in which rs42176310 was identified, is a protein-coding gene. Ryu and Lee [121] performed a genetic association of the marbling score with intragenic nucleotide variants at selection signals of the bovine genome and reported CCDC15 located at the probable selection signal. Ilska-Warner et al. [122] reported CCDC15 as a potential candidate gene for the telomere length and the association with animal fitness.
Thus, it has been shown that the genes in which SNPs associated with the content of trans fatty acids in milk were identified using GWAS were associated with various quantitative traits of dairy and beef cattle and can be used as candidate genes.

4. Conclusions

In recent years, more and more research has been carried out aimed at studying the genetic factors influencing the formation of quantitative and qualitative traits of farm animals. This article reviewed studies on the content of fatty acids in milk, primarily trans fatty acids, since their content in products is controversial. It should be noted that there have been few studies aimed at finding associations of genetic markers with the amount and composition of trans fatty acids in cow’s milk. We analyzed six articles on this topic. They were performed on Holstein, Red Norwegian, and Brown Swiss cattle breeds. Three studies were based on predetermined SNPs, an association of which with milk production had been previously identified. The remaining three experiments were performed using GWAS. The results of the studies show the presence of SNPs’ effects in the genes associated with the milk production traits in dairy cattle with the content of trans fatty acids and expand knowledge about new genomic regions of SNPs which can also affect the content of trans fatty acids in cow’s milk. At the same time, it should be noted that associations of SNPs identified by means of GWAS with various traits, including milk production, fertility, and adaptive qualities, were recorded; in some cases, such positions were under selection pressure. The material presented in this review can be used to create custom SNP panels, designed to assess their effect on phenotypic traits. Further research in this direction will provide new information that will serve as an impetus for the creation of modern breeding technologies and increase the efficiency of the production of high-quality dairy products.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/agriculture13061151/s1, Table S1: Genes in which SNPs have been identified by GWAS, the proteins they encode, and their functions, according to UNIPROT.

Author Contributions

Conceptualization, O.B. and O.K.; methodology, O.B. and O.K.; formal analysis, O.S. and O.K.; investigation O.B., O.S. and O.K.; writing—original draft preparation, O.B., O.S. and O.K.; writing—review and editing, O.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation within Project No. 22-26-00196.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, C.; Cobb, L.K.; Vesper, H.W.; Asma, S. Global Surveillance of trans-Fatty Acids. Prev. Chronic Dis. 2019, 16, E147. [Google Scholar] [CrossRef] [PubMed]
  2. Huang, L.; Federico, E.; Jones, A.; Wu, J.H.Y. Presence of trans fatty acids containing ingredients in pre-packaged foods in Australia in 2018. Aust. N. Z. J. Public Health 2020, 44, 419–420. [Google Scholar] [CrossRef] [PubMed]
  3. Dominguez, L.J.; Barbagallo, M. Chapter 3-Not All Fats Are Unhealthy. In The Prevention of Cardiovascular Disease through the Mediterranean Diet; Academic Press: Cambridge, MA, USA, 2018; pp. 35–58. [Google Scholar] [CrossRef]
  4. Vargas-Bello-Perez, E.; Garnsworthy, P.C. Trans fatty acids and their role in the milk of dairy cows. Cienc. Investig. Agrar. 2013, 40, 449–473. [Google Scholar] [CrossRef]
  5. Hanuš, O.; Samková, E.; Křížová, L.; Hasoňová, L.; Kala, R. Role of Fatty Acids in Milk Fat and the Influence of Selected Factors on Their Variability—A Review. Molecules 2018, 23, 1636. [Google Scholar] [CrossRef] [PubMed]
  6. Pipoyan, D.; Stepanyan, S.; Stepanyan, S.; Beglaryan, M.; Costantini, L.; Molinari, R.; Merendino, N. The Effect of Trans Fatty Acids on Human Health: Regulation and Consumption Patterns. Foods 2021, 10, 2452. [Google Scholar] [CrossRef]
  7. Cruz, V.A.R.; Oliveira, H.R.; Brito, L.F.; Fleming, A.; Larmer, S.; Miglior, F.; Schenkel, F.S. Genome-Wide Association Study for Milk Fatty Acids in Holstein Cattle Accounting for the DGAT1 Gene Effect. Animals 2019, 9, 997. [Google Scholar] [CrossRef]
  8. Shi, L.; Lv, X.; Liu, L.; Yang, Y.; Ma, Z.; Han, B.; Sun, D. A post-GWAS confirming effects of PRKG1 gene on milk fatty acids in a Chinese Holstein dairy population. BMC Genet. 2019, 20, 53. [Google Scholar] [CrossRef]
  9. Palombo, V.; Milanesi, M.; Sgorlon, S.; Capomaccio, S.; Mele, M.; Nicolazzi, E.; Ajmone-Marsan, P.; Pilla, F.; Stefanon, B.; D’Andrea, M. Genome-wide association study of milk fatty acid composition in Italian Simmental and Italian Holstein cows using single nucleotide polymorphism arrays. J. Dairy Sci. 2018, 101, 11004–11019. [Google Scholar] [CrossRef]
  10. Park, Y. Conjugated linoleic acid (CLA): Good or bad trans fat? J. Food Compos. Anal. 2009, 22, 4–12. [Google Scholar] [CrossRef]
  11. MacDonald, H.B. Conjugated linoleic acid and disease prevention: A review of current knowledge. J. Am. Coll. Nutr. 2000, 19, 111–118. [Google Scholar] [CrossRef]
  12. Pariza, M.W.; Park, Y.; Cook, M.E. The biologically active isomers of conjugated linoleic acid. Prog. Lipid Res. 2001, 40, 283–298. [Google Scholar] [CrossRef]
  13. Knutsen, T.M.; Olsen, H.G.; Ketto, I.A.; Sundsaasen, K.K.; Kohler, A.; Tafintseva, V.; Svendsen, M.; Kent, M.P.; Lien, S. Genetic variants associated with two major bovine milk fatty acids offer opportunities to breed for altered milk fat composition. Genet. Sel. Evol. 2022, 54, 35. [Google Scholar] [CrossRef]
  14. Yang, C.; Zhu, B.; Ye, S.; Fu, Z.; Li, J. Isomer-Specific Effects of cis-9,trans-11- and trans-10,cis-12-CLA on Immune Regulation in Ruminal Epithelial Cells. Animals 2021, 11, 1169. [Google Scholar] [CrossRef]
  15. Bessonov, V.V.; Zaitseva, L.V. Trans-isomers of fatty acids: Health risks and ways to reduce consumption. Food Issues 2016, 85, 6–17. [Google Scholar]
  16. Carroll, S.M.; DePeters, E.J.; Taylor, S.J.; Rosenberg, M.; Perez-Monti, H.; Capps, V.A. Milk composition of Holstein, Jersey, and Brown Swiss cows in response to increasing levels of dietary fat. Anim. Feed Sci. Technol. 2006, 131, 451–473. [Google Scholar] [CrossRef]
  17. Mele, M.; Macciotta, N.P.P.; Cecchinato, A.; Conte, G.; Schiavon, S.; Bittante, G. Multivariate factor analysis of detailed milk fatty acid profile: Effects of dairy system, feeding, herd, parity, and stage of lactation. J. Dairy Sci. 2016, 99, 9820–9833. [Google Scholar] [CrossRef]
  18. Maurice-Van Eijndhoven, M.H.T.; Hiemstra, S.J.; Calus, M.P.L. Short communication: Milk fat composition of 4 cattle breeds in the Netherlands. J. Dairy Sci. 2011, 94, 1021–1025. [Google Scholar] [CrossRef]
  19. Ramalho, H.M.; Campos, S.D.; Casal, S.; Alves, R.; Oliveira, M.B. Lipid fraction quality of milk produced by Minhota (Portuguese autochthonous breed) compared to Holstein Friesian cow’s. J. Sci. Food Agric. 2012, 92, 2994–3001. [Google Scholar] [CrossRef]
  20. Olsen, H.G.; Knutsen, T.M.; Kohler, A.; Svendsen, M.; Gidskehaug, L.; Grove, H.; Nome, T.; Sodeland, M.; Sundsaasen, K.K.; Kent, M.P.; et al. Genome-wide association mapping for milk fat composition and fine mapping of a QTL for de novo synthesis of milk fatty acids on bovine chromosome 13. Genet. Sel. Evol. 2017, 49, 20. [Google Scholar] [CrossRef]
  21. Ibeagha-Awemu, E.M.; Akwanji, K.A.; Beaudoin, F.; Zhao, X. Associations between variants of FADS genes and omega-3 and omega-6 milk fatty acids of Canadian Holstein cows. BMC Genet. 2014, 17, 25. [Google Scholar] [CrossRef]
  22. Bovenhuis, H.; Visker, M.H.P.W.; Poulsen, N.A.; Sehested, J.; van Valenberg, H.J.F.; van Arendonk, J.A.M.; Larsen, L.B.; Buitenhuis, A.J. Effects of the diacylglycerol o-acyltransferase 1 (DGAT1) K232A polymorphism on fatty acid, protein, and mineral composition of dairy cattle milk. J. Dairy Sci. 2016, 10, 3168. [Google Scholar] [CrossRef]
  23. Strillacci, M.G.; Frigo, E.; Canavesi, F.; Ungar, Y.; Schiavini, F.; Zaniboni, L.; Reghenzani, L.; Cozzi, M.C.; Samoré, A.B.; Kashi, Y.; et al. Quantitative trait loci mapping for conjugated linoleic acid, vaccenic acid and ∆(9) -desaturase in Italian Brown Swiss dairy cattle using selective DNA pooling. Anim. Genet. 2014, 45, 485–499. [Google Scholar] [CrossRef] [PubMed]
  24. Pegolo, S.; Cecchinato, A.; Mele, M.; Conte, G.; Schiavon, S.; Bittante, G. Effects of candidate gene polymorphisms on the detailed fatty acids profile determined by gas chromatography in bovine milk. J. Dairy Sci. 2016, 99, 4558–4573. [Google Scholar] [CrossRef] [PubMed]
  25. Pegolo, S.; Dadousis, C.; Mach, N.; Ramayo-Caldas, Y.; Mele, M.; Conte, G.; Schiavon, S.; Bittante, G.; Cecchinato, A. SNP co-association and network analyses identify E2F3, KDM5A and BACH2 as key regulators of the bovine milk fatty acid profile. Sci. Rep. 2017, 7, 17317. [Google Scholar] [CrossRef] [PubMed]
  26. Rychtářová, J.; Sztankóová, Z.; Kyselová, J.; Zink, V.; Štípková, M.; Vacek, M.; Štolc, L. Effect of DGAT1, BTN1A1, OLR1, and STAT1 genes on milk production and reproduction traits in the Czech Fleckvieh breed. Czech J. Anim. Sci. 2014, 59, 45–53. [Google Scholar] [CrossRef]
  27. Bykova, O.A.; Chechenikhina, O.S.; Stepanov, A.V.; Kostyunina, O.V.; Shevkunov, O.A.; Kosilov, V.I. A Study on Milk Productivity of Black-and-white Cows Considering Genotypes of DNA Markers CSN2, LGB, CRH, STAT1, TFAM1, and TFAM2. Int. Trans. J. Eng. Manag. Appl. Sci. Technol. 2022, 13, 1–11. [Google Scholar] [CrossRef]
  28. Akhmetov, T.M.; Safina, N.Y.; Alimov, A.M.; Varlamova, M.I. Genetic parameters of milk productivity for three lactations of Holstein cattle with different genotypes of LEP gene. BIO Web Conf. 2020, 27, 00061. [Google Scholar] [CrossRef]
  29. Metin Kiyici, J.; Akyüz, B.; Kaliber, M.; Arslan, K.; Aksel, E.G.; Çinar, M.U. LEP and SCD polymorphisms are associated with milk somatic cell count, electrical conductivity and pH values in Holstein cows. Anim. Biotechnol. 2020, 31, 498–503. [Google Scholar] [CrossRef]
  30. Maletić, M.; Paprikić, N.; Lazarević, M.; Hodžić, A.; Davidović, V.; Stanišić, L.; Stanimirović, Z. Insight in Leptin gene polymorphism and impact on milk traits in autochtonous Busha cattle. Acta Vet.-Beogr. 2019, 69, 153–163. [Google Scholar] [CrossRef]
  31. Samková, E.; Čítek, J.; Brzáková, M.; Hanuš, O.; Večerek, L.; Jozová, E.; Hoštičková, I.; Trávníček, J.; Hasoňová, L.; Rost, M.; et al. Associations among Farm, Breed, Lactation Stage and Parity, Gene Polymorphisms and the Fatty Acid Profile of Milk from Holstein, Simmental and Their Crosses. Animals 2021, 11, 3284. [Google Scholar] [CrossRef]
  32. Haruna, I.L.; Zhou, H.; Hickford, J.G.H. Variation in bovine leptin gene affects milk fatty acid composition in New Zealand Holstein Friesian × Jersey dairy cows. Arch. Anim. Breed. 2021, 64, 245–256. [Google Scholar] [CrossRef]
  33. Ladyka, V.; Drevytska, T.; Pavlenko, Y.; Skliarenko, Y.; Lahuta, T.; Drevytskyi, O.; Dosenko, V. The Comparison of CSN2 (rs43703011) Beta-Casein Gene Options Frequencies in Different Breeds of Ukraine Cows and the Prospect of Creating Herds with the A2/A2 Genotype. Res. Sq. 2021, 1–12. [Google Scholar] [CrossRef]
  34. Vanvanhossou, S.F.U.; Giambra, I.J.; Yin, T.; Brügemann, K.; Dossa, L.H.; König, S. First DNA Sequencing in Beninese Indigenous Cattle Breeds Captures New Milk Protein Variants. Genes 2021, 12, 1702. [Google Scholar] [CrossRef]
  35. Amalfitano, N.; Macedo Mota, L.F.; Rosa, G.M.; Cecchinato, A.; Bittante, G. Role of CSN2, CSN3, and BLG genes and the polygenic background in the cattle milk protein profile. J. Dairy Sci. 2022, 105, 6001–6020. [Google Scholar] [CrossRef]
  36. Sørensen, B.M.; Chris Kazala, E.; Murdoch, G.K.; Keating, A.F.; Cruz-Hernandez, C.; Wegner, J.; Kennelly, J.J.; Okine, E.K.; Weselake, R.J. Effect of CLA and other C18 unsaturated fatty acids on DGAT in bovine milk fat biosynthetic systems. Lipids 2008, 43, 903–912. [Google Scholar] [CrossRef]
  37. Kühn, C.; Thaller, G.; Winter, A.; Bininda-Emonds, O.R.; Kaupe, B.; Erhardt, G.; Bennewitz, J.; Schwerin, M.; Fries, R. Evidence for multiple alleles at the DGAT1 locus better explains a quantitative trait locus with major effect on milk fat content in cattle. Genetics 2004, 167, 1873–1881. [Google Scholar] [CrossRef]
  38. Pausch, H.; Emmerling, R.; Gredler-Grandl, B.; Fries, R.; Daetwyler, H.D.; Goddard, M.E. Meta-analysis of sequence-based association studies across three cattle breeds reveals 25 QTL for fat and protein percentages in milk at nucleotide resolution. BMC Genom. 2017, 18, 853. [Google Scholar] [CrossRef]
  39. Elzaki, S.; Korkuć, P.; Arends, D.; Reissmann, M.; Brockmann, G.A. Effects of DGAT1 on milk performance in Sudanese Butana × Holstein crossbred cattle. Trop. Anim. Health Prod. 2022, 54, 142. [Google Scholar] [CrossRef]
  40. Dudásová, S.; Miluchová, M.; Gábor, M.; Candrák, J.; Dočkalová, K. Effects of the DGAT1 K232A polymorphism on milk production traits in Holstein cattle. Acta Fytotech. Zootech. 2021, 24, 233–237. [Google Scholar] [CrossRef]
  41. Tumino, S.; Criscione, A.; Moltisanti, V.; Marletta, D.; Bordonaro, S.; Avondo, M.; Valenti, B. Feeding System Resizes the Effects of DGAT1 Polymorphism on Milk Traits and Fatty Acids Composition in Modicana Cows. Animals 2021, 11, 1616. [Google Scholar] [CrossRef]
  42. Upadhyay, A.; Chakravarty, A.K.; De, S.; Gupta, A.K.; Singh, A.; Sakthivel Selvan, A. Detection of Single-Nucleotide Polymorphism in AGPAT6 Gene, Associated with Milk Fat Content, using Tetra-Primer ARMS PCR-Based Assay, in Karan Fries Breeding Bulls. Iran J. Biotechnol. 2019, 17, e2084. [Google Scholar] [CrossRef] [PubMed]
  43. Zaalberg, R.M.; Janss, L.; Buitenhuis, A.J. Genome-wide association study on Fourier transform infrared milk spectra for two Danish dairy cattle breeds. BMC Genet. 2020, 21, 9. [Google Scholar] [CrossRef] [PubMed]
  44. Viale, E.; Tiezz, F.; Maretto, F.; De Marchi, M.; Penasa, M.; Cassandro, M. Association of candidate gene polymorphisms with milk technological traits, yield, composition, and somatic cell score in Italian Holstein-Friesian sires. J. Dairy Sci. 2017, 100, 7271–7281. [Google Scholar] [CrossRef] [PubMed]
  45. Li, Y.; Zhou, H.; Cheng, L.; Hodge, M.; Zhao, J.; Tung, R.; Edwards, G.; Hickford, J. Effects of FABP4 variation on milk fatty-acid composition for dairy cattle grazed on pasture in late lactation. J. Dairy Res. 2020, 87, 32–36. [Google Scholar] [CrossRef]
  46. Wang, X.; Maltecca, C.; Tal-Stein, R.; Lipkin, E.; Khatib, H. Association of bovine fibroblast growth factor 2 (FGF2) gene with milk fat and productive life: An example of the ability of the candidate pathway strategy to identify quantitative trait genes. J. Dairy Sci. 2008, 91, 2475–2480. [Google Scholar] [CrossRef]
  47. Li, M.; Gao, Q.; Wang, M.; Liang, Y.; Sun, Y.; Chen, Z.; Zhang, H.; Karrow, N.A.; Yang, Z.; Mao, Y. Polymorphisms in Fatty Acid Desaturase 2 Gene Are Associated with Milk Production Traits in Chinese Holstein Cows. Animals 2020, 10, 671. [Google Scholar] [CrossRef]
  48. Brzáková, M.; Hosnedlová, B.; Svitáková, A.; Vernerová, K.; Veselá, Z.; Čítek, J. Effect of the FGF2 SNP11646 on milk production and fertility traits of Holstein cattle. Czech J. Anim. Sci. 2016, 61, 377–382. [Google Scholar] [CrossRef]
  49. Pant, S.D.; Schenkel, F.S.; Leyva-Baca, I.; Sharma, B.S.; Karrow, N.A. Identification of single nucleotide polymorphisms in bovine CARD15 and their associations with health and production traits in Canadian Holsteins. BMC Genom. 2007, 8, 421. [Google Scholar] [CrossRef]
  50. Wang, W.; Cheng, L.; Yi, J.; Gan, J.; Tang, H.; Fu, M.Z.; Wang, H.; Lai, S.J. Health and production traits in bovine are associated with single nucleotide polymorphisms in the NOD2 gene. Genet. Mol. Res. 2015, 14, 3570–3578. [Google Scholar] [CrossRef]
  51. Selvaggi, M.; Albarella, S.; Dario, C.; Peretti, V.; Ciotola, F. Association of STAT5A Gene Variants with Milk Production Traits in Agerolese Cattle. Biochem. Genet. 2017, 55, 158–167. [Google Scholar] [CrossRef]
  52. Kiyici, J.M.; Akyüz, B.; Kaliber, M.; Arslan, K.; Aksel, E.G.; Cinar, M.U. Association of GH, STAT5A, MYF5 gene polymorphisms with milk somatic cell count, EC and pH levels of Holstein dairy cattle. Anim. Biotechnol. 2022, 33, 401–407. [Google Scholar] [CrossRef]
  53. Khatib, H.; Monson, R.L.; Schutzkus, V.; Kohl, D.M.; Rosa, G.J.; Rutledge, J.J. Mutations in the STAT5A gene are associated with embryonic survival and milk composition in cattle. J. Dairy Sci. 2008, 91, 784–793. [Google Scholar] [CrossRef]
  54. Yang, C.; Lim, W.; Bae, H.; Bazer, F.W.; Song, G. C-C motif chemokine ligand 2 induces proliferation and prevents lipopolysaccharide-induced inflammatory responses in bovine mammary epithelial cells. J. Dairy Sci. 2018, 101, 4527–4541. [Google Scholar] [CrossRef]
  55. Leyva-Baca, I.; Schenkel, F.; Sharma, B.S.; Jansen, G.B.; Karrow, N.A. Identification of single nucleotide polymorphisms in the bovine CCL2, IL8, CCR2 and IL8RA genes and their association with health and production in Canadian Holsteins. Anim. Genet. 2007, 38, 198–202. [Google Scholar] [CrossRef]
  56. Cobanoglu, O.; Kul, E.; Gurcan, E.K.; Abaci, S.H.; Cankaya, S. Determination of the association of GHR/AluI gene polymorphisms with milk yield traits in Holstein and Jersey cattle raised in Turkey. Arch. Anim. Breed. 2021, 64, 417–424. [Google Scholar] [CrossRef]
  57. Ma, Y.N.; He, P.J.; Zhu, J.; Lei, Z.M.; Liu, Z.; Wu, J.P. The effect of polymorphism F279Y of GHR gene on milk production trait in Chinese Holstein cattle. Zhongguo Ying Yong Sheng Li Xue Za Zhi 2013, 29, 400–404. (In Chinese) [Google Scholar]
  58. Sun, D.; Jia, J.; Ma, Y.; Zhang, Y.; Wang, Y.; Yu, Y.; Zhang, Y. Effects of DGAT1 and GHR on milk yield and milk composition in the Chinese dairy population. Anim. Genet. 2009, 40, 997–1000. [Google Scholar] [CrossRef]
  59. Rahmatalla, S.A.; Müller, U.; Strucken, E.M.; Reissmann, M.; Brockmann, G.A. The F279Y polymorphism of the GHR gene and its relation to milk production and somatic cell score in German Holstein dairy cattle. J. Appl. Genet. 2011, 52, 459–465. [Google Scholar] [CrossRef]
  60. Lü, A.; Hu, X.; Chen, H.; Jiang, J.; Zhang, C.; Xu, H.; Gao, X. Single nucleotide polymorphisms in bovine PRL gene and their associations with milk production traits in Chinese Holsteins. Mol. Biol. Rep. 2010, 37, 547–551. [Google Scholar] [CrossRef]
  61. Patel, J.B.; Chauhan, J.B. Polymorphism of the Prolactin Gene and Its Relationship with Milk Production in Gir and Kankrej Cattle. J. Nat. Sci. Biol. Med. 2017, 8, 167–170. [Google Scholar] [CrossRef]
  62. Alipanah, M.; Rodbari, Z.; Esmailizadeh, A.; Javan, I.Y.; Qarari, F. The relationships between PRL/Rsa Ipolymorphism in prolactin gene and milk production in cattle: A Meta-analysis. Large Anim. Rev. 2023, 29, 9–14. [Google Scholar]
  63. Motmain, Z.; Özdemir, M.; Ekinci, K.; Saygili, E.; Bilgin, E. A Meta-Analysis of the Associations Between Prolactin (PRL) Gene Polymorphism and Milk Production Traits in Cattle. Kafkas Univ. Vet. Fak. Derg. 2022, 28, 627–631. [Google Scholar] [CrossRef]
  64. Li, C.; Sun, D.; Zhang, S.; Wang, S.; Wu, X.; Zhang, Q.; Liu, L.; Li, Y.; Qiao, L. Genome Wide Association Study Identifies 20 Novel Promising Genes Associated with Milk Fatty Acid Traits in Chinese Holstein. PLoS ONE 2014, 9, e96186. [Google Scholar] [CrossRef]
  65. Raven, L.A.; Cocks, B.G.; Hayes, B.J. Multibreed genome wide association can improve precision of mapping causative variants underlying milk production in dairy cattle. BMC Genom. 2014, 15, 62. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, H.; Mi, S.; Brito, L.F.; Hu, L.; Wang, L.; Ma, L.; Xu, Q.; Guo, G.; Yu, Y.; Wang, Y. Genomic and transcriptomic analyses enable the identification of important genes associated with subcutaneous fat deposition in Holstein cows. J. Genet. Genom. 2023; in press. [Google Scholar] [CrossRef]
  67. Beak, S.H.; Lee, Y.; Lee, E.B.; Kim, K.H.; Kim, J.G.; Bok, J.D.; Kang, S.K. Study on the fatty acid profile of phospholipid and neutral lipid in Hanwoo beef and their relationship to genetic variation. J. Anim. Sci. Technol. 2019, 61, 69–76. [Google Scholar] [CrossRef]
  68. Li, M.; Lu, X.; Gao, Q.; Wang, M.; Arbab, A.A.I.; Sun, Y.; Chen, Z.; Zhang, H.; Karrow, N.A.; Yang, Z.; et al. A Functional 3′ UTR Polymorphism of FADS2 Affects Cow Milk Composition through Modifying Mir-744 Binding. Animals 2019, 9, 1090. [Google Scholar] [CrossRef]
  69. Bovenhuis, H.; Jibrila, I.; Dijkstra, J. Predicting milk phosphorus content based on genotypic and milk infrared data. In Proceedings of the World Congress on Genetics Applied to Livestock Production, Auckland, New Zealand, 7–11 February 2018; p. 534. Available online: http://www.wcgalp.org/proceedings/2018/predicting-milk-phosphorus-content-based-genotypic-and-milk-infrared-data (accessed on 15 April 2023).
  70. Wang, Q.; Bovenhuis, H. Genome-wide association study for milk infrared wavenumbers. J. Dairy Sci. 2018, 101, 2260–2272. [Google Scholar] [CrossRef]
  71. Sanchez, M.P.; Rocha, D.; Charles, M.; Boussaha, M.; Hozé, C.; Brochard, M.; Delacroix-Buchet, A.; Grosperrin, P.; Boichard, D. 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]
  72. Iung, L.H.S.; Petrini, J.; Ramírez-Díaz, J.; Salvian, M.; Rovadoscki, G.A.; Pilonetto, F.; Dauria, B.D.; Machado, P.F.; Coutinho, L.L.; Wiggans, G.R.; et al. Genome-wide association study for milk production traits in a Brazilian Holstein population. J. Dairy Sci. 2019, 102, 5305–5314. [Google Scholar] [CrossRef]
  73. An, B.; Xu, L.; Xia, J.; Wang, X.; Miao, J.; Chang, T.; Song, M.; Ni, J.; Xu, L.; Zhang, L.; et al. Multiple association analysis of loci and candidate genes that regulate body size at three growth stages in Simmental beef cattle. BMC Genet. 2020, 21, 32. [Google Scholar] [CrossRef]
  74. Hou, Y.; Liu, G.E.; Bickhart, D.M.; Cardone, M.F.; Wang, K.; Kim, E.S.; Matukumalli, L.K.; Ventura, M.; Song, J.; VanRaden, P.M.; et al. Genomic characteristics of cattle copy number variations. BMC Genom. 2011, 12, 127. [Google Scholar] [CrossRef]
  75. 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]
  76. Chen, Z.; Brito, L.F.; Luo, H.; Shi, R.; Chang, Y.; Liu, L.; Guo, G.; Wang, Y. Genetic and Genomic Analyses of Service Sire Effect on Female Reproductive Traits in Holstein Cattle. Front. Genet. 2021, 3, 713575. [Google Scholar] [CrossRef]
  77. Raven, L.A.; Cocks, B.G.; Kemper, K.E.; Chamberlain, A.J.; Vander Jagt, C.J.; Goddard, M.E.; Hayes, B.J. Targeted imputation of sequence variants and gene expression profiling identifies twelve candidate genes associated with lactation volume, composition and calving interval in dairy cattle. Mamm. Genome 2016, 27, 81–97. [Google Scholar] [CrossRef]
  78. Kemper, K.E.; Hayes, B.J.; Daetwyler, H.D.; Goddard, M.E. How old are quantitative trait loci and how widely do they segregate? J. Anim. Breed. Genet. 2015, 132, 121–134. [Google Scholar] [CrossRef]
  79. Lopdell, T.J.; Tiplady, K.; Couldrey, C.; Johnson, T.J.J.; Keehan, M.; Davis, S.R.; Harris, B.L.; Spelman, R.J.; Snell, R.G.; Littlejohn, M.D. Multiple QTL underlie milk phenotypes at the CSF2RB locus. Genet. Sel. Evol. 2019, 51, 3. [Google Scholar] [CrossRef]
  80. Bierman, C.D.; Kim, E.; Shi, X.W.; Weigel, K.; Jeffrey Berger, P.; Kirkpatrick, B.W. Validation of whole genome linkage-linkage disequilibrium and association results, and identification of markers to predict genetic merit for twinning. Anim. Genet. 2010, 41, 406–416. [Google Scholar] [CrossRef]
  81. Mahmoudi, P.; Rostamzadeh, J.; Rashidi, A.; Zergani, E.; Razmkabir, M. A meta-analysis on association between CSN3 gene variants and milk yield and composition in cattle. Anim. Genet. 2020, 51, 369–381. [Google Scholar] [CrossRef]
  82. Alim, M.A.; Dong, T.; Xie, Y.; Wu, X.P.; Zhang, Y.; Zhang, S.; Sun, D.X. Effect of polymorphisms in the CSN3 (κ-casein) gene on milk production traits in Chinese Holstein Cattle. Mol. Biol. Rep. 2014, 41, 7585–7593. [Google Scholar] [CrossRef]
  83. Wang, Q.; Hulzebosch, A.; Bovenhuis, H. Genetic and environmental variation in bovine milk infrared spectra. J. Dairy Sci. 2016, 99, 6793–6803. [Google Scholar] [CrossRef]
  84. Pegolo, S.; Mach, N.; Ramayo-Caldas, Y.; Schiavon, S.; Bittante, G.; Cecchinato, A. Integration of GWAS, pathway and network analyses reveals novel mechanistic insights into the synthesis of milk proteins in dairy cows. Sci. Rep. 2018, 8, 566. [Google Scholar] [CrossRef] [PubMed]
  85. Dadousis, C.; Biffani, S.; Cipolat-Gotet, C.; Nicolazzi, E.L.; Rosa, G.J.M.; Gianola, D.; Rossoni, A.; Santus, E.; Bittante, G.; Cecchinato, A. Genome-wide association study for cheese yield and curd nutrient recovery in dairy cows. J. Dairy Sci. 2017, 100, 1259–1271. [Google Scholar] [CrossRef]
  86. Dadousis, C.; Biffani, S.; Cipolat-Gotet, C.; Nicolazzi, E.L.; Rossoni, A.; Santus, E.; Bittante, G.; Cecchinato, A. Genome-wide association of coagulation properties, curd firmness modeling, protein percentage, and acidity in milk from Brown Swiss cows. J. Dairy Sci. 2016, 99, 3654–3666. [Google Scholar] [CrossRef] [PubMed]
  87. Schopen, G.C.B. Genetic Analysis of Protein Composition of Bovine Milk; Wageningen University: Wageningen, The Netherlands, 2010; 189p, ISBN 9789085856450. [Google Scholar]
  88. Moravčíková, N.; Kasarda, R.; Kadlečík, O. Genetic improvement of cattle through low density SNP panels. In Proceedings of the VIII International Scientific Agriculture Symposium “V AgroSym.”, Jahorina, Bosnia and Herzegovina, 5–8 October 2017; Book of Proceedings. Volume 41, pp. 2212–2219. [Google Scholar]
  89. Moravčíková, N.; Kasarda, R.; Vostrý, L.; Krupová, Z.; Krupa, E.; Lehocká, K.; Olšanská, B.; Trakovická, A.; Nádaský, R.; Židek, R.; et al. Analysis of selection signatures in the beef cattle genome. Czech J. Anim. Sci. 2019, 64, 491–503. [Google Scholar] [CrossRef]
  90. Illa, S.K.; Mukherjee, S.; Nath, S.; Mukherjee, A. Genome-Wide Scanning for Signatures of Selection Revealed the Putative Genomic Regions and Candidate Genes Controlling Milk Composition and Coat Color Traits in Sahiwal Cattle. Front. Genet. 2021, 12, 699422. [Google Scholar] [CrossRef]
  91. Mengistie, D.; Edea, Z.; Tesema, T.; Dejene, G.; Dessie, T.; Jemal, J.; Kim, K.S.; Dadi, H. Genome-Wide Signature of Positive Selection and Linkage Disequilibrium in Ethiopian Indigenous and European Beef Cattle Breeds. Res. Sq. 2022, 1–14. [Google Scholar] [CrossRef]
  92. Ha, S.; Lee, D.; Lee, S.; Chae, J.; Seo, K. Genome-wide association study on immune-response for improving healthiness in Holstein dairy cattle. Korean J. Vet. Serv. 2019, 42, 217–225. [Google Scholar] [CrossRef]
  93. Tribout, T.; Croiseau, P.; Lefebvre, R.; Barbat, A.; Boussaha, M.; Fritz, S.; Boichard, D.; Hoze, C.; Sanchez, M.P. Confirmed effects of candidate variants for milk production, udder health, and udder morphology in dairy cattle. Genet. Sel. Evol. 2020, 52, 55. [Google Scholar] [CrossRef]
  94. Zhou, G.L.; Cao, Y.; Jin, H.G. Identification of SNPs in ME1 gene and association analysis with meat quality traits in Chinese Red cattle. Czech J. Anim. Sci. 2014, 59, 297–301. [Google Scholar] [CrossRef]
  95. Gill, J.L.; Bishop, S.C.; McCorquodale, C.; Williams, J.L.; Wiener, P. Identification of polymorphisms in the malic enzyme 1, NADP(+)-dependent, cytosolic and nuclear receptor subfamily 0, group B, member 2 genes and their associations with meat and carcass quality traits in commercial Angus cattle. Anim. Genet. 2012, 43, 88–92. [Google Scholar] [CrossRef]
  96. Tahir, M.S.; Porto-Neto, L.R.; Gondro, C.; Shittu, O.B.; Wockner, K.; Tan, A.W.L.; Smith, H.R.; Gouveia, G.C.; Kour, J.; Fortes, M.R.S. Meta-Analysis of Heifer Traits Identified Reproductive Pathways in Bos indicus Cattle. Genes 2021, 12, 768. [Google Scholar] [CrossRef]
  97. Guarini, A.R.; Lourenco, D.A.L.; Brito, L.F.; Sargolzaei, M.; Baes, C.F.; Miglior, F.; Misztal, I.; Schenkel, F.S. Genetics and genomics of reproductive disorders in Canadian Holstein cattle. J. Dairy Sci. 2019, 102, 1341–1353. [Google Scholar] [CrossRef]
  98. Salleh, M.S.; Mazzoni, G.; Hoglund, J.K.; Olijhoek, D.W.; Lund, P.; Lovendahl, P.; Kadarmideen, H.N. RNA-Seq transcriptomics and pathway analyses reveal potential regulatory genes and molecular mechanisms in high- and low-residual feed intake in Nordic dairy cattle. BMC Genom. 2017, 18, 258. [Google Scholar] [CrossRef]
  99. Laodim, T.; Elzo, M.A.; Koonawootrittriron, S.; Suwanasopee, T.; Jattawa, D. Pathway enrichment and protein interaction network analysis for milk yield, fat yield and age at first calving in a Thai multibreed dairy population. Asian-Australas. J. Anim. Sci. 2019, 32, 508–518. [Google Scholar] [CrossRef]
  100. Leal-Gutiérrez, J.D.; Elzo, M.A.; Carr, C.; Mateescu, R.G. RNA-seq analysis identifies cytoskeletal structural genes and pathways for meat quality in beef. PLoS ONE 2020, 15, e0240895. [Google Scholar] [CrossRef]
  101. Saatchi, M.; Schnabel, R.D.; Taylor, J.F.; Garrick, D.J. Large-effect pleiotropic or closely linked QTL segregate within and across ten US cattle breeds. BMC Genom. 2014, 15, 442. [Google Scholar] [CrossRef]
  102. Bovo, S.; Schiavo, G.; Kazemi, H.; Moscatelli, G.; Ribani, A.; Ballan, M.; Bonacini, M.; Prandi, M.; Dall’Olio, S.; Fontanesi, L. Exploiting within-breed variability in the autochthonous Reggiana breed identified several candidate genes affecting pigmentation-related traits, stature and udder defects in cattle. Anim. Genet. 2021, 52, 579–597. [Google Scholar] [CrossRef]
  103. Buaban, S.; Lengnudum, K.; Boonkum, W.; Phakdeedindan, P. Genome-wide association study on milk production and somatic cell score for Thai dairy cattle using weighted single-step approach with random regression test-day model. J. Dairy Sci. 2022, 105, 468–494. [Google Scholar] [CrossRef]
  104. Flori, L.; Fritz, S.; Jaffrézic, F.; Boussaha, M.; Gut, I.; Heath, S.; Foulley, J.L.; Gautier, M. The genome response to artificial selection: A case study in dairy cattle. PLoS ONE 2009, 4, e6595. [Google Scholar] [CrossRef]
  105. Li, G.; Yang, R.; Lu, X.; Liu, Y.; He, W.; Li, Y.; Yu, H.; Qin, L.; Cao, Y.; Zhao, Z.; et al. RNA-Seq Analysis Identifies Differentially Expressed Genes in the Longissimus dorsi of Wagyu and Chinese Red Steppe Cattle. Int. J. Mol. Sci. 2022, 24, 387. [Google Scholar] [CrossRef]
  106. Soares, R. Investigation of Genomic Regions Affecting Ketosis in Dairy Cattle. Master’s Thesis, The University of Guelph, Guelph, ON, Canada, 2020; 211p. Available online: https://hdl.handle.net/10214/21280 (accessed on 15 April 2023).
  107. Bögeholz, A.; Falker-Gieske, C.; Guélat, M.; Gurtner, C.; Hunziker, S.; Oevermann, A.; Thaller, G.; Drögemüller, C.; Tetens, J. GWAS Hits for Bilateral Convergent Strabismus with Exophthalmos in Holstein Cattle Using Imputed Sequence Level Genotypes. Genes 2021, 12, 1039. [Google Scholar] [CrossRef] [PubMed]
  108. Barber, A.M.; Helms, A.; Thompson, R.; Whitlock, B.K.; Steffen, D.J.; Petersen, J.L. Whole-genome sequencing to investigate a possible genetic basis of perosomus elumbis in a calf resulting from a consanguineous mating. Transl. Anim. Sci. 2021, 5, S1–S5. [Google Scholar] [CrossRef]
  109. Mastrangelo, S.; Ben Jemaa, S.; Ciani, E.; Sottile, G.; Moscarelli, A.; Boussaha, M.; Montedoro, M.; Pilla, F.; Cassandro, M. Genome-wide detection of signatures of selection in three Valdostana cattle populations. J. Anim. Breed. Genet. 2020, 137, 609–621. [Google Scholar] [CrossRef] [PubMed]
  110. Do, D.N.; Bissonnette, N.; Lacasse, P.; Miglior, F.; Sargolzaei, M.; Zhao, X.; Ibeagha-Awemu, E.M. Genome-wide association analysis and pathways enrichment for lactation persistency in Canadian Holstein cattle. J. Dairy Sci. 2017, 100, 1955–1970. [Google Scholar] [CrossRef]
  111. Yu, S.-L.; Lee, S.-M.; Kang, M.-J.; Jeong, H.-J.; Sang, B.-C.; Jeon, J.-T.; Lee, J.-H. Identification of Differentially Expressed Genes between Preadipocytes and Adipocytes Using Affymetrix Bovine Genome Array. J. Anim. Sci. Technol. 2009, 51, 443–452. [Google Scholar] [CrossRef]
  112. Doyle, J.L.; Berry, D.P.; Veerkamp, R.F.; Carthy, T.R.; Evans, R.D.; Walsh, S.W.; Purfield, D.C. Genomic regions associated with muscularity in beef cattle differ in five contrasting cattle breeds. Genet. Sel. Evol. 2020, 52, 2. [Google Scholar] [CrossRef]
  113. Zeng, X. Angus Cattle at High Altitude: Pulmonary Arterial Pressure, Estimated Breeding Value and Genome-Wide Association Study. Ph.D. Thesis, Colorado State University, Fort Collins, CO, USA, 2016; 259p. Available online: https://mountainscholar.org/bitstream/handle/10217/178953/ZENG_colostate_0053A_13997.pdf (accessed on 19 April 2023).
  114. Nyman, S.; Duchemin, S.I.; de Koning, D.J.; Berglund, B. Genome-wide association study of normal and atypical progesterone profiles in Holstein-Friesian dairy cows. J. Dairy Sci. 2019, 102, 3204–3215. [Google Scholar] [CrossRef]
  115. Lu, X.; Arbab, A.A.I.; Abdalla, I.M.; Liu, D.; Zhang, Z.; Xu, T.; Su, G.; Yang, Z. Genetic Parameter Estimation and Genome-Wide Association Study-Based Loci Identification of Milk-Related Traits in Chinese Holstein. Front. Genet. 2022, 12, 799664. [Google Scholar] [CrossRef]
  116. Houlahan, K.L. Understanding the Genomic Architecture of Feed Efficiency and Implications of Selection for It in Dairy Cattle. Ph.D. Thesis, The University of Guelph, Guelph, ON, Canada, 2021; 137p. [Google Scholar]
  117. Aiken, G.E.; Klotz, J.L.; Looper, M.L.; Tabler, S.F.; Schrick, F.N. Disrupted hair follicle activity in cattle grazing endophyte-infected tall fescue in the summer insulates core body temperatures. Prof. Anim. Sci. 2011, 27, 336–343. [Google Scholar] [CrossRef]
  118. Rowan, T.N.; Durbin, H.J.; Seabury, C.M.; Schnabel, R.D.; Decker, J.E. Powerful detection of polygenic selection and evidence of environmental adaptation in US beef cattle. PLoS Genet. 2021, 17, e1009652. [Google Scholar] [CrossRef]
  119. Doyle, J.L.; Berry, D.P.; Veerkamp, R.F.; Carthy, T.R.; Walsh, S.W.; Evans, R.D.; Purfield, D.C. Genomic Regions Associated with Skeletal Type Traits in Beef and Dairy Cattle Are Common to Regions Associated with Carcass Traits, Feed Intake and Calving Difficulty. Front. Genet. 2020, 11, 20. [Google Scholar] [CrossRef]
  120. Fonseca, P.A.S.; Schenkel, F.S.; Cánovas, A. Genome-wide association study using haplotype libraries and repeated-measures model to identify candidate genomic regions for stillbirth in Holstein cattle. J. Dairy Sci. 2022, 105, 1314–1326. [Google Scholar] [CrossRef]
  121. Ryu, J.; Lee, C. Genetic association of marbling score with intragenic nucleotide variants at selection signals of the bovine genome. Animal 2016, 10, 566–570. [Google Scholar] [CrossRef]
  122. Ilska-Warner, J.J.; Psifidi, A.; Seeker, L.A.; Wilbourn, R.V.; Underwood, S.L.; Fairlie, J.; Whitelaw, B.; Nussey, D.H.; Coffey, M.P.; Banos, G. The Genetic Architecture of Bovine Telomere Length in Early Life and Association with Animal Fitness. Front. Genet. 2019, 10, 1048. [Google Scholar] [CrossRef]
Figure 1. Summary statistics for the identified SNPs.
Figure 1. Summary statistics for the identified SNPs.
Agriculture 13 01151 g001
Figure 2. Protein–protein interactions constructed on genes in which SNPs were identified by using GWAS (STRING db). (a) Network built using 22 identified genes and (b) network built with added genes.
Figure 2. Protein–protein interactions constructed on genes in which SNPs were identified by using GWAS (STRING db). (a) Network built using 22 identified genes and (b) network built with added genes.
Agriculture 13 01151 g002
Table 1. SNPs associated with trans fat content in cow’s milk presented in the Cattle QTLdb (genome version ARS-UCD1.2).
Table 1. SNPs associated with trans fat content in cow’s milk presented in the Cattle QTLdb (genome version ARS-UCD1.2).
No.TraitsSNPBreedBTAPositionConsequenceCandidate GeneAlleleReferences
1C18:1t11rs41569243Norwegian Red cattle145,564,535intergenic variant T/A[20]
2C18:1t11rs29019625Norwegian Red cattle1142,894,021intergenic variantSLC37A1A/G[20]
3C18:2t11,c15rs43232419Brown Swiss144,571,298intron variantTBC1D23T/G[25]
4C18:1t11rs110831311Brown Swiss1147,265,560intergenic variant A/G[25]
5C18:1t16, C18:1t6–8rs43705173Brown Swiss279,518,124downstream gene variantSTAT1G/A[24]
6C18:1t16rs43706906Brown Swiss279,553,228intron variantSTAT1G/C[24]
7C18:2c9,t11rs110614098Brown Swiss2129,910,992intron variantEPHB2G/A[25]
8C18:1t11rs110847444Brown Swiss488,230,965intergenic variant T/C[23]
9C18:2c9,t11, C18:2t11,c15rs29004170Brown Swiss492,436,328upstream gene variantLEPG/C[24]
10MTCFARrs41590819Norwegian Red cattle575,374,050downstream gene variantCSF2RBC/G[20]
11C18:1t10rs41623140Norwegian Red cattle592,068,616intergenic variant T/G[20]
12C18:1t10rs43703011Brown Swiss685,451,298missense variantCSN2T/G[24]
13C18:2c9,t11rs43703013Brown Swiss685,451,132missense variantCSN2C/G[24]
14C18:1t9rs29024681Norwegian Red cattle685,662,337upstream/downstream gene variantCSN3G/A[20]
15C18:1t9rs29024683Norwegian Red cattle685,662,408upstream/downstream gene variantCSN3A/G[20]
16C18:1t9rs29024684Norwegian Red cattle685,662,466upstream/downstream gene variantCSN3C/A[20]
17C18:1t9rs29024685Norwegian Red cattle685,662,516upstream/downstream gene variantCSN3A/G[20]
18C18:1t9rs41587868Norwegian Red cattle672,916,033intergenic variant G/A[20]
19C18:1t9rs41653769Norwegian Red cattle670,588,722intron variantKDRA/G[20]
20C18:1t16rs42960052Brown Swiss620,546,001intergenic variant A/G[25]
21C18:1t6–8rs41651324Brown Swiss676938995intergenic variant C/T[25]
22C18:2t11,c15rs41567758Brown Swiss669,216,812intergenic variant A/C[23]
23C18:1t16rs41573488Brown Swiss83,790,608intergenic variant T/C[25]
24C18:1t10rs41593428Norwegian Red cattle965,223,783intron variantCEP162C/G[20]
25C18:1t10rs41611219Norwegian Red cattle923,244,267intron variantME1C/A[20]
26TFArs108973184Brown Swiss9102,953,472intergenic variant T/C[25]
27C18:1t10rs41568929Norwegian Red cattle1044,720,992intron variantGNG2T/C[18]
28C18:1t11rs41575963Brown Swiss1226,784,974intergenic variant A/C[21]
29MTRANSFArs41662646Norwegian Red cattle1277,525,632intergenic variant G/A[20]
30C18:1t6–8rs109738802Brown Swiss1278,577,324intergenic variant C/T[25]
31C18:1c9,t11rs109326954Jersey, Holstein14611,020missense variantDGAT1C/A[22]
32C18:1t9rs110445169Brown swiss1420,267,955intron variantAGPAT6T/C[24]
33C16:1t9, C18:1t4rs110757796Brown swiss1444,677,959missense variantFABP4T/C[24]
34C18:1T4rs110642420Brown Swiss145,880,036intergenic variant C/T[25]
35C18:1t11rs42357017Brown Swiss1520,213,507intron variantZC3H12CC/T[23]
36MTCFARrs29012314Norwegian Red cattle1528,602,093upstream gene variantTMPRSS13C/T[20]
37MTCFARrs29019684Norwegian Red cattle1557,539,478intergenic variant G/A[20]
38C18:1t10rs41655008Brown Swiss1569,449,535intergenic variant C/T[25]
39C18:1t11rs41578757Brown Swiss1626,461,264intron variantDISP1T/A[23]
40C18:1t11rs109178989Brown Swiss1725,016,349intergenic variant A/G[23]
41C18:1T6–8rs110937773Brown Swiss1734,849,206intron variantFGF2T/C[24]
42MTCFARrs41633197Norwegian Red cattle1746,687,151intron variantRIMBP2A/G[20]
43MTCFARrs41637627Norwegian Red cattle1744,150,664upstream gene variantZNF891G/T[20]
44MTCFARrs41660449Norwegian Red cattle1751,342,347intron variantNCOR2T/G[20]
45C18:1T4rs41596865Brown Swiss1759,367,351intergenic variant C/T[25]
46C18:1T11rs41611446Brown Swiss1720,279,742intergenic variant G/A[25]
47C18:1t16rs43710288Brown Swiss1819,117,794missense variantNOD2A/T[24]
48C18:1T4rs109578101Brown Swiss1942,415,6823 prime UTR variantSTAT5AC/T[24]
49C18:3rs137182814Brown Swiss1942,335,251splice region variant, synonymous variantSTAT5AG/C[24]
50TFA, С18:1t11rs110773010Brown Swiss1910,076,852intron variantTRIM37G/A[25]
51C18:1t4rs41255713Brown Swiss1915,905,377upstream gene variantCCL2G/A[24]
52C18:1t10rs109136815Brown swiss2031,870,046synonymous variantGHRA/G[24]
53C18:3c9,t11,c15rs109794490Brown Swiss2221,455,728intergenic variant C/T[25]
54C18:2t9,t12rs211032652Brown Swiss2335,333,764synonymous variantPRLC/T[24]
55C18:1t10rs41589759Norwegian Red cattle2341,632,536intron variantJARID2G/A[20]
56MTCFARrs41642031Norwegian Red cattle2323,196,063intergenic variant G/C[20]
57C18:1t16rs41641235Brown Swiss2322,032,364intergenic variant T/G[25]
58C18:1t9rs29024014Norwegian Red cattle2423,015,448intron variantNOL4A/G[20]
59C18:1t9rs41644943Norwegian Red cattle246,108,284intergenic variant G/A[20]
60MTRANSFArs41567529Norwegian Red cattle2534,827,662intron variantCUX1T/C[20]
61C18:2c9,t11rs41622946Brown Swiss2532,138,950intergenic variant T/C[25]
62C18:3c9,t11,c15rs41624917Brown Swiss2615,524,965intron variantPLCE1G/A[24]
63C16:1t9rs41650170Brown Swiss2713,628,042intron variantTENM3G/A[25]
64C18:2c9,t11rs3423094014Brown Swiss284,467,528intron variantDISC1G/T[25]
65C18:2t6rs42187261Holstein2940,247,432synonymous variantFADS1G/A[21]
66C18:2t10, C18:2t6rs210169303Holstein2940,386,1733 prime UTR variantFADS2G/A[21]
67C14:1trs109772589Holstein2940,387,3453 prime UTR variantFADS2G/A[21]
68C18:1t11rs42176310Brown Swiss2928,440,651intron variantCCDC15A/C[23]
Grey rows—SNPs identified by using GWAS.
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

Bykova, O.; Shevkunov, O.; Kostyunina, O. Overview of SNPs Associated with Trans Fat Content in Cow’s Milk. Agriculture 2023, 13, 1151. https://doi.org/10.3390/agriculture13061151

AMA Style

Bykova O, Shevkunov O, Kostyunina O. Overview of SNPs Associated with Trans Fat Content in Cow’s Milk. Agriculture. 2023; 13(6):1151. https://doi.org/10.3390/agriculture13061151

Chicago/Turabian Style

Bykova, Olga, Oleg Shevkunov, and Olga Kostyunina. 2023. "Overview of SNPs Associated with Trans Fat Content in Cow’s Milk" Agriculture 13, no. 6: 1151. https://doi.org/10.3390/agriculture13061151

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