Next Article in Journal
Assessing Molecular Mechanisms of Stress Induced Salinity Adaptation in the Juvenile Ornate Spiny Lobster, Panulirus ornatus
Previous Article in Journal
Lactobacillus rhamnosus GG and Lactobacillus paracasei IMPC2.1 Mitigate LPS-Induced Epithelial Barrier Dysfunction: A Focus on Autophagy Regulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Association Study of Daughter Pregnancy Rate in Crossbred Dairy Cows

1
Department of Animal Science, University of Minnesota, Saint Paul, MN 55108, USA
2
College of Animal Science and Technology, Yunnan Agricultural University, Kunming 650201, China
3
Department of Epidemiology and Health Statistics, School of Public Health, Fujian Medical University, Fuzhou 350108, China
4
Department of Animal and Avian Sciences, University of Maryland, College Park, MD 20742, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(22), 11149; https://doi.org/10.3390/ijms262211149
Submission received: 8 September 2025 / Revised: 11 November 2025 / Accepted: 14 November 2025 / Published: 18 November 2025
(This article belongs to the Section Molecular Genetics and Genomics)

Abstract

A genome-wide association study (GWAS) of daughter pregnancy rate (DPR) was conducted using 75,133 SNPs and 40,203 first lactation crossbred dairy cows mostly from Jersey–Holstein crosses. The GWAS analysis detected 6528 additive effects, 65 dominance effects, 1638 additive × additive (A × A) effects, 3 additive × dominance effects, and 18 intra-chromosome dominance × dominance (D × D) effects. Of the 1638 A × A effects, 1634 were intra-chromosome and four were inter-chromosome A × A effects. The distance between two SNPs with intra-chromosome epistasis effects was in the range of 3.61 Kb to 2.68 Mb, and many interacting SNP pairs were within the same genes. The additive and A × A effects were distributed on all chromosomes showing genome-wide involvement in DPR heterosis. The dominance and D × D effects all had homozygous advantages and heterozygous disadvantages. The GWAS results identified four genetic mechanisms underlying DPR heterosis in crossbred dairy cows: complementary additive effects from different breeds and new additive effects due to cross breeding, two-locus allelic interactions between loci and between breeds, within-locus allelic interactions between breeds, and genotype × genotype interactions enabled by allelic interactions between breeds. Results in this study provided a novel understanding about the genetic factors and mechanisms underlying DPR heterosis in crossbred dairy cows.

1. Introduction

Daughter pregnancy rate (DPR) is the percentage of cows that become pregnant during each 21-day period and is an important fertility trait in dairy cows [1]. The Holstein breed is the largest dairy breed in the U.S. with high production levels but low DPR, which was in decline since 1960s until about 2000 when a trend reversal occurred, but the declines since the 1960s were not fully recovered [2]. In contrast to purebred dairy cows, crossbred dairy cows typically have improved fertility and survival rate relative to pure breeds and these advantages are most striking relative to Holsteins [3,4,5,6]. The phenotypic advantage of crossbreds over the phenotypic performance of their parental breeds is termed as heterosis [7,8]. The previously reported DPR heterosis was also confirmed in this study (see Section 3). Several data-driven hypotheses about the genetic mechanism of heterosis have been proposed, including more superior alleles in hybrids than in parental lines [9]; heterozygous advantage where heterozygous genotypes outperformed homozygous genotypes; a genetic phenomenon termed as dominance [10] or overdominance [11,12]; the dominance model where the genotypes with one or two dominant alleles per locus were more favorable than the homozygous recessive genotypes [13]; and negative dominance and dominance × dominance epistasis effects [14]. These different hypotheses indicated that the genetic mechanism of heterosis could be different for different traits and species and that more studies on the genetic mechanism of heterosis are needed to build a consensus towards understanding the genetic mechanism of heterosis. The U.S. dairy genomic evaluation database has accumulated a large dairy crossbred population with DPR phenotypic observations and genotypic data of genome-wide single nucleotide polymorphism (SNP) markers. Utilizing these resources, this study aimed at identifying genetic mechanisms of DPR heterosis through a genome-wide association study (GWAS) of additive, dominance, and pairwise epistasis effects.

2. Results and Discussion

The GWAS analysis detected 6528 additive effects and 65 dominance effects with log10(1/p) > 8, and 1659 pairwise epistasis effects with log10(1/p) > 12 for DPR in the 40,203 crossbred dairy cows. Of the 1659 pairwise effects, 1638 were additive × additive (A × A) effects, 3 were additive × dominance (A × D) or dominance × additive (D × A) effects, and 18 were dominance × dominance (D × D) effects. Of the 1638 A × A effects, 1634 were intra-chromosome and four were inter-chromosome A × A effects. The only inter-chromosome pairwise epistasis effect was a D × A effect. Details of these effects are described below.

2.1. Additive Effects

The 6528 additive effects were distributed over all chromosomes (Figure 1a, Table S1). SNP rs108997231 in the SLX4IP gene of Chr13 had the most significant additive effect with log10(1/p) = 30.49, followed by SNPs on chromosomes 1, 17, 20, and 26 (Figure 1a, Table 1). In general, the difference between the statistical significance of any two additive effects ranked next to each other were relatively small. Therefore, the results of additive effects indicated genome-wide contributions of additive effects to DPR in the crossbred dairy cows in this study. Among the top 1000 significant SNP effects, negative allelic effects on average had larger sizes (average −0.957) than positive allelic effects (average 0.629), but the positive alleles had higher frequencies (average 0.60) than the frequencies of the negative alleles (average 0.40) (Table S1). SNP rs42052893 in the KDSR gene of Chr24 had the largest positive allelic effect (2.62) with # 663 ranking in statistical significance, whereas SNP rs110314053 upstream of LOC107132214 of Chr01 had the largest negative allelic effect (–2.17) with #562 ranking in statistical significance (Table S1, Figure 1b). Four examples of the distribution of allelic effects of the top 1000 additive effects on different chromosomes are shown in Figure 1c–f. Compared to the Holstein DPR studies, this crossbred study had a larger number of additive effects but lacked chromosome regions that were much more significant than other chromosome regions. The 6528 additive effects in this crossbred study with 40,203 cows were more than six times as many as the 1126 additive effects from the 2019 Holstein study using 245,214 cows [15] that were more than nine times as many as the 40,203 cows in this crossbred study, and were only 14% fewer than the 7567 additive effects from the Holstein study using 1,194,736 cows [16] that were 29 times as many as the 40,203 cows in this crossbred study. Based on the comparison between the crossbred and the 2019 Holstein studies, the number of significant additive effects in the crossbred study was at least six times as many as in Holsteins, indicating that additive effects were a contributing factor to DPR heterosis in crossbred dairy cows.

2.2. Dominance Effects

The 65 dominance effects were distributed over 19 autosomes and the X chromosomes with Chr20 having the largest number (17) of dominance effects (Table S2, Figure 2a). SNP rs109688013 in the MC1R gene of Chr18 had the most significant dominance effect with log10(1/p) = 18.19, the most positive dominance value (d22 = 1.92) and the most negative dominance value (d12 = –1.28), and the largest negative dominance effect, δ = d12 − 0.5 (d11 + d22) = –2.55. Other SNPs with significant dominance effects following rs109688013 included SNPs on chromosomes X, 10, 20, and 12 (Figure 2a, Table 2). The number of dominance effects (65) is much smaller than the number of additive effects (6528) in this study but is far more than the two significant dominance effects from the 2019 Holstein study using 245,214 cows [15].
A striking feature of the dominance effects was the consistent pattern of homozygous advantage and heterozygous disadvantage according to the three dominance values (dij) of each significant dominance effect (δ), showing that the homozygous genotypes all had positive dominance values, one of the two homozygous genotypes of each SNP had the highest dominance value, and the heterozygous SNP genotypes of all 65 SNPs had the lowest dominance values that were all negative (Figure 2b,c; Table S2). Since each dominance value is a deviation of the genotypic value from the population mean of all three genotypic values and the additive value of the SNP genotype, the pattern of each dominance value could be affected by the estimate of the additive value of the genotype. Therefore, the phenomenon of homozygous advantage and heterozygous disadvantage detected by dominance values was further validated by the genotypic means of two types of phenotypic values: the original phenotypic values without removing pedigree additive (breeding) values (yij) and the corrected phenotypic values after removing pedigree additive (breeding) values (gij). The results of the original phenotypic values confirmed the phenomenon of homozygous advantage and heterozygous disadvantage for all 65 dominance effects, i.e., y11 or y22 had the highest genotypic mean whereas y12 had the lowest genotypic mean. The results of the corrected phenotypic values had only four SNPs that did not have the lowest genotypic means for the heterozygous genotypes (g12) but one of the homozygous genotypes of every SNP was still the most positive genotype for all 65 dominance effects. Therefore, the dominance and original phenotypic values confirmed that all dominance effects had homozygous advantage and heterozygous disadvantage whereas the corrected phenotypic values confirmed homozygous advantage for all 65 dominance effects and heterozygous disadvantage for 61 of the 65 dominance effects (Figure 2d, Table S2).

2.3. Additive × Additive (A × A) Epistasis Effects

The 1658 significant A × A effects detected in the 40,203 crossbred cows were far more than just one significant A × A effect detected in the 2021 Holstein study using 245,214 cows [17]. Of the 1658 A × A effects, 1654 were intra-chromosome and four inter-chromosome A × A effects (Table S3). The 1658 A × A epistasis effects were distributed on all chromosomes but lacked chromosome regions with A × A effects far more significant than in other regions (Figure 3). The significance and distribution patterns of the A × A effects indicated genome-wide contributions of A × A effects to DPR heterosis, like the contributions of the additive effects. Although all chromosomes had intra-chromosome A × A epistasis effects, different chromosomes had large variations in the number of intra-chromosome A × A effects, e.g., Ch20 had 110 effects whereas Chr28 only had five effects. The distance between two interacting SNPs on the same chromosome was in the range of 6702 bp to 2.68 Mb, and many interacting SNP pairs were within the same genes. The A × A epistasis effects had an interesting pattern: the most positive and the most negative allelic combinations always involved one allele of one SNP and two alleles at the other interacting SNP (Table S3). The top 20 A × A effects are shown in Table 3.
Within each chromosome, some chromosome regions had large concentrations of intra-chromosome A × A epistasis effects and six examples of such large concentrations are shown in Figure 4. The 1.08–3.71 Mb region of Chr13 had 31 pairs of A × A epistasis effects mostly involving six genes, PLCB1, PLCB4, PAK5, ANKEF1, SNAP25, and SLX4IP (Table S3, Figure 4a). Through interaction with neighboring genes, the 31 pairs of A × A epistasis effects formed an interacting chromosome region about 2.63 Mb in size. It was interesting to note that the two genes at the two ends of this region, PLCB1 at the upstream end and SLX4IP at the downstream end, were reported to be associated with fertility, with PLCB1 affecting embryonic implantation failure in mice and SLX4IP affecting embryogenesis in zebrafish [18]. It was also interesting to note that SLX4IP had the most significant additive effect in this study. The 31–51 Mb region of Chr14 (Figure 4b), the 24–74Mb region of Chr05 (Figure 4c), and the 22–32 Mb region of Chr17 (Figure 4d) each had a large cluster of A × A epistasis effects; Chr01 (Figure 4e) and Chr03 (Figure 4f) each had A × A epistasis effects covering over half of the chromosome.
Only four significant inter-chromosome A × A effects were detected between Chr02 and Chr12, Chr08 and Chr19, Chr07 and Chr18, and Chr20 and Chr26 (Table S3). The A × A values of the four inter-chromosome A × A effects all had the same pattern as that of the intra-chromosome: the most positive and negative A × A values were due to interactions between one allele of one SNP and two alleles of the other interacting SNP.

2.4. Additive × Dominance (A × D and D × A) Epistasis Effects

The GWAS analysis detected two intra-chromosome A × D effects on Chr20 and one inter-chromosome and D × A effect between Chr04 and Chr19 (Table S4). The HECW1 gene with a D × A effect also had two other SNPs with an intra-chromosome A × A effect. Although these results showed the existence of A × D and D × A effects in the crossbred cows, A × D and D × A effects were unlikely to be an important mechanism underlying DPR heterosis in crossbred cows due to the small number of the A × D and D × A effects.

2.5. Dominance × Dominance (D × D) Epistasis Effects

The GWAS analysis detected 18 intra-chromosome D × D effects on nine chromosomes with Chr20 having the largest number (six) of D × D effects (Table 4). The chromosome distance between two interaction loci was in the range of 3.61 Kb to 1.77 Mb, and only one pair of interacting loci were within the same gene, MAGI1 of Chr22. All 18 D × D effects had a heterozygous disadvantage where any two-locus genotypes with a heterozygous genotype at one locus and a homozygous genotype at the other locus had a negative D × D value. The doubly heterozygous genotype and 68 of the 72 two-locus homozygous genotypes had positive D × D values (Table S5). These D × D results added evidence to the phenomenon of heterozygous disadvantage and homozygous advantage observed for single-locus dominance effects. The doubly heterozygous genotype (AaBb) in fact had a positive value and was an interesting exception of the heterozygous disadvantage. Of the 18 doubly heterozygous genotypes, seven had DD3 values (3rd most positive values), eight had DD4 values, and one had a DD5 value, noting that all 18 DD3 and 18 DD4 values were positive, and DD5 had fourteen positive values and four negative values (Table S5). The positive D × D values of AaBb underwent negative switching twice: the heterozygous genotype at one locus turned a positive value into a negative value and the heterozygous genotype at the other locus turned the negative value back to positive, but the actual reason for the positivity of AaBb is yet to be understood.

2.6. Genetic Mechanism of DPR Heterosis in Dairy Cattle

The GWAS results indicated that the genetic mechanism of DPR heterosis in crossbred dairy cows included a larger collection of favorable alleles than in Holsteins, and allelic and genotypic interactions. These genetic mechanisms can be summarized as the following: complementary additive effects from different breeds, allelic interactions between different breeds, homozygous advantage due to interaction between the same alleles from different breeds and heterozygous disadvantage, and genotype × genotype interactions.

2.6.1. Complementary Additive Effects from Different Breeds and New Additive Effects Due to Cross Breeding

This crossbred study with 40,203 cows had more than six times as many additive effects as the Holstein study on additive and dominance effects using 245,214 cows [15], as described earlier. This result supports the hypothesis that the large number of additive effects were due to complementary additive effects between different breeds, where a chromosome region lacked favorable alleles in one breed, but another breed had favorable alleles in the same region. For the hypothetical example of Figure 5, the chromosome region from the breed represented in green has nine favorable alleles whereas the breed represented in blue has only three favorable alleles in the same region, but the crossbred from these two breeds has 12 favorable alleles, more than those in either breed. The distribution of the additive effects over the entire genome indicated that chromosome regions with complementary additive effects from different breeds widely existed on the cattle genome. Given that Holsteins and Jerseys had the largest genetic contributions (combined 97.11% breed contributions, see Section 3) to the crossbred cows, the additive results indicated the broad existence of complementary additive effects between Holsteins and Jerseys. However, the summation of additive effects in both Holsteins and Jerseys will unlikely reach the number of additive effects observed in this study, noting that the number of additive effects for the DPR in Jerseys was unknown. Therefore, cross breeding likely activated new additive effects that were either undetectable or non-existent in the pure breeds.

2.6.2. Two-Locus Allelic Interactions Between Loci and Between Breeds

This crossbred study with 40,203 cows detected 1658 significant additive A × A effects, compared to only one significant A × A effect from the Holstein study on epistasis effects using 245,214 cows [17], and the only significant Holstein A × A effect was not detected in the crossbred cows. As the genetic effect type with the second largest number of significant effects, A × A effects were the second most common effects in the crossbred cows after additive effects. Two-locus allelic interactions are the genetic interpretation of pairwise A × A effects and such allelic interactions in this study were allelic interactions between breeds.
A significant A × A effect requires at least one pair of alleles, one allele from each locus interacting with each other as described in Figure 6. For two biallelic loci with each locus having one allele from Jersey and one allele from Holstein (Figure 6a), four allelic combinations are possible, AB from Jersey, ab from Holstein, and the Ab and aB allelic combinations each have one allele from Jersey and one allele from Holstein (Figure 6b). The A × A effect of the two loci is a contrast of the mean genotypic values of the four allelic combinations in the population, and allelic interactions must be present for at least one of the four allelic combinations for the A × A effect to be statistically significant (Figure 6c). Although the method in this study does not identify which alleles are interacting with each other, the interacting alleles could be deduced. Since none of the A × A effects in crossbred cows was detected in Holstein cows, the ab allelic combination from Holstein could be assumed to have no interaction between the two alleles. Since Jersey epistasis study was unavailable and Jersey also had lower DPR levels than in crossbred cows, it should be reasonable to assume most of the A × A effects in the crossbred cows were insignificant in Jersey cows. Under this assumption, the AB allelic combination from Jersey also had no interaction between the two alleles, leaving only Ab and aB allelic combinations to be the candidates with allelic interactions responsible for the significant A × A effect, noting that each of these two allelic combinations has one allele from Jersy for one locus and one allele from Holstein for the other locus. Therefore, the mostly likely reason for the significant A × A effects in the crossbred cows was allelic interactions between loci and between breeds.

2.6.3. Within-Locus Allelic Interaction Between Different Breeds Underlying Homozygous Advantage and Heterozygous Disadvantage

Within-locus allelic interactions of a biallelic SNP locus are reflected in the dominance values of the three genotypes of the locus including the two homozygous genotypes and the heterozygous genotype. Assuming a Jersey–Holstein cross, each SNP genotype of a crossbred cow consists of one Jersey allele and one Holstein allele (Figure 7). The dominance value (dij) of a genotype is the deviation of the genotypic value (gij) from the additive value (aij) of the genotype and the population average of all three genotypic values (µ), where the additive value of the genotype is the sum of the two allelic effects. A large dominance value of a genotype, positive or negative, indicates the presence of allelic interactions within the genotype. The example of Figure 7 is the dominance values of rs109688013 in the MC1R gene of Chr18 with the most significant dominance effect (Table 2 and Table S2). The two homozygous genotypes (AjAh or ajah) had positive dominance values (d22 = 1.92 and d11 = 0.629) and the heterozygous genotype (ajAh or Ajah, not distinguishable in this study) had a negative dominance value (d12 = –1.28). Since the two alleles of every genotype originated from different breeds, the within-locus allelic interactions were allelic interactions between different breeds. These dominance values indicated that every genotype had within-locus allelic interaction between breeds that was responsible for the homozygous advantage and heterozygous disadvantage. All the advantages and disadvantages revealed by the dominance values were further confirmed by the original and corrected phenotypic values, as shown in Figure 2d. The interaction between two same alleles from different breeds within a homozygous genotype (AjAh or ajah) was interesting because the concept of dominance is typically associated with the heterozygous genotype. For this reason, the terminology ‘within-locus interaction’ is a more general description than ‘dominance’ for interpreting the dominance values, particularly those of the homozygous genotypes. The quantitative genetics approach of dominance values in this study led to the finding of allelic interactions between breeds within both homozygous and heterozygous genotypes but had no indication as to why two same alleles from different breeds interacted with each other or why homozygous genotypes had advantages and heterozygous genotypes had disadvantages. Compared to the additive and A × A effects, the number of significant dominance effects is relatively small and should have limited contributions to the DPR heterosis.

2.6.4. Genotype × Genotype Interactions Enabled by Allelic Interactions Between Breeds

Genotype × genotype (G × G) interactions are the reasons for the 18 significant D × D effects which also had homozygous advantage and heterozygous advantage. Figure 8 is the example of the most significant D × D effect between rs42670220 and rs41942366 of Chr20 showing G × G interactions reflected in the nine D × D values of DD1-DD9 ranked from the most positive to the most negative. Any two-locus genotype with a heterozygous genotype at one locus and a homozygous genotype at the other locus was negative (highlighted in yellow) and all homozygous genotypes were positive. The D × D values clearly demonstrated the presence of G × G interactions in crossbred cows. The AAbb genotype had the most positive D × D value of DD1 = 1.22, but AA and bb individually were also associated with very negative D × D values: the Aabb genotype had the most negative D × D value of DD9 = –1.36 and the AABb genotype had the second most negative D × D value of DD8 = –1.22. Since parents do not transmit any genotype to the next generations and each breed transmits only one allele per locus to the offspring, the G × G interactions were not direct interactions between breeds but were enabled by the allelic interactions between the parental breeds. Moreover, no G × G interaction was detected in the 2021 Holstein study using 245,214 cows [17]. Therefore, the G × G interaction in crossbred cows enabled by the allelic interactions between the parental breeds was a genetic mechanism underlying DPR heterosis in crossbred cows.

2.7. Candidate Genes with Fertility and Reproductive Functions

The significant genetic effects detected in this study were associated with many candidate genes known to have fertility and reproductive functions. We searched for such genes among the top 20 additive effects, the top 20 dominance effects, the top 20 intra-chromosome A × A effects, the three inter-chromosome A × A effects, the three A × D and D × A effects’ top 20 dominance effects, and the 18 D × D effects. In addition, we noted that some genes with highly significant DPR effects in Holstein cows [15,16] were also significant in the crossbred cows, although the effects of those genes in crossbred cows were not highly ranked. Together, 82 candidate genes known to affect many aspects of fertility and reproduction were identified (Appendix A), indicating that fertility and reproductive genes likely were broadly involved in DPR heterosis in crossbred dairy cows. Important fertility and reproductive functions of the candidate genes included embryo development (KALRN, DAZL, IFNG, SERPINE3, BMP7, PCDH10, FOXC2, SIPA1L3, NDUFS4, GHR, DROSHA, EOMES, KDSR, and RARB), embryonic implantation failure (PCLB1), embryogenesis (SLX4IP), maintenance of pregnancy (IGFBP7 and IGFBP3), placentation and pregnancy maintenance (EIF2S1), uterine contraction and pregnancy maintenance (OXTR), progesterone synthesis (FGF12), early pregnancy (SLC4A4), delayed puberty (DLG2), oogenesis (HECW1 and LARP1B), oocyte maturation and female fertility (CNOT2), sex hormone production (DIAPH1), gonadotropin-releasing hormone (HOMER1), growth hormone (GHR), and thyroid hormone (THRB). Some genes with highly significant effects on the Holstein DPR were also significant but not ranked high in the crossbred cows, including SLC4A4 and GC of Chr06, KALRN of Chr01, and PEPD, CHST8, and SIPA1L3 of Chr18. The KALRN, PEPD, and CHST8 genes had extremely negative recessive alleles for the DPR in Holstein cows [16] but those extremely negative effects were not observed in the crossbred cows.

3. Materials and Methods

3.1. Dairy Crossbred Population and SNP Data

The dairy crossbred population in this study had 40,203 cows with first lactation phenotypic observations on DPR resulting from crosses among five dairy breeds: Ayrshire, Brown Swiss, Guernsey, Holstein, and Jersey. Based on breed base representation (BBR), a measure for identifying breed origin, a cow or bull is considered a purebred animal if the animal’s BBR values is 94–100% [19]. The breed contributions to the 40,203 crossbred cows measured by average BBR per breed as the percentage of all BBR values of the five breeds were 51.80% from Jersey, 45.31% Holstein, 1.60% Ayrshire, 0.99% Brow Swiss, and 0.30% from Guernsey. Therefore, the crossbreeds were primarily due to the Jersey × Holstein cross given that Jersey and Holstein had a combined 97.11% breed contribution to the crossbred cows. Ignoring the ‘0’ BBR values that every breed had, the range of the BBR values was 2.5–89.5% for Jersey, Holstein, Ayrshire, and Brown Swiss; and 2.5–89.2% for Guernsey, indicating varying degrees of purebred genetic contributions among different crossbred cows that should provide necessary contrasts between favorable and unfavorable genetic factors underlying DPR in the crossbred cows. The phenotypic values used in the GWAS analysis were the phenotypic residuals after removing fixed non-genetic effects available from the December 2023 U.S. Holstein genomic evaluation by the Council on Dairy Cattle Breeding (CDCB). Among Holstein, Jersey, and crossbred cows, crossbreds had the highest average DPR values (54.97), followed by Jersey (53.43) and Holstein (49.06) (Table 5). Since 1% DPR is equivalent to four fewer days open, crossbred cows on average had about 24 fewer days open than Holstein cows and six fewer days open than Jersey cows, showing that crossbred cows had DPR heterosis over Holstein and Jersey cows. The SNP genotypes were from 32 SNP chips with various densities and were imputed into 78,964 SNPs via the FindHap algorithm [20] as a routine procedure for genomic evaluation by CDCB [21]. The SNP genotyping quality control by CDCB had checks and requirements at the individual and SNP levels, including call rate, parent–progeny conflicts, sex verification using X-specific SNPs, and Hardy-Weinberg equilibrium [22,23]. In addition, we applied minor allele frequency (MAF) of 5% for SNP filtering in this study. With the requirement of 5% MAF, the number of SNPs for the GWAS analysis was 75,133. The threshold p-value for declaring significant effects for the Bonferroni correction with 0.05 genome-wide false positives was 0.05/[(2)(75,133)] = 10−8 or log10(1/p) = 8 for additive and dominance effects and was 0.05/[(4)(75,133)(75,132)/2] = 10−12 or log10(1/p) = 12 for pairwise epistasis effects. The SNP and gene positions were those from the ARS-UCD1.3 cattle genome assembly [24]. Genes containing or in proximity to highly significant SNP effects were identified as candidate genes for DPR in the crossbred dairy cows.

3.2. GWAS Analysis

The GWAS analysis used an approximate generalized least squares (AGLS) method. The AGLS method combines the least squares (LS) tests implemented by EPISNP1mpi [25,26] with the phenotypic correction using the estimated breeding values from routine genetic evaluation using the entire U.S. Holstein population. For each SNP pair, the statistical model was
y = µ1 + Xgg + Za + e = Xb + Za + e
where y = column vector of phenotypic deviation after removing fixed nongenetic effects such as heard–year–season (termed as ‘yield deviation’ for any trait) using a standard procedure for the CDCB genetic and genomic evaluation; µ = common mean; 1 = column vector of 1’s; g = column vector of genotypic values of the nine genotypes of the SNP pair; Xg = model matrix of g; b = ( μ , g ) ; X = (1, Xg); a = column vector of pedigree additive values; Z = model matrix of a; and e = column vector of random residuals. The first and second moments of Equation (1) are E(y) = Xb and v a r y = V = Z G Z + R = σ a 2 Z A Z + σ e 2 I , where σ a 2 = additive variance, A = additive relationship matrix, σ e 2   = residual variance, and I = identity matrix. The problem of estimating the b vector that includes SNP genotypic values in Equation (1) is the requirement of inverting the V if the generalized least squares (GLS) method is used or solving the mixed model equations (MMEs) [27]. Either the GLS or MME method for each of the genome-wide SNPs is computationally challenging. To avoid these computing difficulties, the GWAS used the AGLS method that replaces the pedigree additive values (a) with the best linear unbiased prediction based on pedigree relationships [15] that were readily available from CDCB’s routine genetic and genomic evaluations. The AGLS method is based on the following results:
b ^ = ( X V 1 X ) X V 1 y
b ^ = X R 1 X X R 1 y X R 1 Z a ^ = X X X ( y Z a ^ ) = X X X y
where y = y Z a ^ and a ^ is the best linear unbiased prediction (BLUP) of a. Equation (2) is the GLS solution and Equation (3) is the MME solution of b. These two equations yield identical results and b ^ from either equation is termed the best linear unbiased estimator (BLUE) [27]. If a ^ is known, the LS version of BLUE given by Equation (3) is computationally efficient relative to the GLS of Equation (2) that requires the V inverse or the joint MME solutions of b ^ and a ^ . The AGLS method uses two approximations. The first approximation is to use a ~ from routine genetic evaluation as an approximation of a ^ in Equation (3):
b ^ = X X X y Z a ~ = X X X y
where y = y Z a ~ and a ~ is the column vector of 2(PTA) with PTA being the predicted transmitting ability from the routine genetic evaluation. Equation (4) achieves the benefit of sample stratification correction from mixed models using pedigree relationships without the computing difficulty of inverting V or the joint MME solutions of b ^ and a ^ for every SNP. The second approximation of the AGLS approach is the t-test using the LS rather than the GLS formula of the t-statistic to avoid using the V inverse in the GLS formula. The significance tests were used for the t-tests of the contrasts of the estimated SNP genotypic values of the nine genotypes of each SNP pair where locus 1 had AA, Aa, and aa genotypes and locus 2 had BB, Bb, and bb genotypes [25,28]. The t-statistic of the AGLS was calculated as
t j = L j v a r ( L j ) = s j g ^ v s j ( X X ) g g s j , j = A , D , A × A , A × D , D × A , D × D
where Lj = effect contrast; v a r ( L j )   = standard deviation of the effect contrast; s j = row vector of the contrast coefficients for the j t h effect type; g ^ = column vector of the AGLS estimates of the nine SNP genotypic values for two loci from Equation (4); A = additive; D = dominance; A × A = additive × additive; A × D = additive × dominance; D × A = dominance × additive; D × D = dominance × dominance; v 2 = ( y X b ^ ) ( y X b ^ ) / ( n k )   = estimated residual variance; ( X X ) g g = submatrix of X X corresponding to g ^ ; n = number of observations; and k = rank of X. The formula of s a defined above allows Hardy-Weinberg disequilibrium [28]. To understand the genetic effects, the genetic values for each effect contrast (Lj) were calculated based on Kempthorne’s definitions [29,30] with an extension to allow for Hardy–Weinberg and linkage disequilibria [28] implemented by the EPISNP1mpi computing tool [25,26]:
a i = µ i μ , i = A ,   a
a k = µ k μ , k = B ,   b
d i j = µ i j μ a i a j , ij = AA ,   Aa ,   aa
d k l = µ k l μ a k a l , kl = BB ,   Bb ,   bb
a i k = µ i k μ a i a k , ik = AB ,   Ab ,   aB ,   ab
( a d ) i k l = µ i k l μ     a i a k a l d k l ( a a ) i k ( a a ) i l i = A ,   a ;   kl = BB ,   Bb ,   bb
( d a ) i j k = µ i j k     μ     a i a j a k d i j ( a a ) i k ( a a ) j k , ij = AA ,   Aa ,   aa ;   k = B ,   b
( d d ) i j k l = g i j k l μ a i a j   a k a l d i j d k l   ( a a ) i k ( a a ) i l   ( a a ) j k ( a a ) i l ( a d ) i k l ( a d ) j k l ( d a ) i j k ( d a ) i j l , ij = AA ,   Aa ,   aa ;   kl = BB ,   Bb ,   bb
where a i = allelic effect of allele i at locus 1 (i = A, a), a k = allelic effect of allele k at locus 2 (k = B, b), d i j = dominance value of genotype ij at locus 1, d k l = dominance value of genotype kl at locus 2, (aa)ik = additive × additive epistasis value of genotypes with alleles ik, ( a d ) i k l = additive × dominance epistasis value of genotypes with alleles ikl, ( d a ) i j k = dominance × additive epistasis value of genotypes with alleles ijk, ( d d ) i j k l = dominance × dominance epistasis value of ijkl genotype, μ = the mean genotypic value in the population, µ i = the marginal mean of genotypic values for individuals with allele i (i = A, a), µ k = the marginal mean of genotypic values for individuals with allele k (k = B, b), µ i j = the marginal mean of genotypic values for individuals with genotype ij at locus 1 (ij = AA, Aa, and aa), µ i k = the marginal mean of genotypic values for individuals with genotype kl at locus 2 (kl = BB, Bb, and bb), µ i k = the marginal mean of genotypic values for individuals with allele i at locus 1 and allele k at locus 2 (i = A, a; k = B, b), µ i k l = the marginal mean of genotypic values for individuals with allele i at locus 1 and genotype kl at locus 2 (i = A, a; kl = BB, Bb, and bb), and µ i j k = the marginal mean of genotypic values for individuals with genotype ij at locus 1 and allele k at locus 2 (ij = AA, Aa, aa; k = B, b).

4. Conclusions

The GWAS using 40,203 crossbred dairy cows showed that genome-wide complementary additive effects between breeds and A × A epistasis effects due to between-breed and between-locus allelic interactions were the primary genetic mechanism of DPR heterosis. Interesting and complex genetic mechanisms of DPR heterosis were observed, including dominance effects due to between-breed and within-locus allelic interactions with homozygous advantage and heterozygous disadvantage and D × D effects enabled by between-breed and between-locus allelic interactions also with homozygous advantage and heterozygous disadvantage.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms262211149/s1.

Author Contributions

Y.D. and L.M. conceived this study. R.Y. and Z.L. conducted the data analysis. D.P. contributed to data and graphical work. L.M. contributed to computing tools and manuscript reviews. Y.D. and R.Y. prepared the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Institutes of Health’s National Human Genome Research Institute, grant R01HG012425 as part of the NSF/NIH Enabling Discovery through GEnomics (EDGE) Program; grant 2020-67015-31133 from the USDA National Institute of Food and Agriculture; and project MIN-16-124 of the Agricultural Experiment Station at the University of Minnesota. The funders had no role in the study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Institutional Review Board Statement

Ethical review and approval were waived because this study used existing data only and did not involve the use of live animals.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original genotype data are owned by third parties and maintained by the Council on Dairy Cattle Breeding (CDCB). A request to the CDCB is necessary for obtaining data access on the research, which may be sent to: João Dürr, CDCB Chief Executive Officer (joao.durr@cdcb.us). All other relevant data are available in the manuscript and Supplementary Materials.

Acknowledgments

Members of the Council on Dairy Cattle Breeding (CDCB) and the Cooperative Dairy DNA Repository (CDDR) are acknowledged for providing the dairy genomic evaluation data.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Candidate Genes with Fertility and Reproductive Functions

SNPChrPositionCandidate
Gene
Effect TypeFunction
rs110391667168682461KALRNAAintraEmbryo development [31], Holstein cow conception rate, and DPR [16]
rs133450760168949628KALRNAEmbryo development [31], Holstein cow conception rate, and DPR [16]
rs110468756175346472FGF12AProgesterone synthesis [32]
rs136701022175755758FGF12AAintraProgesterone synthesis [32]
rs110789800176375060GMNCAGrowth and fertility [33]
rs41640538181200460ETV5ASpermatogenesis and folliculogenesis [34,35]
rs432784531142383844UMODL1DDintraOocyte quality [36]
rs415836971146007587PCNTDFemale fertility [37]
rs1107838981153664085DAZLAAintraGerm cells development in embryos [38]
rs110991306294592808GPR1AAintraSteroidogenesis in ovaries [39]
rs1100270852100793281ERBB4ARegulation of sertoli and germ cell adhesion [40]
rs433152192106815376WNT6AAintraStromal proliferation during decidualization [41]
rs1100993082122454418PUM1DEstablishment of the primordial follicle pool [42]
rs1343359642128057688RUNX3AAintraRegulates ovarian functions and ovulation [43]
rs110808151319209742TUFT1AAintraMale fertility [44]
rs111002920327938935NGFAAintraFolliculogenesis and spermatogenesis [45,46]
rs43712201397857026SPATA6AEssential for sperm formation [47]
BTB-01885061438854540HGFAMale and female fertility [48]
rs43406699476069283IGFBP3AAintraFollicle and pregnancy maintenance [49,50]
rs42664831477659486HECW1AAintraOogenesis [51]
rs110544309477811545HECW1DAinterOogenesis [51]
BTB-01477571543254214CNOT2AOocyte maturation and female fertility [52]
rs43438147543898744FRS2AEpididymal development [53]
rs136525396545641750IFNGAAintraSuccess of fertilization and embryo development [54]
rs134371136547933174HMGA2AMale and female infertility [55]
rs42516642575694606RAC2ARemodeling of the endometrial stroma [56]
rs110601510670246689KITAMale infertility [57]
rs41654346672581115IGFBP7A, DMaintenance of pregnancy [58]
rs134948894681156919EPHA5AAintraFemale fertility [59]
rs3423226650685656834CSN3AHolstein DPR [16]
rs43474151686508590SLC4A4AHolstein DPR [16]
rs110434046687184768GCAHolstein cow conception rate and DPR [16]
rs109034709687316810NPFFR2AHolstein cow conception rate and DPR [16]
rs43475934692430962CNOT6LAFemale fertility [60]
rs110103910752664571DIAPH1AAinterSex hormone production [61]
rs136410320761491583PDGFRBAAintraEmbryonic testis development [62]
rs43549012836818848PTPRDAAinterGestational diabetes risk in pregnant women [63]
rs109347873882976354HSD17B3AOvary morphological traits of cows [64]
rs1350503438105307551PAPPAAOvarian steroidogenesis and female fertility [65]
UA-IFASA-5710915743940IMPG1DMale fertility [66]
rs290199531010520078HOMER1AAintraGonadotropin [67]
rs1342579231079265033EIF2S1DPlacentation and pregnancy maintenance [68]
rs1101478811220815304SERPINE3DFollicle count and embryo development [69]
rs42391346131641651PCLB1AAintraEmbryonic implantation failure [70]
rs108997231133906114SLX4IPADNA repair and maintenance [71] and embryogenesis [18]
rs41605729133710010SLX4IPAAintraDNA repair and maintenance [71] and embryogenesis [18]
rs1326977511325821584ENKURDDintraSperm motility [72]
rs417045491358979512BMP7AAintraEmbryo development [73]
rs415872521376013752NCOA3AGerm cell and fertility-related phenotypes [74]
rs34230933201435595531TRPA1AAintraSpermatogenesis and uterine function [75,76]
rs1340520041440390594PEX2A, AAintraMale fertility [77]
rs1106003241575950150PEX16ASpermatocyte maturation [78]
rs415769591615915181BRINP3DInfluencing horse reproductive traits [79]
rs1103486661725067804PCDH10AAintraEmbryo development [80]
rs1097789581729595964LARP1BAOogenesis [81]
rs1092892431766235451TPST2DDintraMale infertility [82]
rs1369938001812377354FOXC2AAintraEmbryo development [83]
rs1096880131814705671MC1RDFertility and prenatal development [84]
rs1378106051816507273ABCC12AAinterMale fertility [85]
rs1350722521840169055HYDINAAintraSperm-tail movement [86]
rs1334299141843836519PEPDAHolstein cow conception rate and DPR [16]
rs418795571844031531CHST8AHolstein cow conception rate, DPR [16], and productive life [87]
rs1100617001847814171SIPA1L3AEmbryo development [31], Holstein cow conception rate, and DPR [16]
rs1337063731856085584PRMT1DMale fertility and embryo development [88,89]
rs416003571958869903SOX9AAintraTestis development and fertility [90]
rs1108238451961080519MAP2K6ATestis determination [91]
rs416384042025229625NDUFS4AEmbryo development [92]
rs422786492029647234HCN1ADintraFunctional significance in ovaries [93]
rs1358532962032077537GHRAOvarian follicular growth [94]
rs1090141482033024411PLCXD3DDintraSpermatogenic dysfunction [95]
rs412570662035956386LIFRAEmbryo implantation [96]
rs419446642041658534PDZD2DRisk factor to recurrent miscarriage [97]
rs1101167852041999241DROSHADEmbryo development [98]
rs1095350192046964318CDH9ADintraSperm capacitation and egg—sperm binding [99]
rs429864482051571771CDH12AAinterFemale fertility [100]
rs1099013142159873899DICER1AAintraFemale fertility [101]
rs41993852221986320EOMESAEmbryo development [102]
rs1099723682217708467OXTRAAintraUterine contraction and pregnancy maintenance [103]
rs1375909822217910318LMCD1AMale fertility [104]
rs1361646852217954298LMCD1AAintraMale fertility [104]
rs137797676239112456TCP11AAintraSperm capacitation, motility, and acrosome reaction [105,106]
rs1100808472421456850INO80CASpermatogenesis [107]
rs420528932461612123KDSRAEmbryo development [108]
rs1100280072740780768RARBAAintraEmbryo development [109]
rs1347694622741433477THRBA, DFemale fertility [110]
rs1335073152911356071DLG2A, AAintraDelayed puberty [111]
In ‘Effect Type’ column, A = additive effect, D = dominance effect, AAintra = intra-chromosome A × A effect, AAinter = inter-chromosome A × A effect, ADintra = intra-chromosome A × D effect, ADinter = inter-chromosome A × D effect, and DDintra = intra-chromosome D × D effect, DAinter = inter-chromosome D × A effect.

References

  1. VanRaden, P.; Sanders, A.; Tooker, M.; Miller, R.; Norman, H.; Kuhn, M.; Wiggans, G. Development of a national genetic evaluation for cow fertility. J. Dairy Sci. 2004, 87, 2285–2292. [Google Scholar] [CrossRef]
  2. CDCB. Impact on U.S. Dairy. Available online: https://uscdcb.com/impact/ (accessed on 7 September 2025).
  3. Auldist, M.; Pyman, M.; Grainger, C.; Macmillan, K. Comparative reproductive performance and early lactation productivity of Jersey × Holstein cows in predominantly Holstein herds in a pasture-based dairying system. J. Dairy Sci. 2007, 90, 4856–4862. [Google Scholar] [CrossRef]
  4. Begley, N.; Pierce, K.; Buckley, F. Milk production, udder health, body condition score and fertility performance of Holstein-Friesian, Norwegian Red and Norwegian Red × Holstein-Friesian cows on Irish dairy farms. In Breeding for Robustness in Cattle; Wageningen Academic: Wageningen, The Netherlands, 2009; pp. 191–198. [Google Scholar]
  5. Hazel, A.; Heins, B.J.; Seykora, A.J.; Hansen, L.B. Production, fertility, survival, and body measurements of Montbéliarde-sired crossbreds compared with pure Holsteins during their first 5 lactations. J. Dairy Sci. 2014, 97, 2512–2525. [Google Scholar] [CrossRef]
  6. Heins, B.J.; Hansen, L.B.; Hazel, A.; Seykora, A.J.; Johnson, D.; Linn, J. Jersey × Holstein crossbreds compared with pure Holsteins for body weight, body condition score, fertility, and survival during the first three lactations. J. Dairy Sci. 2012, 95, 4130–4135. [Google Scholar] [CrossRef]
  7. Shull, G.H. What is” heterosis”? Genetics 1948, 33, 439. [Google Scholar] [CrossRef]
  8. Birchler, J.A.; Yao, H.; Chudalayandi, S.; Vaiman, D.; Veitia, R.A. Heterosis. Plant Cell 2010, 22, 2105–2112. [Google Scholar] [CrossRef] [PubMed]
  9. Huang, X.; Yang, S.; Gong, J.; Zhao, Y.; Feng, Q.; Gong, H.; Li, W.; Zhan, Q.; Cheng, B.; Xia, J. Genomic analysis of hybrid rice varieties reveals numerous superior alleles that contribute to heterosis. Nat. Commun. 2015, 6, 6258. [Google Scholar] [CrossRef]
  10. Akanno, E.C.; Chen, L.; Abo-Ismail, M.K.; Crowley, J.J.; Wang, Z.; Li, C.; Basarab, J.A.; MacNeil, M.D.; Plastow, G.S. Genome-wide association scan for heterotic quantitative trait loci in multi-breed and crossbred beef cattle. Genet. Sel. Evol. 2018, 50, 48. [Google Scholar] [CrossRef]
  11. Hollick, J.B.; Chandler, V.L. Epigenetic allelic states of a maize transcriptional regulatory locus exhibit overdominant gene action. Genetics 1998, 150, 891–897. [Google Scholar] [CrossRef] [PubMed]
  12. Krieger, U.; Lippman, Z.B.; Zamir, D. The flowering gene SINGLE FLOWER TRUSS drives heterosis for yield in tomato. Nat. Genet. 2010, 42, 459–463. [Google Scholar] [CrossRef] [PubMed]
  13. Hashimoto, S.; Wake, T.; Nakamura, H.; Minamiyama, M.; Araki-Nakamura, S.; Ohmae-Shinohara, K.; Koketsu, E.; Okamura, S.; Miura, K.; Kawaguchi, H. The dominance model for heterosis explains culm length genetics in a hybrid sorghum variety. Sci. Rep. 2021, 11, 4532. [Google Scholar] [CrossRef]
  14. Boeven, P.H.; Zhao, Y.; Thorwarth, P.; Liu, F.; Maurer, H.P.; Gils, M.; Schachschneider, R.; Schacht, J.; Ebmeyer, E.; Kazman, E. Negative dominance and dominance-by-dominance epistatic effects reduce grain-yield heterosis in wide crosses in wheat. Sci. Adv. 2020, 6, eaay4897. [Google Scholar] [CrossRef] [PubMed]
  15. Jiang, J.; Ma, L.; Prakapenka, D.; VanRaden, P.M.; Cole, J.B.; Da, Y. A large-scale genome-wide association study in US Holstein cattle. Front. Genet. 2019, 10, 412. [Google Scholar] [CrossRef]
  16. Liang, Z.; Prakapenka, D.; VanRaden, P.M.; Jiang, J.; Ma, L.; Da, Y. A Million-cow genome-wide association study of three fertility traits in US Holstein cows. Int. J. Mol. Sci. 2023, 24, 10496. [Google Scholar] [CrossRef] [PubMed]
  17. Prakapenka, D.; Liang, Z.; Jiang, J.; Ma, L.; Da, Y. A Large-scale genome-wide association study of epistasis effects of production traits and daughter pregnancy rate in US Holstein cattle. Genes 2021, 12, 1089. [Google Scholar] [CrossRef] [PubMed]
  18. Ingham, A.; Ramaswami, M.; Ramangoudr-Bhojappa, R.; Pladevall-Morera, D.; De Santis, F.; Terriente, J.; Muñoz, I.M.; Rouse, J.; Chandrasekharappa, S.C.; Lopez-Contreras, A.J. Loss of SLX4IP leads to Common Fragile Sites instability and compromises DNA interstrand crosslink repair in vivo. J. Biol. Chem. 2025, 301, 110244. [Google Scholar] [CrossRef]
  19. CDCB. CDCB Genomic Dictionary. Available online: https://redmine.uscdcb.com/projects/cdcb-customer-service/wiki/CDCB_Genomic_Dictionary (accessed on 7 September 2025).
  20. VanRaden, P.M.; Sun, C.; O’Connell, J.R. Fast imputation using medium or low-coverage sequence data. BMC Genet. 2015, 16, 82. [Google Scholar] [CrossRef]
  21. CDCB Genomic Evaluations. Available online: https://uscdcb.com/genomic-evaluations/ (accessed on 7 September 2025).
  22. Wiggans, G.; VanRaden, P.; Cooper, T. The genomic evaluation system in the United States: Past, present, future. J. Dairy Sci. 2011, 94, 3202–3211. [Google Scholar] [CrossRef]
  23. CDCB. QC Metrics for Genotyping Laboratories. Available online: https://redmine.uscdcb.com/projects/cdcb-customer-service/wiki/QC_Metrics_for_Genotyping_Laboratories (accessed on 7 September 2025).
  24. National Library of Medicine (NCBI). Available online: https://www.ncbi.nlm.nih.gov/genome/82?genome_assembly_id=1850378 (accessed on 7 September 2025).
  25. Ma, L.; Runesha, H.B.; Dvorkin, D.; Garbe, J.; Da, Y. Parallel and serial computing tools for testing single-locus and epistatic SNP effects of quantitative traits in genome-wide association studies. BMC Bioinform. 2008, 9, 315. [Google Scholar] [CrossRef]
  26. Weeks, N.T.; Luecke, G.R.; Groth, B.M.; Kraeva, M.; Ma, L.; Kramer, L.M.; Koltes, J.E.; Reecy, J.M. High-performance epistasis detection in quantitative trait GWAS. Int. J. High Perform. Comput. Appl. 2016, 32, 1094342016658110. [Google Scholar] [CrossRef]
  27. Henderson, C. Applications of Linear Models in Animal Breeding; University of Guelph: Guelph, ON, Canada, 1984. [Google Scholar]
  28. Mao, Y.; London, N.R.; Ma, L.; Dvorkin, D.; Da, Y. Detection of SNP epistasis effects of quantitative traits using an extended Kempthorne model. Physiol. Genom. 2006, 28, 46–52. [Google Scholar] [CrossRef]
  29. Kempthorne, O. The correlation between relatives in a random mating population. Proc. R. Soc. London. Ser. B-Biol. Sci. 1954, 143, 103–113. [Google Scholar] [CrossRef]
  30. Kempthorne, O. An Introduction to Genetic Statistics; Wiley: New York, NY, USA, 1957. [Google Scholar]
  31. Parnell, E.; Shapiro, L.P.; Voorn, R.A.; Forrest, M.P.; Jalloul, H.A.; Loizzo, D.D.; Penzes, P. KALRN: A central regulator of synaptic function and synaptopathies. Gene 2021, 768, 145306. [Google Scholar] [CrossRef]
  32. Yang, W.L.; Yu, S.Q.; Peng, J.Z.; Chang, P.H.; Chen, X.Y. FGF12 regulates cell cycle gene expression and promotes follicular granulosa cell proliferation through ERK phosphorylation in geese. Poult. Sci. 2023, 102, 102937. [Google Scholar] [CrossRef]
  33. Terré, B.; Piergiovanni, G.; Segura-Bayona, S.; Gil-Gómez, G.; Youssef, S.A.; Attolini, C.S.O.; Wilsch-Bräuninger, M.; Jung, C.; Rojas, A.M.; Marjanovic, M.; et al. GEMC1 is a critical regulator of multiciliated cell differentiation. EMBO J. 2016, 35, 942–960. [Google Scholar] [CrossRef]
  34. Eo, J.; Song, H.; Lim, H.J. Etv5, a transcription factor with versatile functions in male reproduction. Clin. Exp. Reprod. Med. 2012, 39, 41–45. [Google Scholar] [CrossRef]
  35. Eo, J.; Shin, H.; Kwon, S.; Song, H.; Murphy, K.M.; Lim, J.H. Complex ovarian defects lead to infertility in Etv5-/- female mice. Mol. Hum. Reprod. 2011, 17, 568–576. [Google Scholar] [CrossRef]
  36. Wang, W.; Liu, H.-C.; Jung, M.; He, Z.; Rosenwaks, Z. Functional uromodulin-like 1 (Umodl1), a glycosylated surface protein is essential to oocyte quality. Fertil. Steril. 2008, 90, S107–S108. [Google Scholar] [CrossRef]
  37. Baumann, C.; Wang, X.; Yang, L.; Viveiros, M.M. Error-prone meiotic division and subfertility in mice with oocyte-conditional knockdown of pericentrin. J. Cell Sci. 2017, 130, 1251–1262. [Google Scholar] [CrossRef]
  38. Lin, Y.; Page, D.C. Dazl deficiency leads to embryonic arrest of germ cell development in XY C57BL/6 mice. Dev. Biol. 2005, 288, 309–316. [Google Scholar] [CrossRef]
  39. Yang, Y.-L.; Sun, L.-F.; Yu, Y.; Xiao, T.-X.; Wang, B.-B.; Ren, P.-G.; Tang, H.-R.; Zhang, J.V. Deficiency of Gpr1 improves steroid hormone abnormality in hyperandrogenized mice. Reprod. Biol. Endocrinol. 2018, 16, 50. [Google Scholar] [CrossRef]
  40. Naillat, F.; Veikkolainen, V.; Miinalainen, I.; Sipila, P.; Poutanen, M.; Elenius, K.; Vainio, S.J. ErbB4, a receptor tyrosine kinase, coordinates organization of the seminiferous tubules in the developing testis. Mol. Endocrinol. 2014, 28, 1534–1546. [Google Scholar] [CrossRef]
  41. Wei, M.; Zhang, C.; Tian, Y.; Du, X.; Wang, Q.; Zhao, H. Expression and function of WNT6: From development to disease. Front. Cell Dev. Biol. 2020, 8, 558155. [Google Scholar] [CrossRef]
  42. Mak, W.; Fang, C.; Holden, T.; Dratver, M.B.; Lin, H. An Important role of pumilio 1 in regulating the development of the mammalian female germline. Biol. Reprod. 2016, 94, 134. [Google Scholar] [CrossRef]
  43. Ojima, F.; Saito, Y.; Tsuchiya, Y.; Kayo, D.; Taniuchi, S.; Ogoshi, M.; Fukamachi, H.; Takeuchi, S.; Takahashi, S. Runx3 transcription factor regulates ovarian functions and ovulation in female mice. J. Reprod. Dev. 2016, 62, 479–486. [Google Scholar] [CrossRef]
  44. Lewandowski, J.P.; Dumbović, G.; Watson, A.R.; Hwang, T.; Jacobs-Palmer, E.; Chang, N.; Much, C.; Turner, K.M.; Kirby, C.; Rubinstein, N.D. The Tug1 lncRNA locus is essential for male fertility. Genome Biol. 2020, 21, 237. [Google Scholar] [CrossRef]
  45. Dissen, G.A.; Romero, C.; Hirshfield, A.N.; Ojeda, S.R. Nerve growth factor is required for early follicular development in the mammalian ovary. Endocrinology 2001, 142, 2078–2086. [Google Scholar] [CrossRef]
  46. Stabile, A.M.; Pistilli, A.; Moretti, E.; Bartolini, D.; Ruggirello, M.; Rende, M.; Castellini, C.; Mattioli, S.; Ponchia, R.; Tripodi, S.A.; et al. A possible role for nerve growth factor and its receptors in human sperm pathology. Biomedicines 2023, 11, 3345. [Google Scholar] [CrossRef]
  47. Yuan, S.Q.; Stratton, C.J.; Bao, J.Q.; Zheng, H.L.; Bhetwal, B.P.; Yanagimachi, R.; Yan, W. Spata6 is required for normal assembly of the sperm connecting piece and tight head-tail conjunction. Proc. Natl. Acad. Sci. USA 2015, 112, E430–E439. [Google Scholar] [CrossRef]
  48. Mi, X.; Chen, C.Y.; Feng, C.; Qin, Y.Y.; Chen, Z.J.; Yang, Y.J.; Zhao, S.D. The functions and application prospects of hepatocyte growth factor in reproduction. Curr. Gene Ther. 2024, 24, 347–355. [Google Scholar] [CrossRef] [PubMed]
  49. Amato, G.; Izzo, A.; Tucker, A.; Bellastella, A. Insulin-like growth factor binding protein-3 reduction in follicular fluid in spontaneous and stimulated cycles. Fertil. Steril. 1998, 70, 141–144. [Google Scholar] [CrossRef]
  50. Mense, K.; Heidekorn-Dettmer, J.; Wirthgen, E.; Brockelmann, Y.; Bortfeldt, R.; Peter, S.; Jung, M.; Hoflich, C.; Hoeflich, A.; Schmicke, M. Increased concentrations of insulin-like growth factor binding protein (IGFBP)-2, IGFBP-3, and IGFBP-4 are associated with fetal mortality in pregnant cows. Front. Endocrinol. 2018, 9, 310. [Google Scholar] [CrossRef]
  51. Fajner, V.; Giavazzi, F.; Sala, S.; Oldani, A.; Martini, E.; Napoletano, F.; Parazzoli, D.; Cesare, G.; Cerbino, R.; Maspero, E. Hecw controls oogenesis and neuronal homeostasis by promoting the liquid state of ribonucleoprotein particles. Nat. Commun. 2021, 12, 5488. [Google Scholar] [CrossRef] [PubMed]
  52. Soeda, S.; Oyama, M.; Kozuka-Hata, H.; Yamamoto, T. The CCR4-NOT complex suppresses untimely translational activation of maternal mRNAs. Development 2023, 150, dev201773. [Google Scholar] [CrossRef] [PubMed]
  53. Xu, B.; Yang, L.; Hinton, B.T. The Role of fibroblast growth factor receptor substrate 2 (FRS2) in the regulation of two activity levels of the components of the extracellular signal-regulated kinase (ERK) pathway in the mouse epididymis. Biol. Reprod. 2013, 89, 48. [Google Scholar] [CrossRef]
  54. Seaward, A.V.; Burke, S.D.; Croy, B.A. Interferon gamma contributes to preimplantation embryonic development and to implantation site structure in NOD mice. Hum. Reprod. 2010, 25, 2829–2839. [Google Scholar] [CrossRef] [PubMed]
  55. Lee, M.O.; Li, J.; Davis, B.W.; Upadhyay, S.; Al Muhisen, H.M.; Suva, L.J.; Clement, T.M.; Andersson, L. Hmga2 deficiency is associated with allometric growth retardation, infertility, and behavioral abnormalities in mice. G3 2022, 12, jkab417. [Google Scholar] [CrossRef]
  56. Grewal, S.; Carver, J.; Ridley, A.J.; Mardon, H.J. Human endometrial stromal cell Rho GTPases have opposing roles in regulating focal adhesion turnover and embryo invasion in vitro. Biol. Reprod. 2010, 83, 75–82. [Google Scholar] [CrossRef]
  57. Lin, S.; Yang, M.; Zhu, W.; Yang, C.; Chen, Y.; Cong, P.; Liu, X.; He, Z. Heterozygous deletion of exon 17 of the Kit gene impairs mouse spermatogenesis by attenuating MAPK-ERK signaling. Biol. Res. 2025, 58, 28. [Google Scholar] [CrossRef]
  58. Liu, Z.K.; Wang, R.C.; Han, B.C.; Yang, Y.; Peng, J.P. A novel role of IGFBP7 in mouse uterus: Regulating uterine receptivity through Th1/Th2 lymphocyte balance and decidualization. PLoS ONE 2012, 7, e45224. [Google Scholar] [CrossRef]
  59. Buensuceso, A.V.; Son, A.I.; Zhou, R.P.; Paquet, M.; Withers, B.M.; Deroo, B.J. Ephrin-A5 is required for optimal fertility and a complete ovulatory response to gonadotropins in the female mouse. Endocrinology 2016, 157, 942–955. [Google Scholar] [CrossRef]
  60. Horvat, F.; Fulka, H.; Jankele, R.; Malik, R.; Jun, M.; Solcova, K.; Sedlacek, R.; Vlahovicek, K.; Schultz, R.M.; Svoboda, P. Role of in maternal mRNA turnover. Life Sci. Alliance 2018, 1, e201800084. [Google Scholar] [CrossRef]
  61. Zhao, X.; Fan, C.; Qie, T.; Fu, X.; Chen, X.; Wang, Y.; Wu, Y.; Fu, X.; Shi, K.; Yan, W.; et al. Diaph1 knockout inhibits mouse primordial germ cell proliferation and affects gonadal development. Reprod. Biol. Endocrinol. 2024, 22, 82. [Google Scholar] [CrossRef]
  62. Uzumcu, M.; Dirks, K.A.; Skinner, M.K. Inhibition of platelet-derived growth factor actions in the embryonic testis influences normal cord development and morphology. Biol. Reprod. 2002, 66, 745–753. [Google Scholar] [CrossRef]
  63. Du, R.H.; Chen, H.Y.; Gao, L. Roles of protein tyrosine phosphatases in reproduction and related diseases. Reprod. Dev. Med. 2023, 7, 252–256. [Google Scholar] [CrossRef]
  64. Li, J.; Zhang, S.; Shen, C.; Niu, Z.; Yang, H.; Zhang, K.; Liu, Z.; Wang, Y.; Lan, X. Indel mutations within the bovine HSD17B3 gene are significantly associated with ovary morphological traits and mature follicle number. J. Steroid Biochem. Mol. Biol. 2021, 209, 105833. [Google Scholar] [CrossRef]
  65. Nyegaard, M.; Overgaard, M.T.; Su, Y.Q.; Hamilton, A.E.; Kwintkiewicz, J.; Hsieh, M.; Nayak, N.R.; Conti, M.; Conover, C.A.; Giudice, L.C. Lack of functional pregnancy-associated plasma protein-A (PAPPA) compromises mouse ovarian steroidogenesis and female fertility. Biol. Reprod. 2010, 82, 1129–1138. [Google Scholar] [CrossRef]
  66. Ebrahimi, A.; Ghavi, D.; Mirzaei, Z.; Barati, T.; Mansoori, S. Novel differentially expressed male infertility-associated genes in sperm as prospective diagnostic biomarkers. Hum. Gene 2024, 39, 201255. [Google Scholar] [CrossRef]
  67. Wang, Q.; Chikina, M.D.; Pincas, H.; Sealfon, S.C. Homer1 alternative splicing is regulated by gonadotropin-releasing hormone and modulates gonadotropin gene expression. Mol. Cell Biol. 2014, 34, 1747–1756. [Google Scholar] [CrossRef]
  68. Shi, J.X.; Yang, L.; Gan, J.; Gu, W.W.; Gu, Y.; Shi, Y.; Jiang, H.Y.; Xu, H.R.; Yang, S.H.; Zhang, X.; et al. MiR-3074-5p regulates trophoblasts function via EIF2S1/GDF15 pathway in recurrent miscarriage. Reprod. Sci. 2024, 31, 1290–1302. [Google Scholar] [CrossRef]
  69. Stavros, S.; Potiris, A.; Mavrogianni, D.; Moustakli, E.; Tsiorou, K.; Zikopoulos, A.; Kathopoulis, N.; Theofanakis, C.; Loutradis, D.; Domali, E.; et al. Exploring the potential impact of SERPINE gene expression in cumulus cells during fertility treatments: A single center study. Clin. Pract. 2025, 15, 83. [Google Scholar] [CrossRef]
  70. Filis, P.; Kind, P.C.; Spears, N. Implantation failure in mice with a disruption in Phospholipase C beta 1 gene: Lack of embryonic attachment, aberrant steroid hormone signalling and defective endocannabinoid metabolism. Mol. Hum. Reprod. 2013, 19, 290–301. [Google Scholar] [CrossRef]
  71. Svendsen, J.M.; Smogorzewska, A.; Sowa, M.E.; O’Connell, B.C.; Gygi, S.P.; Elledge, S.J.; Harper, J.W. Mammalian BTBD12/SLX4 assembles a holliday junction resolvase and is required for DNA repair. Cell 2009, 138, 63–77. [Google Scholar] [CrossRef]
  72. Jungnickel, M.K.; Sutton, K.A.; Baker, M.A.; Cohen, M.G.; Sanderson, M.J.; Florman, H.M. The flagellar protein Enkurin is required for mouse sperm motility and for transport through the female reproductive tract. Biol. Reprod. 2018, 99, 789–797. [Google Scholar] [CrossRef]
  73. Solloway, M.J.; Robertson, E.J. Early embryonic lethality in Bmp5; Bmp7 double mutant mice suggests functional redundancy within the 60A subgroup. Development 1999, 126, 1753–1768. [Google Scholar] [CrossRef]
  74. Percharde, M.; Lavial, F.; Ng, J.H.; Kumar, V.; Tomaz, R.A.; Martin, N.; Yeo, J.C.; Gil, J.; Prabhakar, S.; Ng, H.H.; et al. Ncoa3 functions as an essential Esrrb coactivator to sustain embryonic stem cell self-renewal and reprogramming. Gene Dev. 2012, 26, 2286–2298. [Google Scholar] [CrossRef] [PubMed]
  75. Saha, S.; Sucharita, S.; Majhi, R.K.; Tiwari, A.; Ghosh, A.; Pradhan, S.K.; Patra, B.K.; Dash, R.R.; Nayak, R.N.; Giri, S.C.; et al. TRPA1 is selected as a semi-conserved channel during vertebrate evolution due to its involvement in spermatogenesis. Biochem. Biophys. Res. Commun. 2019, 512, 295–302. [Google Scholar] [CrossRef]
  76. Pohoczky, K.; Kun, J.; Szalontai, B.; Szoke, E.; Saghy, E.; Payrits, M.; Kajtar, B.; Kovacs, K.; Kornyei, J.L.; Garai, J.; et al. Estrogen-dependent up-regulation of TRPA1 and TRPV1 receptor proteins in the rat endometrium. J. Mol. Endocrinol. 2016, 56, 135–149. [Google Scholar] [CrossRef]
  77. Wang, S.; Yang, H.; Fu, Y.; Teng, X.; Wang, C.; Xu, W. The Key Role of Peroxisomes in Follicular Growth, Oocyte Maturation, Ovulation, and Steroid Biosynthesis. Oxid. Med. Cell Longev. 2022, 2022, 7982344. [Google Scholar] [CrossRef]
  78. Nakayama, M.; Sato, H.; Okuda, T.; Fujisawa, N.; Kono, N.; Arai, H.; Suzuki, E.; Umeda, M.; Ishikawa, H.O.; Matsuno, K. Drosophila carrying pex3 or pex16 mutations are models of Zellweger syndrome that reflect its symptoms associated with the absence of peroxisomes. PLoS ONE 2011, 6, e22984. [Google Scholar] [CrossRef]
  79. Asadollahpour Nanaei, H.; Ayatollahi Mehrgardi, A.; Esmailizadeh, A. Whole-genome sequence analysis reveals candidate genomic footprints and genes associated with reproductive traits in Thoroughbred horse. Reprod. Domest. Anim. 2020, 55, 200–208. [Google Scholar] [CrossRef]
  80. Lavagi, I.; Krebs, S.; Simmet, K.; Beck, A.; Zakhartchenko, V.; Wolf, E.; Blum, H. Single-cell RNA sequencing reveals developmental heterogeneity of blastomeres during major genome activation in bovine embryos. Sci. Rep-UK 2018, 8, 4071. [Google Scholar] [CrossRef]
  81. Nykamp, K.; Lee, M.H.; Kimble, J. C. elegans La-related protein, LARP-1, localizes to germline P bodies and attenuates Ras-MAPK signaling during oogenesis. RNA 2008, 14, 1378–1389. [Google Scholar] [CrossRef]
  82. Borghei, A.; Ouyang, Y.-B.; Westmuckett, A.D.; Marcello, M.R.; Landel, C.P.; Evans, J.P.; Moore, K.L. Targeted disruption of tyrosylprotein sulfotransferase-2, an enzyme that catalyzes post-translational protein tyrosine O-sulfation, causes male infertility. J. Biol. Chem. 2006, 281, 9423–9431. [Google Scholar] [CrossRef] [PubMed]
  83. Xia, S.; Yu, W.; Menden, H.; Younger, S.T.; Sampath, V. FOXC2 autoregulates its expression in the pulmonary endothelium after endotoxin stimulation in a histone acetylation-dependent manner. Front. Cell Dev. Biol. 2021, 9, 657662. [Google Scholar] [CrossRef] [PubMed]
  84. Kinsler, V.A.; Abu-Amero, S.; Budd, P.; Jackson, I.J.; Ring, S.M.; Northstone, K.; Atherton, D.J.; Bulstrode, N.W.; Stanier, P.; Hennekam, R.C.; et al. Germline melanocortin-1-receptor genotype is associated with severity of cutaneous phenotype in congenital melanocytic nevi: A role for MC1R in human fetal development. J. Investig. Dermatol. 2012, 132, 2026–2032. [Google Scholar] [CrossRef]
  85. Chambers, I.G.; Kumar, P.; Lichtenberg, J.; Wang, P.; Yu, J.; Phillips, J.D.; Kane, M.A.; Bodine, D.; Hamza, I. MRP5 and MRP9 play a concerted role in male reproduction and mitochondrial function. Proc. Natl. Acad. Sci. USA 2022, 119, e2111617119. [Google Scholar] [CrossRef]
  86. Olbrich, H.; Schmidts, M.; Werner, C.; Onoufriadis, A.; Loges, N.T.; Raidt, J.; Banki, N.F.; Shoemark, A.; Burgoyne, T.; Al Turki, S. Recessive HYDIN mutations cause primary ciliary dyskinesia without randomization of left-right body asymmetry. Am. J. Hum. Genet. 2012, 91, 672–684. [Google Scholar] [CrossRef]
  87. Liang, Z.; Prakapenka, D.; Zaabza, H.B.; VanRaden, P.M.; Van Tassell, C.P.; Da, Y. A million-cow genome-wide association study of productive life in US Holstein cows. Genet. Sel. Evol. 2024, 56, 67. [Google Scholar] [CrossRef]
  88. Azhar, M.; Xu, C.; Jiang, X.; Li, W.; Cao, Y.; Zhu, X.; Xing, X.; Wu, L.; Zou, J.; Meng, L.; et al. The arginine methyltransferase Prmt1 coordinates the germline arginine methylome essential for spermatogonial homeostasis and male fertility. Nucleic Acids Res. 2023, 51, 10428–10450. [Google Scholar] [CrossRef]
  89. Hashimoto, M.; Fukamizu, A.; Nakagawa, T.; Kizuka, Y. Roles of protein arginine methyltransferase 1 (PRMT1) in brain development and disease. Biochim. Biophys. Acta Gen. Subj. 2021, 1865, 129776. [Google Scholar] [CrossRef]
  90. Stewart, M.K.; Mattiske, D.M.; Pask, A.J. Estrogen suppresses SOX9 and activates markers of female development in a human testis-derived cell line. BMC Mol. Cell Biol. 2020, 21, 66. [Google Scholar] [CrossRef]
  91. Warr, N.; Siggers, P.; Carre, G.A.; Wells, S.; Greenfield, A. Genetic analyses reveal functions for MAP2K3 and MAP2K6 in mouse testis determination. Biol. Reprod. 2016, 94, 103. [Google Scholar] [CrossRef]
  92. Wang, M.; Huang, Y.P.; Wu, H.; Song, K.; Wan, C.; Chi, A.N.; Xiao, Y.M.; Zhao, X.Y. Mitochondrial complex I deficiency leads to the retardation of early embryonic development in knockout mice. PeerJ 2017, 5, e3339. [Google Scholar] [CrossRef]
  93. Yeh, J.; Kim, B.S.; Gaines, L.; Peresie, J.; Page, C.; Arroyo, A. The expression of hyperpolarization activated cyclic nucleotide gated (HCN) channels in the rat ovary are dependent on the type of cell and the reproductive age of the animal: A laboratory investigation. Reprod. Biol. Endocrinol. 2008, 6, 35. [Google Scholar] [CrossRef]
  94. Bachelot, A.; Monget, P.; Imbert-Bolloré, P.; Coshigano, K.; Kopchick, J.J.; Kelly, P.A.; Binart, N. Growth hormone is required for ovarian follicular growth. Endocrinology 2002, 143, 4104–4112. [Google Scholar] [CrossRef] [PubMed]
  95. Li, Z.; Zheng, Z.; Ruan, J.; Li, Z.; Zhuang, X.; Tzeng, C.M. Integrated analysis miRNA and mRNA profiling in patients with severe oligozoospermia reveals miR-34c-3p downregulates PLCXD3 expression. Oncotarget 2016, 7, 52781–52796. [Google Scholar] [CrossRef] [PubMed]
  96. Terakawa, J.; Nakamura, S.; Ohtomo, M.; Uehara, S.; Kawata, Y.; Takarabe, S.; Sugita, H.; Namiki, T.; Kageyama, A.; Noguchi, M. LIFR-mediated ERBB2 signaling is essential for successful embryo implantation in mice. Biomolecules 2025, 15, 698. [Google Scholar] [CrossRef]
  97. Nagirnaja, L.; Palta, P.; Kasak, L.; Rull, K.; Christiansen, O.B.; Nielsen, H.S.; Steffensen, R.; Esko, T.; Remm, M.; Laan, M. Structural genomic variation as risk factor for idiopathic recurrent miscarriage. Hum. Mutat. 2014, 35, 972–982. [Google Scholar] [CrossRef]
  98. Zhang, C.; Long, X.; Ding, Y.; Chen, X.; He, J.; Liu, S.; Geng, Y.; Wang, Y.; Liu, X. Expression of DROSHA in the uterus of mice in early pregnancy and its potential significance during embryo implantation. Reprod. Sci. 2016, 23, 154–162. [Google Scholar] [CrossRef]
  99. Matsunaga, E.; Nambu, S.; Oka, M.; Iriki, A. Complex and dynamic expression of cadherins in the embryonic marmoset cerebral cortex. Dev. Growth Differ. 2015, 57, 474–483. [Google Scholar] [CrossRef]
  100. Golawski, K.; Soczewica, R.; Kacperczyk-Bartnik, J.; Manka, G.; Kiecka, M.; Lipa, M.; Warzecha, D.; Spaczynski, R.; Piekarski, P.; Banaszewska, B.; et al. The role of cadherin 12 (CDH12) in the peritoneal fluid among patients with endometriosis and endometriosis-related infertility. Int. J. Environ. Res. Public Health 2022, 19, 11586. [Google Scholar] [CrossRef] [PubMed]
  101. Hong, X.M.; Luense, L.J.; McGinnis, L.K.; Nothnick, W.B.; Christenson, L.K. Dicer1 is essential for female fertility and normal development of the female reproductive system. Endocrinology 2008, 149, 6207–6212. [Google Scholar] [CrossRef]
  102. Talbot, C.D.; Walsh, M.D.; Cutty, S.J.; Elsayed, R.; Vlachaki, E.; Bruce, A.E.E.; Wardle, F.C.; Nelson, A.C. Eomes function is conserved between zebrafish and mouse and controls left-right organiser progenitor gene expression via interlocking feedforward loops. Front. Cell Dev. Biol. 2022, 10, 982477. [Google Scholar] [CrossRef]
  103. Akdemir, N.; Cinemre, F.B.; Cinemre, H.; Sevinc, L.; Aydemir, B.; Coban, B.; Cevrioglu, A.S.; Ozden, S. Polymorphism of the Oxytocin Receptor (OXTR) gene affects the circulating oxytocin receptor levels in late-term pregnancy in a Turkish population. Gynecol. Obstet. Investig. 2020, 85, 343–351. [Google Scholar] [CrossRef]
  104. Jin, X.H.; Zhang, S.; Ding, T.B.; Zhao, P.T.; Zhang, C.L.; Zhang, Y.X.; Li, W. Testicular Lmcd1 regulates phagocytosis by Sertoli cells through modulation of NFAT1/Txlna signaling pathway. Aging Cell 2020, 19, e13217. [Google Scholar] [CrossRef]
  105. Liu, Y.Y.; Jiang, M.; Li, C.; Yang, P.; Sun, H.Q.; Tao, D.C.; Zhang, S.H.; Ma, Y.X. Human t-complex protein 11 (TCP11), a testis-specific gene Product, is a potential determinant of the sperm morphology. Tohoku J. Exp. Med. 2011, 224, 111–117. [Google Scholar] [CrossRef]
  106. Fraser, L.R.; Hosseini, R.; Hanyalogou, A.; Talmor, A.; Dudley, R.K. TCP-11, the product of a mouse t-complex gene, plays a role in stimulation of capacitation and inhibition of the spontaneous acrosome reaction. Mol. Reprod. Dev. Inc. Gamete Res. 1997, 48, 375–382. [Google Scholar] [CrossRef]
  107. Serber, D.W.; Runge, J.S.; Menon, D.U.; Magnuson, T. The Mouse INO80 chromatin-remodeling complex is an essential meiotic factor for spermatogenesis. Biol. Reprod. 2016, 94, 8. [Google Scholar] [CrossRef]
  108. Park, K.H.; Ye, Z.W.; Zhang, J.; Hammad, S.M.; Townsend, D.M.; Rockey, D.C.; Kim, S.H. 3-ketodihydrosphingosine reductase mutation induces steatosis and hepatic injury in zebrafish. Sci. Rep. 2019, 9, 1138. [Google Scholar] [CrossRef]
  109. Janesick, A.; Tang, W.Y.; Nguyen, T.T.L.; Blumberg, B. RARβ2 is required for vertebrate somitogenesis. Development 2017, 144, 1997–2008. [Google Scholar] [CrossRef] [PubMed]
  110. Ai, N.; Han, C.R.; Zhao, H.; Cheng, S.Y.; Ge, W. Disruption of thyroid hormone receptor thrab leads to female infertility in Zebrafish. Endocrinology 2024, 165, bqae037. [Google Scholar] [CrossRef] [PubMed]
  111. Jee, Y.H.; Won, S.; Lui, J.C.; Jennings, M.; Whalen, P.; Yue, S.; Temnycky, A.G.; Barnes, K.M.; Cheetham, T.; Boden, M.G.; et al. DLG2 variants in patients with pubertal disorders. Genet. Med. 2020, 22, 1329–1337. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Additive effects of DPR. (a) Manhattan plot of genome-wide additive effects, showing genome-wide distribution of additive effects and the lack of a sharply significant chromosome region; (b) positive and negative allelic effects of the top 1000 most significant additive effects; (c) allelic effects of Chr01 among the top 1000 additive effects; (d) allelic effects of Chr04 among the top 1000 additive effects; (e) allelic effects of Chr13 among the top 1000 additive effects; (f) allelic effects of Chr24 among the top 1000 additive effects. ‘ae+’ is the allelic effect of the positive allele. ‘ae-’ is the allelic effect of the negative allele. Green double-arrowed line indicates larger effect size of the positive allele than the negative allele. Red double-arrowed line indicates larger effect size of the negative allele than the positive allele.
Figure 1. Additive effects of DPR. (a) Manhattan plot of genome-wide additive effects, showing genome-wide distribution of additive effects and the lack of a sharply significant chromosome region; (b) positive and negative allelic effects of the top 1000 most significant additive effects; (c) allelic effects of Chr01 among the top 1000 additive effects; (d) allelic effects of Chr04 among the top 1000 additive effects; (e) allelic effects of Chr13 among the top 1000 additive effects; (f) allelic effects of Chr24 among the top 1000 additive effects. ‘ae+’ is the allelic effect of the positive allele. ‘ae-’ is the allelic effect of the negative allele. Green double-arrowed line indicates larger effect size of the positive allele than the negative allele. Red double-arrowed line indicates larger effect size of the negative allele than the positive allele.
Ijms 26 11149 g001
Figure 2. Dominance effects and the homozygous advantage and heterozygous disadvantage of significant dominance effects of DPR. (a) Manhattan plot of genome-wide dominance effects, showing that the MC1R gene had a far more significant dominance effect than in other regions; (b) dominance values of top-ranked significant dominance effects; (c) dominance values of bottom-ranked significant dominance effects; (d) examples of SNP genotypic mean of phenotypic values. dij is the dominance value of the ij SNP genotype, ij = 1,2. yij is the SNP genotypic mean of the original phenotypic values without removing pedigree additive (breeding) values. gij is the SNP genotypic mean of the corrected phenotypic values after removing pedigree additive (breeding) values. µy is the population mean of the original phenotypic values without removing pedigree additive (breeding) values. µg is the population mean of the corrected phenotypic values after removing pedigree additive (breeding) values.
Figure 2. Dominance effects and the homozygous advantage and heterozygous disadvantage of significant dominance effects of DPR. (a) Manhattan plot of genome-wide dominance effects, showing that the MC1R gene had a far more significant dominance effect than in other regions; (b) dominance values of top-ranked significant dominance effects; (c) dominance values of bottom-ranked significant dominance effects; (d) examples of SNP genotypic mean of phenotypic values. dij is the dominance value of the ij SNP genotype, ij = 1,2. yij is the SNP genotypic mean of the original phenotypic values without removing pedigree additive (breeding) values. gij is the SNP genotypic mean of the corrected phenotypic values after removing pedigree additive (breeding) values. µy is the population mean of the original phenotypic values without removing pedigree additive (breeding) values. µg is the population mean of the corrected phenotypic values after removing pedigree additive (breeding) values.
Ijms 26 11149 g002
Figure 3. Manhattan plot of A × A epistasis effects. The significant A × A epistasis effects were distributed on all chromosomes but lacked chromosome regions with A × A effects far more significant than in other regions.
Figure 3. Manhattan plot of A × A epistasis effects. The significant A × A epistasis effects were distributed on all chromosomes but lacked chromosome regions with A × A effects far more significant than in other regions.
Ijms 26 11149 g003
Figure 4. Circular plot of intra-chromosome A × A epistasis effects for six selected chromosomes. (a) The 1.08–3.71 Mb region of Chr13 had 31 pairs of A × A epistasis effects with the PLCB1 and SLX4IP genes at the two ends of this region; (b) the 34–74 Mb region of Chr05 had a large cluster of A × A effects; (c) the 31–51 Mb region of Chr14 had a large cluster of A × A effects; (d) the 22–32 Mb region of Chr17 had a large cluster of A × A effects; (e) Chr01 had A × A effects covering over half of the chromosome; (f) Chr03 had A × A effects covering over half of the chromosome.
Figure 4. Circular plot of intra-chromosome A × A epistasis effects for six selected chromosomes. (a) The 1.08–3.71 Mb region of Chr13 had 31 pairs of A × A epistasis effects with the PLCB1 and SLX4IP genes at the two ends of this region; (b) the 34–74 Mb region of Chr05 had a large cluster of A × A effects; (c) the 31–51 Mb region of Chr14 had a large cluster of A × A effects; (d) the 22–32 Mb region of Chr17 had a large cluster of A × A effects; (e) Chr01 had A × A effects covering over half of the chromosome; (f) Chr03 had A × A effects covering over half of the chromosome.
Ijms 26 11149 g004
Figure 5. Complementary additive effects between two breeds. Capital letter represents favorable allele, and lower-case letter represents unfavorable or neutral allele. Green represents one breed, and blue represents the other breed. The green breed has nine favorable alleles, the blue breed has three favorable alleles, and the crossbred between the two breeds has 12 favorable alleles.
Figure 5. Complementary additive effects between two breeds. Capital letter represents favorable allele, and lower-case letter represents unfavorable or neutral allele. Green represents one breed, and blue represents the other breed. The green breed has nine favorable alleles, the blue breed has three favorable alleles, and the crossbred between the two breeds has 12 favorable alleles.
Ijms 26 11149 g005
Figure 6. A × A effects in crossbred cows are allelic interaction effects between two loci.
Figure 6. A × A effects in crossbred cows are allelic interaction effects between two loci.
Ijms 26 11149 g006
Figure 7. Example of dominance values (deviations) of crossbred cows from Jersey–Holstein cross. Jersey alleles are Aj and aj and Holstein alleles are Ah and ah. dij = dominance value (deviation), gij = genotypic value, aij = additive value, i, j = 1, 2, µ = population mean of the three genotypic values.
Figure 7. Example of dominance values (deviations) of crossbred cows from Jersey–Holstein cross. Jersey alleles are Aj and aj and Holstein alleles are Ah and ah. dij = dominance value (deviation), gij = genotypic value, aij = additive value, i, j = 1, 2, µ = population mean of the three genotypic values.
Ijms 26 11149 g007
Figure 8. Example of genotype × genotype interactions for different D × D values of nine two-locus SNP genotypes. The two alleles of each SNP assumed to originate from Jersey are indicated by subscript j and Holstein are indicated by subscript h. Every two-locus genotype with a heterozygous genotype at either SNP (highlighted in yellow) had a negative D × D value (highlighted in yellow), whereas the doubly heterozygous genotype (AaBb) and all homozygous genotypes had positive D × D values. DD1–DD9 are the nine D × D values of the most significant D × D effect from Table S5.
Figure 8. Example of genotype × genotype interactions for different D × D values of nine two-locus SNP genotypes. The two alleles of each SNP assumed to originate from Jersey are indicated by subscript j and Holstein are indicated by subscript h. Every two-locus genotype with a heterozygous genotype at either SNP (highlighted in yellow) had a negative D × D value (highlighted in yellow), whereas the doubly heterozygous genotype (AaBb) and all homozygous genotypes had positive D × D values. DD1–DD9 are the nine D × D values of the most significant D × D effect from Table S5.
Ijms 26 11149 g008
Table 1. Top 20 additive effects of daughter pregnancy rate (DPR).
Table 1. Top 20 additive effects of daughter pregnancy rate (DPR).
SNPChrPositionCandidate Geneαal+ae+f_al+al-ae-f_al-log10(1/p)
rs108997231133906114SLX4IP2.11610.966 0.543 2−1.150 0.457 30.49
rs110789800176375060LOC614619; GMNC−2.00520.655 0.673 1−1.350 0.327 27.00
rs34230934962629435606MIR2285K-4(d)−1.89720.847 0.553 1−1.050 0.447 26.69
rs1097789581729595964LARP1B1.92810.698 0.638 2−1.230 0.362 26.49
rs416384042025229625ARL15; NDUFS41.87010.750 0.600 2−1.120 0.400 26.30
rs1105601192622679137ARMH31.83510.735 0.599 2−1.100 0.401 25.19
rs43292973222416125CIR11.87010.790 0.577 2−1.080 0.423 25.10
rs111018678143308936TRAPPC9−1.91320.553 0.711 1−1.360 0.289 24.43
rs109798047105830788LOC100138889; LOC781028−1.84520.958 0.481 1−0.887 0.519 24.41
rs1105115541036374183PPP1R14D1.85610.991 0.466 2−0.865 0.534 24.33
rs1106003241575950150PEX161.86911.070 0.427 2−0.799 0.573 23.80
rs43438147543898744FRS2−1.78120.681 0.617 1−1.100 0.383 23.65
rs34231418182103153421ABCA12−1.81920.845 0.535 1−0.974 0.465 23.64
rs34231650113118672880LOC101907483; LOC1124460441.81110.651 0.640 2−1.160 0.360 23.63
rs43708865945754622LOC785087; LOC112448181−1.96920.589 0.701 1−1.380 0.299 23.59
rs29019531197643741LOC112448275; LOC107132196−1.71820.809 0.529 1−0.909 0.471 23.45
rs1100808472421456850GALNT1; INO80C−1.94420.674 0.653 1−1.270 0.347 23.40
rs43464313648853315LOC100298058; LOC1124471861.79010.710 0.603 2−1.080 0.397 23.37
rs3423092932141854457ADGRB11.75710.597 0.660 2−1.160 0.340 23.09
rs42438948177548169LRBA−1.88320.633 0.663 1−1.250 0.337 22.87
‘d’ indicates that the SNP is downstream of the gene. ‘α’ is the additive effect of the SNP as the difference between allelic effects of ‘allele 1’ and ‘allele 2’. ‘ae+’ is the allelic effect of the positive allele. ‘ae-’ is the allelic effect of the negative allele. ‘f_al+’ is the frequency of the positive allele. ‘f_al-’ is the frequency of the negative allele.
Table 2. Top 20 dominance effects of daughter pregnancy rate (DPR).
Table 2. Top 20 dominance effects of daughter pregnancy rate (DPR).
SNPChrPositionCandidate Geneδd_12f_12d_11f_11d_22f_22log10(1/p)
rs1096880131814705671MC1R−2.555 −1.280 0.421 0.629 0.445 1.920 0.134 18.19
rs109588756X56705838VSIG1−2.136 −0.846 0.529 1.850 0.092 0.730 0.379 11.83
rs1342579231079265033EIF2S1−1.973 −0.854 0.511 0.667 0.368 1.570 0.121 11.17
rs1356343362027116615-−1.826 −0.784 0.563 0.953 0.292 1.130 0.145 10.87
rs1096586032026870070-−1.873 −0.775 0.560 1.340 0.121 0.855 0.319 10.60
rs290188842027087053-−1.848 −0.767 0.559 0.841 0.320 1.320 0.121 10.54
rs1101478811220815304SERPINE3−1.751 −0.783 0.553 0.973 0.183 0.962 0.264 10.37
rs1353127622028614378EMB−1.988 −0.788 0.511 1.790 0.088 0.610 0.402 10.35
rs1110048992035745330LOC104975274; LOC101906686−1.742 −0.750 0.567 1.020 0.151 0.964 0.282 10.03
rs1095132932947636529SHANK2−1.714 −0.740 0.570 0.857 0.201 1.090 0.230 9.92
rs416081721327232535TRNAE-UUC_63 (d)−1.694 −0.752 0.554 0.971 0.168 0.912 0.278 9.83
rs1096288242027104916-−1.791 −0.745 0.559 0.821 0.320 1.270 0.121 9.78
rs1096804561242526721LOC107132995 (d)−1.693 −0.713 0.571 0.660 0.235 1.300 0.193 9.74
rs41653703476449707RAMP3−1.677 −0.729 0.567 0.924 0.165 0.972 0.268 9.63
rs109588503476953836NUDCD3−1.675 −0.727 0.568 0.922 0.165 0.973 0.268 9.59
rs1337063731856085584PRMT1−1.723 −0.761 0.543 0.803 0.308 1.120 0.148 9.58
rs416326421580951917SELENOH−1.651 −0.754 0.544 0.950 0.245 0.844 0.211 9.53
rs418894351856040467PRR12; RRAS−1.659 −0.747 0.549 0.905 0.261 0.918 0.189 9.46
rs41654346672581115IGFBP7; TRNAK-UUU_12−1.681 −0.754 0.536 0.773 0.314 1.080 0.150 9.40
rs1094892281078362097LOC112448513−1.930 −0.786 0.512 1.670 0.097 0.617 0.391 9.39
‘d’ indicates that the SNP is downstream of the gene. ‘δ’ is the dominance effect of the SNP as the difference between the heterozygous dominance value and the average of the two homozygous dominance values. ‘d_ij’ is the dominance value (deviation) of the ij SNP genotype. ‘f_ij’ is the frequency of the ij genotype, i = 1, 2 and j = 1, 2.
Table 3. Top 20 intra-chromosome A × A effects.
Table 3. Top 20 intra-chromosome A × A effects.
SNP-1ChrPosition-1Candidate Gene-1SNP-2Position-2Candidate Gene-2ααlog10(1/p)
rs1361646852217954298LMCD1rs10913354517997196LOC1124435042.880 24.42
rs34230933201435595531LOC101903752; TRPA1rs342309361335751572TRPA1; LOC1124495773.290 21.27
rs415797851914094861ENSBTAG00000053331rs342346424214110279LOC104969035; HEATR63.490 21.09
rs34233634501435558190LOC101903752; TRPA1rs342309361335751572TRPA1; LOC1124495773.290 21.07
rs1334912401723169353LOC783956 (u)rs11034866625067804PCDH102.600 20.97
rs111002920327938935LOC782058; NGFrs11002097828078792NGF; LOC1124459673.410 20.52
rs416566971010474235HOMER1rs2901995310520078HOMER1−2.350 20.16
rs429828332045269875LOC112443070 (u)rs13761938945663409LOC112443070; LOC1124429553.330 20.04
rs1098742401142636491TFF1; TMPRSS3rs42141252142787751SLC37A1−2.810 19.93
rs109520109542304422CPNE8rs4343696643506809LOC104972423; MYRFL−2.700 19.76
rs1340520041440390594PEX2; LOC101904449rs4162593440600797LOC101904449; LOC782385−3.830 19.75
rs42445061360310725TTLL7(d)rs13421259560964021-2.470 19.70
rs111002920327938935LOC782058; NGFrs10989672028047336NGF; LOC1124459673.430 19.61
rs109621977348432235ALG14rs10984966049515221ABCA4−2.390 19.21
rs109619154298688612LOC101905908; PRSS23rs1100888079023536ME32.920 19.09
rs1350722521840169055HYDINrs11078571740219814HYDIN; LOC1124424002.730 19.09
rs109367522542871428PTPRRrs4279135143371061LOC104972423; MYRFL−2.700 19.09
rs110991306294592808GPR1; ZDBF2rs11097597995586874LOC112442377; LOC1008476662.360 19.06
rs431834102045022442LOC112443070 (u)rs13761938945663409LOC112443070; LOC112442955−3.230 19.02
rs34230933201435595531LOC101903752; TRPA1rs342309391435815088LOC112449577; LOC1124495162.340 18.83
‘u’ indicates the SNP is upstream of the gene. ‘αα’ is the A × A effect of the SNP pair.
Table 4. Significant D × D effects.
Table 4. Significant D × D effects.
SNP-1Chr-1Position-1Candidate Gene-1SNP-2Chr-2Position-2Candidate Gene-2δδlog10(1/p)
rs426702202039445232RAI14rs419423662041041011LOC104975283; SUB15.010 15.33
rs110573919253828387DNAAF8rs134957967253920360UBN14.350 14.32
rs1092892431766235451TPST2rs1092552411766309374LOC100847159; LOC6148814.710 14.28
rs1326977511325821584ENKURrs1093528191325911339GPR1584.750 14.25
rs419421222034603769LOC782462; LOC112443045rs1096338972034759872LOC782462; LOC1124430455.230 13.94
rs418747081832330449LOC112442399; CDH11rs416372821832771279CDH114.950 13.90
rs13291204446268973LOC101904266 (d)rs342316410546379484LOC101904266 (d)4.600 13.87
rs1104050281814014113APRTrs1101805801814050260CBFA2T34.500 13.48
rs41594638775473623LOC100296952; LOC112447598rs41667214776297401LOC112447528 (u)4.250 13.29
rs426401572058364680ANKHrs1108407252058566498OTULINL3.930 12.93
rs1090141482033024411PLCXD3rs430018582033323766C65.010 12.77
rs432784531142383844UMODL1rs1361386901143444786CRYAA; LOC1124482674.130 12.72
rs419513352058357512ANKHrs1108407252058566498OTULINL3.900 12.68
rs1340108352235456434MAGI1rs1107765232235535752MAGI14.950 12.53
rs418942821854627066BICRA; EHD2rs34234547871854680124SELENOW; LOC1019027665.650 12.52
rs1351766711144551556TSPEARrs1098400081146320854PRMT24.030 12.37
rs137324140132145874PLCB4rs134403374132232129ENSBTAG000000541744.470 12.14
rs34234491472067492982LOC104969150; ICE1rs419641972068697504LOC112443018 (d)4.880 12.03
‘d’ indicates the SNP is downstream of the gene. ‘u’ indicates the SNP is upstream of the gene. ‘δδ’ is the D × D effect of the SNP pair.
Table 5. Daughter pregnancy rate (DPR) of crossbred, Holstein, and Jersey cows based on the CDCB first lactation DPR data of 2015–2023.
Table 5. Daughter pregnancy rate (DPR) of crossbred, Holstein, and Jersey cows based on the CDCB first lactation DPR data of 2015–2023.
CrossbredJerseyHolstein
Mean54.9703 ± 33.17853.434 ± 33.19949.06 ± 32.01
n31,338722,9494,956,093
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

Yang, R.; Liang, Z.; Prakapenka, D.; Ma, L.; Da, Y. Genome-Wide Association Study of Daughter Pregnancy Rate in Crossbred Dairy Cows. Int. J. Mol. Sci. 2025, 26, 11149. https://doi.org/10.3390/ijms262211149

AMA Style

Yang R, Liang Z, Prakapenka D, Ma L, Da Y. Genome-Wide Association Study of Daughter Pregnancy Rate in Crossbred Dairy Cows. International Journal of Molecular Sciences. 2025; 26(22):11149. https://doi.org/10.3390/ijms262211149

Chicago/Turabian Style

Yang, Ruifei, Zuoxiang Liang, Dzianis Prakapenka, Li Ma, and Yang Da. 2025. "Genome-Wide Association Study of Daughter Pregnancy Rate in Crossbred Dairy Cows" International Journal of Molecular Sciences 26, no. 22: 11149. https://doi.org/10.3390/ijms262211149

APA Style

Yang, R., Liang, Z., Prakapenka, D., Ma, L., & Da, Y. (2025). Genome-Wide Association Study of Daughter Pregnancy Rate in Crossbred Dairy Cows. International Journal of Molecular Sciences, 26(22), 11149. https://doi.org/10.3390/ijms262211149

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