- Research
- Open access
- Published:
Influence of MHC on genetic diversity and testicular expression of linked olfactory receptor genes
BMC Genomics volume 26, Article number: 115 (2025)
Abstract
Background
Olfactory receptor (OR) genes are highly polymorphic and form extensive families that recognize a wide range of vertebrate odorants. To explore the genetic diversity of MHC-linked OR genes and their connection to MHC genes, we conducted a combined haplotype analysis of MHC-linked OR and MHC class I genes to determine the influence of MHC on OR diversity, which could be associated with MHC-based mate selection.
Results
We selected nine MHC-linked OR genes based on their expression levels in pig testes and developed a sequence-based typing method for these genes. We then performed high-resolution typing of these OR genes, along with three major classical MHC class I genes (SLA-1, -2, and − 3), in 48 pigs across six breeds. We observed significantly higher allelic diversity (P < 0.01) in ORs with strong linkage disequilibrium (LD) to SLA compared to those with weak or no LD, and we identified 48 SLA class I-OR haplotypes using the expectation-maximization algorithm. The genetic diversity of SLA-linked ORs was positively correlated with their expression levels in the testis. Specifically, SLA-linked ORs with higher testicular expression (FPKM ≥ 0.1) exhibited an increase in the number of codons under mutually diversifying selection with SLA compared to those with lower expression (FPKM < 0.1).
Conclusions
The presence of evolutionary interactions between MHC and linked OR genes supports the potential involvement of MHC-linked ORs in MHC-based mate selection. The use of combined haplotype information for MHC and linked ORs could provide new insights into the reproductive biology of animals.
Background
Olfactory receptors (ORs) are G protein-coupled receptors (GPCRs) responsible for odorant recognition, primarily expressed in olfactory neurons [1]. OR genes are highly diverse, forming extensive gene families that recognize a wide range of odorants and are organized into clusters across multiple chromosomes in the vertebrate genome [2,3,4,5]. The number of OR genes and the extent of pseudogenization vary significantly among species, reflecting the importance of olfaction for their survival. For example, pigs (Sus scrofa), cattle (Bos taurus), mice (Mus musculus), and humans (Homo sapiens) possess 1,301 (14.5% pseudogenes), 1,071 (17.7% pseudogenes), 1,483 (23.1% pseudogenes), and 874 OR genes (55.5% pseudogenes), respectively [6,7,8]. Notably, marine mammals exhibit a reduced OR repertoire compared to terrestrial animals [9,10,11,12].
ORs are also implicated in functions beyond odor recognition and olfactory signal transduction, as they are expressed in a variety of tissues and cells, including the brain, tongue, heart, spleen, and pancreas [13,14,15]. Noncanonical OR expression has been associated with the modulation of cell-cell recognition, migration, proliferation, apoptosis, exocytosis, and pathfinding processes [16,17,18,19,20]. Furthermore, several ORs are expressed in the testes and sperm cells, where they detect specific odorant molecules that trigger Ca2+ signals, which are crucial for changes in sperm motility and chemotaxis [21,22,23].
Intriguingly, ORs expressed in the testes are genetically linked to highly variable major histocompatibility complexes (MHC) and GABBR1, which are involved in recognizing foreign antigens for the adaptive immune response and play a role in sperm acrosome reactions, respectively [15, 24,25,26,27]. The conserved architecture of OR-GABBR1-MHC genes across vertebrate species may contribute to MHC genotype-based mating preferences, favoring individuals with dissimilar MHC genotypes [28,29,30,31,32]. However, our understanding of this phenomenon remains limited. Some studies have reported that MHC-OR haplotypes are highly variable and that specific OR-MHC allele combinations occur more frequently than expected by chance [33, 34]. Although ORs expressed in the testes appear to be genetically linked to the MHC, limited information is available on the diversity of MHC-linked ORs compared to ORs located in other chromosomal regions. Comparing the effects of different MHC-OR haplotypes on reproduction may provide new insights into mate selection. Consequently, new breeding strategies could be designed accordingly.
To elucidate the genetic and functional relationships between MHC and MHC-linked ORs, we analyzed the evolutionary conservation and linkage disequilibrium (LD) between MHC and MHC-linked ORs by typing nine MHC-linked ORs and three classical MHC class I genes in pigs. We performed haplotype analysis and compared allelic diversity between MHC-linked and unlinked ORs to assess the influence of MHC genes on OR diversity. Additionally, we examined the testicular expression patterns of orthologous MHC-linked ORs in pigs and humans using transcriptome data. Our study provides interesting insights into the evolutionary and phenotypic correlations between mammalian MHC and linked OR genes, suggesting that these findings may have implications for improving reproductive efficiency through the identification of ligands for MHC-linked ORs.
Results
Identification of MHC-linked orthologous OR genes across humans, mice, cattle, and pigs
The current genome assemblies for humans, mice, pigs, and cattle reveal that the numbers of OR genes mapped to MHC-residing chromosomes are 34, 67, 33, and 90, respectively (Supplementary Table 1). To determine the orthologous relationships of the ORs in the MHC-linked region among the four species, we identified 11 non-OR orthologous framework loci across these species using information from the NCBI Ortholog Database (www.ncbi.nlm.nih.gov/gene). These loci include ABT1, ZNF322, POM121L2, ZNF184, H1-5, ZSCAN26, ZSCAN12, GPX6, GABBR1, MOG, and ZFP57 (Fig. 1). Subsequently, we established the orthologous relationships of the ORs adjacent to the framework loci through interspecies pairwise BLAST analysis (Supplementary Table 2). OR gene pairs with e-values < 10− 100 (E-100) and bit scores > 500 between species, along with their syntenic relationships from the comparative analysis of framework loci, were identified as putative OR orthologs.
Synteny plot of the MHC and linked OR gene cluster region in mice, humans, pigs, and cattle. Chromosome numbers for each species including Mus musculus chromosome (MMU), Homo sapiens autosome (HSA), Sus scrofa chromosome (SSC), and Bos taurus autosome (BTA) are shown on the left. Eleven evolutionary conserved framework (orange) and OR (black) genes are indicated with vertical lines. The OR and MHC gene clusters are indicated with light green and light blue rectangles, respectively. The syntenic relationships are traced in gray curves. The nine pig MHC-linked OR genes typed in this study and their orthologous genes across species are marked in pink and connected with pink lines. OR gene symbols of LOC100514111, LOC100516618, LOC100157348, LOC100515036, LOC100156552, LOC100522686, LOC100516811, LOC615902, and LOC112443785 for pigs or cattle are shown only with the last three-digit numbers of the genes. The gene maps were depicted relative to the size (bp) of the region for each species except the MHC region (H2, HLA, SLA, and BoLA for mice, humans, pigs, and cattle, respectively). The scale bar is shown at the bottom
The number of OR genes within each OR cluster varied among species. OR genes with multiple matches in a given species from the BLAST analysis were considered to result from recent gene duplications. We observed three OR clusters on S. scrofa chromosome 7 (SSC7) and B. taurus chromosome 23 (BTA23), whereas H. sapiens chromosome 6 (HSA6) contained two clusters. An additional OR gene located between ABT1 and ZNF322 in Euungulata, but absent in other species, indicated an expansion of MHC-linked ORs within this lineage. In contrast, a single cluster of OR genes was present on M. musculus chromosome 17 (MMU17) due to the translocation of the OR cluster onto MMU13 (Fig. 1). Consequently, only a GABBR1-linked OR cluster was common to MHC-bearing chromosomes in all four species. The ORs between ZSCAN26 and H1-5, as well as between ZSCAN12 and GPX6, were conserved across cattle, humans, and pigs.
Development of a high-resolution typing method for nine MHC-linked OR genes in pigs
The high frequency of particular allele combinations at different loci may suggest the presence of LD or a functional association between them. To investigate the associations between MHC and linked ORs, we selected nine pig MHC-linked OR genes, including LOC100514111, LOC100516618, LOC100157348, LOC100515036, LOC100156552, LOC100522686, LOC100516811, OLF42-3, and OLF42-1, based on their cluster location, testicular expression level, and the presence of human orthologs (Supplementary Table 2). To develop a comprehensive typing method for these OR genes, we designed primers corresponding to the 500 bp upstream and downstream regions of each OR gene. Through iterative primer design and confirmation of amplification specificity, we identified a set of primers that enabled comprehensive amplification of the nine ORs (Table 1). We then developed sequencing primers for each amplicon, resulting in a sequence-based high-resolution typing method for ORs. For allelic determination, homozygous typing results were classified as novel alleles when observed for the first time. Heterozygous samples were separated by molecular cloning and sequenced to resolve the alleles. As a result, we successfully determined the allelic status of all typing results (n = 576) without any exceptions (Supplementary Tables 3 and 4). The sequence information for OR alleles was submitted to NCBI under accession numbers PP768337 to PP768442 (Supplementary Table 3). The number of alleles for ORs ranged from 5 to 22 (Supplementary Tables 3 and 5). Typing results for LOC100515036 revealed more than two alleles per individual across multiple samples, suggesting an association between gene duplications and copy number variations (CNVs).
Higher genetic diversity of SLA-linked ORs compared to SLA-unlinked ORs
Extreme genetic polymorphisms of the MHC can influence the diversity of linked genes. To assess the impact of LD with the MHC on the diversity of OR genes, we performed sequence-based typing of SLA-1, SLA-2, and SLA-3 in animals with existing OR typing data, using previously established methods [35,36,37]. High-resolution typing results were obtained for all three SLA class I genes, identifying 34, 29, and 17 alleles for SLA-1, SLA-2, and SLA-3, respectively. CNV was observed in SLA-1, consistent with previous findings [35, 38] (Supplementary Table 4). No novel alleles were detected for SLA-1 and SLA-2; however, a new non-functional allele was identified for SLA-3 (PQ037598), which contains a premature stop codon. Regarding genetic diversity, the observed heterozygosity (Ho) for SLA-1, SLA-2, and SLA-3 ranged from 0.646 to 0.813, with a mean of 0.750, while the expected heterozygosity (He) ranged from 0.942 to 0.961, with a mean of 0.954, consistent with previously reported diversity of SLA class I genes [35,36,37] (Supplementary Table 5).
When comparing the allelic diversity between nine SLA-linked ORs typed in this study and 20 SLA-unlinked ORs reported in a previous study, which examined mostly the same populations as in this study [39], the mean allele numbers for SLA-linked and SLA-unlinked ORs were 11.78 ± 5.38 and 6.65 ± 3.01, respectively (Fig. 2, Supplementary Table 5), indicating a significant difference between the two groups (P-value < 0.05). The mean Ho and He of SLA-linked OR genes were also significantly higher than those of SLA-unlinked ORs, with Ho = 0.68 ± 0.08 vs. 0.33 ± 0.18 (P-value < 0.01) and He = 0.81 ± 0.10 vs. 0.57 ± 0.24 (P-value < 0.01) (Fig. 2, Supplementary Table 5). Notably, three SLA-linked ORs—OLF42-1, OLF42-3, and LOC100515036—exhibited particularly high allele numbers. CNVs were observed in LOC100515036, suggesting gene duplication at this locus. However, the number of alleles was also high for the SLA-unlinked OR sOR6T3 (15 alleles), indicating that the specific functions of certain ORs may also contribute to genetic diversity, although we were unable to determine functional differences among these ORs.
Comparison of allele numbers and heterozygosity among SLA and OR genes. OR genes are classified into SLA-linked and SLA-unlinked based on genetic linkage to SLA genes, distal and proximal to SLA class I, and FPKM ≥ 0.1 and FPKM < 0.1 for testicular expression. Genetic diversity was compared for the number of alleles (A), observed heterozygosity (Ho; B), and expected heterozygosity (He; C) among different groups. The group names corresponding to each bar are identical through (A), (B), and (C) and are only shown on the x-axis of (C). The y-axis indicates the mean values of the group, and the standard deviations are indicated by vertical lines above the bar. Statistical significance using Student’s t-test is indicated with horizontal lines with * and ** denoting P-value < 0.05 and P-value < 0.01, respectively
Determination of 48 haplotypes for SLA-1, SLA-2, SLA-3, and nine SLA-linked ORs
We identified a combined total of 106 alleles for nine ORs and 80 alleles for three classical SLA class I genes by typing 48 individuals from six pig breeds (Supplementary Tables 3 and 4). The allele numbers for SLA-1 and LOC100515036, which exhibit CNVs, ranged from 0 (null) to 2 and 1 to 3 per haplotype, respectively, while the remaining genes showed one allele per haplotype, as expected. To determine haplotypes, preliminary haplotype phasing was performed without CNV-related genes to reduce the complexity of haplotype determination, using the expectation-maximization (EM) algorithm. Subsequently, final haplotypes were manually refined by incorporating the genotypes of the CNV-related genes, following the strategy described in Supplementary Fig. 1. As a result, 48 combined haplotypes were established for SLA-1, SLA-2, SLA-3, and nine SLA-linked ORs (Supplementary Table 6). When the haplotype sequences were translated into amino acid sequences, the number of haplotypes was reduced by only one, indicating a high rate of nonsynonymous substitutions in this region (Supplementary Table 6).
Identification of five putative breakpoint hotspots at locus junctions within the SLA-OR region
We identified candidate recombination breakpoint sites based on codon alignment analysis of 48 haplotype sequences within the SLA-OR region using RDP software (Fig. 3). Five sites emerged as prominent breakpoint hotspots with > 99% confidence, as determined by the observed haplotype patterns in the region (Fig. 3, Supplementary Fig. 2). The deduced breakpoint hotspots within the SLA and linked OR gene regions were primarily located at the locus boundaries between LOC100516618 and LOC100157348, LOC100156552 and LOC100522686, and OLF42-1 and SLA genes. Consistent with these findings, when pairwise LD among three SLA class I and nine linked OR genes was estimated by calculating normalized entropy differences between two loci (ε) using eLD software, LOC100157348, LOC100515036, LOC100156552, OLF42-3, and OLF42-1 showed higher LD with SLA-1, -2, and − 3 (mean ε ≥ 0.3) compared to other analyzed ORs (Fig. 4).
Prediction of recombination breakpoint hotspots for the SLA-OR haplotype-defined region. Recombination breakpoints were inferred from 48 pig SLA-OR haplotypes using haplotype codon alignment. Breakpoint positions (bp) were shown based on the sequence of a representative haplotype (h36) and indicated on the x-axis. The statistical significance of the breakpoints was tested by random permutations (n = 1000) in a 200 bp window. The P-values indicating statistical confidence in recombination breakpoint hotspots are shown on the y-axis. Breakpoint hotspots were defined as regions with P-value < 0.01. The names of OR genes including LOC100514111, LOC100516618, LOC100157348, LOC100515036, LOC100156552, LOC100522686, OLF42-3, and OLF42-1 are indicated on top together with SLA-1, SLA-2, and SLA-3. Only the last three digits of the gene symbols are shown for OR genes with the gene symbol starting with ‘LOC’. Arrows on tops indicate the chromosomal locations of the coding sequence of each gene constituting the OR-SLA haplotype in this study. Purple vertical lines below the gene names indicate variant positions in the haplotype sequence alignment
Pairwise LD plot for three SLA class I and nine OR genes. The pairwise LD levels among three SLA class I and nine OR genes were estimated according to normalized entropy differences between two loci (ε) using eLD software. The ε values for each gene pair are presented in a matrix format and shown at the intersection of the two genes. The cells corresponding to ε values are color-coded in blue and red, with color gradients from low to high LD, respectively, as described in the color key on the bottom left. Gene names and chromosomal locations are indicated on the top and the genes are arranged according to their relative positions on pig chromosome 7 (NC_010449). The names of OR genes starting with “LOC” shown on the gene map are indicated only with the last three digits of the gene symbols in the LD matrix
Similarity in the testicular expression of MHC-linked ORs between humans and pigs
To investigate the expression patterns of MHC-linked ORs, we analyzed the expression levels of nine SLA-linked ORs using publicly available transcriptome data from pig testes (from three 120-day-old pigs, GEO accession: GSE171756) and olfactory epithelium (OE) (from four newborn pigs, GEO accession: GSE197184) (Table 2, Supplementary Table 7). The expression levels of OLF42-1 and OLF42-3 in the proximal OR cluster of SLA class I and LOC100515036 and LOC100156552 in the distal OR cluster were higher (FPKM 0.102–0.742) than those of other ORs (FPKM 0–0.055; Table 2), indicating differential expression levels in the testis. However, this pattern differed from that of ORs expressed in OE, suggesting tissue-specific regulation of gene expression in these ORs (Table 2, Supplementary Table 7). Additionally, we analyzed the expression patterns of putative human OR orthologs of the nine pig ORs using publicly available transcriptome data from human tissues (GEO accession: GSE30611 for human testis and GSE80249 for human OE) and compared them with their expression in pigs. The expression of OR2H1 and OR2H2, human orthologs of pig OLF42-1 and OLF42-3, respectively, was also higher in the human testis (mean FPKM = 0.603) compared to other ORs (mean FPKM = 0.007), reflecting a similar expression pattern to that observed in pigs. Notably, OR2B8P, the putative human ortholog corresponding to pig LOC100515036 and LOC100156552 (likely duplicated in pigs), has been pseudogenized in humans. These results indicate that the testicular expression pattern of the analyzed functional HLA-linked ORs is similar between pigs and humans (Table 2, Supplementary Table 7, Supplementary Fig. 3).
Positive correlation between heterozygosity levels and testicular expression of SLA-linked ORs
From the expression data of nine SLA-linked pig OR genes in regions conserved between pigs and humans, we categorized the genes into two groups based on testicular expression levels exceeding 0.1 FPKM, using the round-off mean value (mean FPKM = 0.098) for expressed ORs in pig testes (Supplementary Table 7, Supplementary Fig. 3). LOC100515036, LOC100156552, OLF42-3, and OLF42-1 showed mean expression levels of ≥ 0.1 FPKM (n = 3), while LOC100514111, LOC100516618, LOC100157348, LOC100522686, and LOC100516811 had mean expression levels of < 0.1 FPKM (n = 3). When comparing these expression levels to the genetic diversity of each locus, Ho differed between the two groups, with mean Ho = 0.750 for FPKM ≥ 0.1 versus Ho = 0.617 for FPKM < 0.1 (P-value < 0.01; Fig. 2 and Supplementary Table 5). The mean Ho of ORs with FPKM ≥ 0.1 (Ho = 0.750) was similar to that of SLA class I genes (Ho = 0.750; P-value = 0.5) but differed from ORs with FPKM < 0.1 (Ho = 0.617; P-value < 0.05). Additionally, allele numbers were greater for ORs with FPKM ≥ 0.1 (mean n = 16.25) compared to those with FPKM < 0.1 (mean n = 8.20), suggesting a positive correlation between allele diversity and testicular expression levels for SLA-linked ORs.
Positive correlation between testicular expression and the number of codons under diversifying selection for SLA and linked ORs
We estimated selection pressure for 3,032 codons constituting nine OR and three SLA genes using a random-site model in PAML. We also calculated LD between all possible SNP pairs from 550 SNPs identified in 48 OR-SLA haplotype sequences (Supplementary Table 8). Our analysis revealed that 41 codons from nine ORs had dN/dS (ω) values > 1.00 according to both M2 and M8 models, indicating diversifying selection. Among these 41 diversifying codons, 25 were in significant LD (D’ > 0.5, LOD > 2) with diversifying codons of SLA genes (Table 3, Supplementary Table 9). Further analysis showed that 15 out of 16 (94%) functional OR genes with FPKM ≥ 0.1 (LOC100515036, LOC100156552, OLF42-3, OLF42-1) were in significant LD with SLA, compared to only 10 out of 25 (40%) for ORs with FPKM < 0.1 (LOC100514111, LOC100516618, LOC100157348, LOC100522686), except for LOC100522686 (89%). This finding suggests a correlation between OR expression in the testis and evolutionary selection between SLA and SLA-linked ORs.
Breed differences in SLA-OR haplotypes
We analyzed population differences in MHC-OR haplotype repertoires for SLA-1, -2, -3, and nine linked OR regions among six pig breeds: Berkshire (BER), Duroc (DUR), Landrace (LAN), Yorkshire (YOR), Korean native pigs (KNP), and SNU minipigs (SNU). Principal component analysis (PCA) of 48 haplotypes showed a high degree of genetic relatedness among European breeds (BER, DUR, LAN, and YOR), whereas KNP and SNU were more distantly related to the European pig cluster (Fig. 5). The mean FST value was highest for SNU (mean FST = 0.265), followed by KNP (mean FST = 0.222), and BER (mean FST = 0.212), with DUR (mean FST = 0.181), YOR (mean FST = 0.172), and LAN (mean FST = 0.156) showing lower values (Supplementary Fig. 4). This indicates that BER has relatively distinct MHC-OR haplotypes compared to other European breeds. In the minimum spanning network (MSN) plot, LAN occupied a central position with the largest number of haplotypes (n = 12) among the six breeds (Supplementary Fig. 5). LAN haplotypes h1, h15, and h28 were shared with YOR and DUR. BER was closely positioned to other European breeds in its haplotype composition but did not directly share haplotypes with them. KNP shared haplotype h7 with DUR. SNU pigs had the lowest number of haplotypes in the inbred population.
Genetic relationships of six pig breeds from principal component analysis using the typing results of three SLA class I and nine linked ORs. The results of principle component analysis using principal components (PC) 1 and 2 are shown. The relative eigenvalues of each component are indicated on the x- and y-axis, respectively. Each individual (n = 48) is represented by a dot, and different breeds are indicated by different colors, as shown on the upper left of the plot. The inset eigenvalue chart indicates the percentage of explained variance in the PCA plot, with PC1 and PC2 highlighted in black
Discussion
MHC molecules have evolved to be highly polymorphic and initiate immune responses against an almost unlimited number of foreign molecules [40,41,42,43]. Such evolutionary forces of MHC could affect the diversity of neighboring genes through mechanisms such as hitchhiking or functional association. Many OR genes, including 26 functional and 7 pseudogenes, are present in SLA-linked regions on pig chromosome 7 (Supplementary Table 7). Thus, we investigated the genetic and evolutionary influence of MHC on the diversity of neighboring OR genes by comparing genetic variations between SLA-linked and unlinked ORs.
The typing results of SLA-linked ORs in this study showed a significant increase in the genetic diversity of SLA-linked ORs compared to unlinked ORs (Fig. 2, Supplementary Table 5), suggesting MHC as a possible driving force for increasing the genetic diversity of neighboring ORs. The influence of MHC on the genetic diversity of linked ORs was also evidenced by the increased allelic diversity and observed heterozygosity of ORs with strong LD to SLA compared to those with weak or no LD to SLA (Fig. 2, Supplementary Table 5).
In this study, we conducted a comparative analysis of MHC-linked ORs using current genome assemblies and annotations. Thus, we provided an updated and more detailed picture of the comparative OR-MHC architecture among cattle, humans, mice, and pigs. For example, several OR genes within the extended MHC (xMHC) region were mapped differently in the previous genome assembly (Sscrofa10.2) than in the current genome assembly, Sscrofa11.1 (Supplementary Table 10). In addition, the number of GABBR1-linked pig ORs mapped to this region appeared to be smaller than those of other species (Fig. 1). However, we confirmed the structural conservation of the major OR cluster linked to MHC among the four species, despite the presence of species specificity.
Pig ORs, OLF42-3 and OLF42-1, which showed higher genetic diversity and testicular expression than the other ORs, were also present in the species-conserved region. A previous study suggested that the strong genetic linkage between GABBR1-linked ORs and HLA could be related to the functional interactions of these molecules [25, 44]. The orthologs of pig OLF42-1 and OLF42-3 which are highly expressed in pig testes, are also expressed in human sperm [22]. HLA class I expression is observed in granulosa cells surrounding oocytes in humans [26, 45]. These results suggested that MHC-linked ORs play a role in the animal reproduction system.
In studies on human testicular expression of ORs, the majority of OR genes expressed in the testis were genetically linked to HLA [46,47,48]. These testicular HLA-linked OR genes show significant variations in expression levels and exhibit a non-canonical expression pattern through long-distance splicing and exon sharing among OR genes [24, 49]. Interestingly, this promiscuous expression plays a critical role in the establishment of self-tolerance in T cells [50]. Therefore, MHC-linked ORs that are highly expressed in the testes or germ cells may play a role in mediating MHC genotypes through strong LD between MHC and OR genes. This is consistent with our results in which the frequency of diversifying codons in significant LD with SLA was much higher for functional OR genes of FPKM ≥ 0.1 than those of FPKM < 0.1 (Table 3). Furthermore, several testicular MHC-linked ORs are expressed in sperm [22, 48], and GABBR1, which is involved in OR signaling, is also expressed in the testis [27, 51]. These findings suggest that MHC diversity plays a role in OR function in testes and sperm.
A study of the human OR2E1P pseudogene and OR2H1, the orthologs of pig LOC100522686 and OLF42-1, respectively, revealed that OR2E1P can form spliced variant transcripts with OR2H1 [49]. Interestingly, the alternative splice sites involved in the exon splicing of OR2H1 were also conserved in porcine OLF42-1 (Supplementary Fig. 6), demonstrating the interspecies conservation of evolutionary and functional characteristics among MHC-linked ORs.
Vomeronasal receptors (VRs), a class of ORs that function as pheromone receptors, recognize the diversity of MHC peptide ligands and trigger responses from vomeronasal sensory neurons expressing these receptors [52, 53]. It has been suggested that MHC peptide ligands can serve as olfactory cues [54]. However, it is unclear whether MHC-linked OR genes expressed in the testes and sperm recognize MHC particles and elicit responses.
The genetic linkage between MHC and MHC-linked OR genes, along with their distinct expression patterns, suggests the presence of an underlying biological mechanism connecting them. A compelling hypothesis is that MHC-linked ORs contribute to mate selection by recognizing potential partners with different MHC genotypes [25, 26]. However, there is no experimental evidence supporting the idea that MHC-linked ORs act as receptors for MHC particles as ligands. Differences in the sensitivity of MHC-linked ORs to specific odorants may also contribute to mate selection due to the linkage effect between MHC and linked ORs. Therefore, elucidating the ligand molecules for MHC-linked ORs is critical to understanding the mechanism associated with MHC-based mate selection.
Higher genetic diversity of MHC-linked ORs compared to unlinked ORs could contribute to more efficient differentiation of specific odorants. Mate selection may have resulted from an evolutionary outcome that enhances immune surveillance through the heterozygote advantage of MHC molecules. Increased genetic diversity of receptors for either foreign antigens or odorant molecules should be beneficial for animal survival. Therefore, we believe that the phenomenon of increased genetic diversity of MHC-linked ORs is likely to be observed in other animals although direct interactions between MHC and OR genes have not been reported in other species.
We identified 48 haplotypes for the nine SLA-linked ORs and three major SLA class I genes using sequence-based genotyping of 48 animals from six pig breeds (Supplementary Tables 3 and 4). Sampling bias can be a major cause of error in haplotype estimation using the EM algorithm [55]. To overcome the concerns of a small sample size and large number of loci for haplotype estimation in our dataset, we conducted indirect validation of our haplotype analysis results by comparing the population distribution of deduced haplotypes in this study with the outcomes of previous studies independently conducted for 20 MHC-unlinked ORs and the same set of SLA class I genes from populations of mostly overlapping breeds of pigs [39, 56, 57]. The allelic and haplotype distributions of SLA-linked ORs for each breed in this study were highly consistent with those of SLA-unlinked ORs reported in a previous study, supporting the reliability of our OR-SLA haplotype information.
Conclusions
Genetic analysis of ORs in the vertebrate genome is challenging because of the presence of high sequence homology, CNV, and pseudogenes, as well as the large number of loci in the gene family. Genetic and evolutionary characteristics in relation to MHC class I genes in pigs indicated a strong influence of SLA diversity on that of the linked OR genes. Our results show the conservation of testicular expression patterns of MHC-linked orthologous ORs between humans and pigs, suggesting possible interactions between MHC molecules and their linked ORs in relation to animal reproduction. Outcomes of future studies on the identification of ligands for MHC-linked ORs and the OR-MHC haplotypes defined in this study contribute to elucidating the mechanism underlying MHC haplotype-based mate selection. Identifying differences in the binding affinity of ligands to MHC-linked ORs among haplotypes and applying these genetic differences to animals could enhance the efficiency of animal breeding and reproduction.
Materials and methods
Animals and tissues
The informed consent for animal use was obtained from a local farm and the university facility. Tissue samples of six pig breeds, BER, DUR, LAN, YOR, KNP, and SNU were collected via ear notching and stored at -80 °C in previous studies [35, 58]. Eight individuals from each breed were randomly selected. animal facility. All experimental procedures were approved (KU20229) and performed accordance with the guidelines and regulations set by the Institute of Animal Care and Use Committee (IACUC) and the Center for Research Ethincs of Konkuk University.
Preparation of genomic DNA
Genomic DNA was extracted from tissues (~ 0.5 g) using a standard protocol. Briefly, tissues were incubated with 700 µL lysis buffer (50 mM Tris pH 8.0, 0.1 M EDTA pH 8.0, 0.5% (w/v) sodium dodecyl sulfate, 20 µg/mL DNase-free pancreatic RNase) and 20 µL of proteinase K (20 mg/mL) at 50 °C overnight. Subsequently, gDNA was purified by phenol: chloroform: isoamyl alcohol (pH 8.0) extraction, precipitated using ethanol, and dissolved in 50 mM Tris-EDTA buffer (pH 8.0).
Determination of syntenic relationship
The genome assemblies and annotations used in this study were GRCh38.p14, Sscrofa11.1, ARS-UCD2.0, and GRCm39 for humans, pigs, cattle, and mice, respectively. The construction of a synteny plot was carried out using the Python package “pyGenomeViz” (github.com/moshi4/pyGenomeViz) on Python version 3.8.15 (www.python.org/).
Primer design
Primers that specifically amplify the entire coding sequence (CDS) of each of the nine MHC-linked OR genes were designed using NCBI Primer-BLAST (www.ncbi.nlm.nih.gov/tools/primer-blast/). The 500-bp upstream and downstream regions of the selected OR genes, based on the pig genome assembly (NC_010449.5, Sscrofa11.1), were used as queries. The maximum product size was set to 2500 bp. Primer sites containing nucleotide variations reported for dbSNPs (www.ncbi.nlm.nih.gov/snp/) were excluded. Off-target or multiple-target amplification of the primers was performed using BLAST analysis. Primer design was repeated until the optimal primers for the specific amplification of the target ORs were obtained. Primer information for the analysis of SLA-1, -2, and − 3 was based on previous studies [35,36,37] and is described in Table 1.
Polymerase chain reaction and sequencing
For a 10 µL reaction for nine OR genes including LOC100514111, LOC100516618, LOC100157348, LOC100515036, LOC100156552, LOC100522686, LOC100516811, OLF42-3, OLF42-1, the reaction mixture consisted of 1 µM of locus-specific primers (Table 1), 250 µM dNTPs, 1 unit of Supertherm™ Taq DNA polymerase (JMR Holdings, Kent, UK), 10× reaction buffer containing 15 mM MgCl2, and 25 ng of DNA. The amplification reaction was conducted on an ABI9700 thermocycler (Applied Biosystems, Foster City, CA) with an initial pre-denaturation step at 94 °C for 3 min, followed by 30 cycles of denaturation at 94 °C for 30 s, annealing at the specific primer annealing temperature (Table 1) for 45 s, and elongation at 72 °C for 90 s. Subsequently, the final elongation was conducted at 72 °C for 10 min. The amplicons were electrophoresed on 1% agarose gel to confirm target amplification. To sequence OR amplicons, sequencing primers with lengths of 15–20 bp were designed at conserved regions in both the forward and reverse directions with an overlap to cover the entire CDS of each OR gene using the Primer Designer of CLC Main Workbench version 7.8.1 (CLC bio, Aarhus, Denmark) (Table 1). Before sequencing, 5 µL of PCR products were mixed with 0.25 U of shrimp alkaline phosphatase (USB Corporation, Cleveland, OH), 15 U of exonuclease I (Fermentas, Massachusetts, USA), and incubated at 37 ℃ for 30 min. The sequencing reaction was performed using the Applied Biosystems BigDye® Terminator v3.1 Cycle Sequencing Kit (Applied Biosystems, Massachusetts, USA) with 2 pmol of a sequencing primer under the conditions of pre-denaturation at 96 °C for 1 min, 25 cycles of 96 °C for 10 s, 50 °C for 5 s, and 60 °C for 4 min. The reaction products were purified using ethanol precipitation, resuspended in 10 µL of Hi-Di™ Formamide (Applied Biosystems, Massachusetts, USA), and analyzed on an ABI3730 DNA Analyzer (Applied Biosystems). Specific amplification and sequencing of SLA-1, -2, and − 3 were conducted as previously described (Table 1) [35,36,37].
Allelic differentiation and haplotype determination
In principle, alleles that exist as homozygotes in sequence-based typing are considered novel alleles. Alleles that existed as heterozygotes in the typing results were separated into individual alleles through TA cloning using pGEM®-T Easy Vector Systems (Promega, Wisconsin, USA). For cloning, the ligation products were transformed into DH10B competent cells (Thermo Fisher Scientific, Waltham, MA, USA) using electrotransformation. Target inserts were amplified in 10 µL of colony PCR mixture containing a piece of a single bacterial colony as a DNA template, 1 µM of T7 and SP6 universal primers (Table 1), 250 µM dNTP, 1U of Supertherm™ Taq DNA polymerase (JMR Holdings, Kent, UK), and 10× reaction buffer with 15 mM MgCl2 (JMR Holdings, Kent, UK). The thermal profile for the colony PCR consisted of 5 min of pre-denaturation at 94 °C, 30 cycles of denaturation at 94 °C for 30 s, primer annealing at 50 °C for 30 s, and elongation at 72 °C for 60 s, followed by a final elongation at 72 °C for 7 min. The PCR amplicons were sequenced using primers specific to each OR gene (Table 1). At least eight independent clones were sequenced for allelic determination using the T7 and SP6 universal primers. Allele differentiation for SLA-1, -2, and − 3 was conducted as described in previous studies [35,36,37]. Haplotype determination of SLA- and SLA-linked OR genes was conducted using the haplotype inference function, employing the EM algorithm provided in Arlequin version 3.5.2.2 [59]. The detailed procedure is shown in Supplementary Fig. 1.
Calculation of genetic diversity
The observed (Ho) and expected heterozygosities (He) of the loci were calculated using Arlequin version 3.5.2.2 [59]. Data normality and equality in variance distribution for the observed indices were tested using the Shapiro-Wilk normality test and an equal-variance test [60, 61]. Differences in the average values of the observed indices were tested using the Student’s t-test [62].
Population genetic analyses
LD index (ε) between loci was calculated after allele differentiation of genotyping results using eLD R script [63] on R version 4.1.2 (www.r-project.org). To generate an MSN plot, haplotype data was prepared for R library “adegenet” version 2.1.6 [64], and generation of a distance matrix and visualization was conducted using R library “poppr” version 2.9.4 [65]. A dataset for PCA was prepared using the genotype data of nine OR and three SLA class I genes from 48 individuals of six pig breeds, using the adegenet R package. PCA was conducted using the dudi.pca function of the R library “ade4” version 1.7.19 [66].
Estimation of haplotype breakpoints
To generate input nucleotide sequences for haplotype codon alignment, stop codons at the end of the full-length coding sequences (CDS) for 11 genes were removed, and the sequences were sequentially connected to produce a single sequence contig relative to the chromosomal order, except LOC100516811, a pseudogene. The input sequences were aligned using the codon alignment function of MUSCLE in MEGA version 11.0.10 [67]. The generated codon alignment was used as an input file for RDP v4.101 [68]. Haplotype breakpoints were deduced using RDP, GENECONV, MaxChi, BootScan, and SiScan in RDP software. Statistical significance of the breakpoints was tested using a breakpoint P distribution plot (1000 permutations and window size = 200 bp). Breakpoints with an estimated P-value < 0.01 were suggested as breakpoint hotspots.
Codon selection test and LD calculation
The haplotype phylogenetic tree was constructed from the haplotype codon alignment using the “Create Tree” feature of CLC Main Workbench version 7.8.1 (CLC bio, Aarhus, Denmark). The neighbor-joining method and Kimura 80 nucleotide substitution model were applied with 5000 bootstrap replications. The created haplotype phylogenetic tree was unrooted using the R library “ape” version 5.6.2 [69]. and was used as the input file for CodeML of pamlX version 1.3.1 [70]. All analysis options were set to default except for codon frequency, which was configured using the 2:F3x4 model. To test selection signatures of codons under Random-sites models, we used site models to allow the dN/dS (ω) ratio to vary among sites using Models 0, 1, 2, 7, and 8 under different assumptions on the distribution of ω ratio [70]. Empirical Bayes analysis using Models 2 and 8 was used to deduce the positively selected codons [71]. Codons with a confidence level > 95% were identified as positively selected codons. LD values between haplotype SNPs were estimated using Haploview version 4.2 from the haplotype codon alignment [72].
Comparison of OR expression
Expression data for SLA-linked OR genes were obtained from the NCBI Gene Expression Omnibus database (www.ncbi.nlm.nih.gov/geo/; accession numbers GSE171756 (pig testis) and GSE197184 (pig olfactory epithelium (OE)). Expression data for human ORs were retrieved from the same database with accession numbers GSE30611 (human testis) and GSE80249 (human OE). Because the OR expression data for porcine OE were available only as read counts, the values were converted to relative FPKM values for expression comparison using the following formula:
\(\eqalign{{\rm{OR}}\,{\rm{ gene}}\,{\rm{ FPKM}}\,{\rm{ in}}\,{\rm{ OE}} & = {\rm{ OR}}\,{\rm{ gene}}\,{\rm{ read }}\,{\rm{count }}\,{\rm{in}}\,{\rm{ OE }} \cr & \times \left( {{{{\rm{OR}}\,{\rm{ gene}}\,{\rm{ FPKM}}\,{\rm{ in}}\,{\rm{ testis}}} \over {{\rm{OR}}\,{\rm{ gene }}\,{\rm{read}}\,{\rm{ count }}\,{\rm{in}}\,{\rm{ testis}}}}} \right) \cr} \)
Data availability
A total of 106 SLA-linked OR allele sequences analyzed in this study were submitted to the NCBI GenBank (www.ncbi.nlm.nih.gov/genbank/) under the accession numbers listed in Supplementary Table 3 (PP768351-PP768442). All other information can be found in the text and supporting information.
References
Buck L, Axel R. A novel multigene family may encode odorant receptors: a molecular basis for odor recognition. Cell. 1991;65(1):175–87.
Glusman G, Bahar A, Sharon D, Pilpel Y, White J, Lancet D. The olfactory receptor gene superfamily: data mining, classification, and nomenclature. Mamm Genome. 2000;11(11):1016–23.
Glusman G, Yanai I, Rubin I, Lancet D. The complete human olfactory subgenome. Genome Res. 2001;11(5):685–702.
Zhang X, Firestein S. The olfactory receptor gene superfamily of the mouse. Nat Neurosci. 2002;5(2):124–33.
Niimura Y. Olfactory receptor multigene family in vertebrates: from the viewpoint of evolutionary genomics. Curr Genomics. 2012;13(2):103–14.
Nguyen DT, Lee K, Choi H, Choi MK, Le MT, Song N, Kim JH, Seo HG, Oh JW, Lee K, et al. The complete swine olfactory subgenome: expansion of the olfactory gene repertoire in the pig genome. BMC Genomics. 2012;13(1):584.
Lee K, Nguyen DT, Choi M, Cha SY, Kim JH, Dadi H, Seo HG, Seo K, Chun T, Park C. Analysis of cattle olfactory subgenome: the first detail study on the characteristics of the complete olfactory receptor repertoire of a ruminant. BMC Genomics. 2013;14(1):596.
Barnes IHA, Ibarra-Soria X, Fitzgerald S, Gonzalez JM, Davidson C, Hardy MP, Manthravadi D, Van Gerven L, Jorissen M, Zeng Z, et al. Expert curation of the human and mouse olfactory receptor gene repertoires identifies conserved coding regions split across two exons. BMC Genomics. 2020;21(1):196.
Hayden S, Bekaert M, Crider TA, Mariani S, Murphy WJ, Teeling EC. Ecological adaptation determines functional mammalian olfactory subgenomes. Genome Res. 2010;20(1):1–9.
McGowen MR, Clark C, Gatesy J. The vestigial olfactory receptor subgenome of odontocete whales: phylogenetic congruence between gene-tree reconciliation and supermatrix methods. Syst Biol. 2008;57(4):574–90.
Kishida T, Kubota S, Shirayama Y, Fukami H. The olfactory receptor gene repertoires in secondary-adapted marine vertebrates: evidence for reduction of the functional proportions in cetaceans. Biol Lett. 2007;3(4):428–30.
Liu A, He F, Shen L, Liu R, Wang Z, Zhou J. Convergent degeneration of olfactory receptor gene repertoires in marine mammals. BMC Genomics. 2019;20(1):977.
Kang N, Koo J. Olfactory receptors in non-chemosensory tissues. BMB Rep. 2012;45(11):612–22.
Feldmesser E, Olender T, Khen M, Yanai I, Ophir R, Lancet D. Widespread ectopic expression of olfactory receptor genes. BMC Genomics. 2006;7(1):121.
Flegel C, Manteniotis S, Osthold S, Hatt H, Gisselmann G. Expression profile of ectopic olfactory receptors determined by deep sequencing. PLoS ONE. 2013;8(2):e55368.
Aisenberg WH, Huang J, Zhu W, Rajkumar P, Cruz R, Santhanam L, Natarajan N, Yong HM, De Santiago B, Oh JJ, et al. Defining an olfactory receptor function in airway smooth muscle cells. Sci Rep. 2016;6(1):38231.
Braun T, Voland P, Kunz L, Prinz C, Gratzl M. Enterochromaffin cells of the human gut: sensors for spices and odorants. Gastroenterology. 2007;132(5):1890–901.
Manteniotis S, Wojcik S, Brauhoff P, Mollmann M, Petersen L, Gothert JR, Schmiegel W, Duhrsen U, Gisselmann G, Hatt H. Functional characterization of the ectopically expressed olfactory receptor 2AT4 in human myelogenous leukemia. Cell Death Discov. 2016;2(1):15070.
Maßberg D, Hatt H. Human olfactory receptors: Novel Cellular functions outside of the nose. Physiol Rev. 2018;98(3):1739–63.
Sanz G, Leray I, Dewaele A, Sobilo J, Lerondel S, Bouet S, Grebert D, Monnerie R, Pajot-Augy E, Mir LM. Promotion of cancer cell invasiveness and metastasis emergence caused by olfactory receptor stimulation. PLoS ONE. 2014;9(1):e85110.
Eisenbach M, Giojalas LC. Sperm guidance in mammals - an unpaved road to the egg. Nat Rev Mol Cell Biol. 2006;7(4):276–85.
Flegel C, Vogel F, Hofreuter A, Schreiner BS, Osthold S, Veitinger S, Becker C, Brockmeyer NH, Muschol M, Wennemuth G, et al. Characterization of the olfactory receptors expressed in human spermatozoa. Front Mol Biosci. 2015;2:73.
Spehr M, Gisselmann G, Poplawski A, Riffell JA, Wetzel CH, Zimmer RK, Hatt H. Identification of a testicular odorant receptor mediating human sperm chemotaxis. Science. 2003;299(5615):2054–8.
Volz A, Ehlers A, Younger R, Forbes S, Trowsdale J, Schnorr D, Beck S, Ziegler A. Complex transcription and splicing of odorant receptor genes. J Biol Chem. 2003;278(22):19691–701.
Ziegler A, Santos PS, Kellermann T, Uchanska-Ziegler B. Self/nonself perception, reproduction and the extended MHC. Self Nonself. 2010;1(3):176–91.
Ziegler A, Dohr G, Uchanska-Ziegler B. Possible roles for products of polymorphic MHC and linked olfactory receptor genes during selection processes in reproduction. Am J Reprod Immunol. 2002;48(1):34–42.
Burrello N, Vicari E, D’Amico L, Satta A, D’Agata R, Calogero AE. Human follicular fluid stimulates the sperm acrosome reaction by interacting with the gamma-aminobutyric acid receptors. Fertil Steril. 2004;82(Suppl 3):1086–90.
Andrews PW, Boyse EA. Mapping of Anh-2-linked gene that influences mating preference in mice. Immunogenetics. 1978;6(1):265–8.
Yamazaki K, Boyse EA, Mike V, Thaler HT, Mathieson BJ, Abbott J, Boyse J, Zayas ZA, Thomas L. Control of mating preferences in mice by genes in the major histocompatibility complex. J Exp Med. 1976;144(5):1324–35.
Roberts SC, Gosling LM. Genetic similarity and quality interact in mate choice decisions by female mice. Nat Genet. 2003;35(1):103–6.
Bruce HM, A BLOCK TO PREGNANCY IN THE MOUSE CAUSED BY PROXIMITY OF STRANGE MALES. Reproduction. 1960;1(1):96–103.
Lovlie H, Gillingham MA, Worley K, Pizzari T, Richardson DS. Cryptic female choice favours sperm from major histocompatibility complex-dissimilar males. Proc Biol Sci. 2013;280(1769):20131296.
Santos PS, Seki Uehara CJ, Ziegler A, Uchanska-Ziegler B, Bicalho Mda G. Variation and linkage disequilibrium within odorant receptor gene clusters linked to the human major histocompatibility complex. Hum Immunol. 2010;71(9):843–50.
da Silva JS, Wowk PF, Poerner F, Santos PS, Bicalho Mda G. Absence of strong linkage disequilibrium between odorant receptor alleles and the major histocompatibility complex. Hum Immunol. 2013;74(12):1619–23.
Le MT, Choi H, Lee H, Le VCQ, Ahn B, Ho C-S, Hong K, Song H, Kim J-H, Park C. SLA-1 genetic diversity in pigs: extensive analysis of Copy Number Variation, Heterozygosity, expression, and Breed specificity. Sci Rep. 2020;10(1):1–12.
Choi H, Le MT, Lee H, Choi MK, Cho HS, Nagasundarapandian S, Kwon OJ, Kim JH, Seo K, Park JK, et al. Sequence variations of the locus-specific 5’ untranslated regions of SLA class I genes and the development of a comprehensive genomic DNA-based high-resolution typing method for SLA-2. Tissue Antigens. 2015;86(4):255–66.
Youk S, Le MT, Kang M, Ahn B, Choi M, Kim K, Kim TH, Kim JH, Ho CS, Park C. Development of a high-resolution typing method for SLA-3, swine MHC class I antigen 3. Anim Genet. 2022;53(1):166–70.
Tanaka-Matsuda M, Ando A, Rogel-Gaillard C, Chardon P, Uenishi H. Difference in number of loci of swine leukocyte antigen classical class I genes among haplotypes. Genomics. 2009;93(3):261–73.
Kang M, Ahn B, Youk S, Jeon H, Soundarajan N, Cho ES, Park W, Park C. Individual and population diversity of 20 representative olfactory receptor genes in pigs. Sci Rep. 2023;13(1):18668.
Erickson RP. Natural history of the major histocompatibility complex. Am J Hum Genet. 1987;40(5):468–9.
Robinson J, Malik A, Parham P, Bodmer JG, Marsh SG. IMGT/HLA database–a sequence database for the human major histocompatibility complex. Tissue Antigens. 2000;55(3):280–7.
Maccari G, Robinson J, Ballingall K, Guethlein LA, Grimholt U, Kaufman J, Ho CS, de Groot NG, Flicek P, Bontrop RE, et al. IPD-MHC 2.0: an improved inter-species database for the study of the major histocompatibility complex. Nucleic Acids Res. 2017;45(D1):D860–4.
Reche PA, Reinherz EL. Sequence variability analysis of human class I and class II MHC molecules: functional and structural correlates of amino acid polymorphisms. J Mol Biol. 2003;331(3):623–41.
Santos PS, Kellermann T, Uchanska-Ziegler B, Ziegler A. Genomic architecture of MHC-linked odorant receptor gene repertoires among 16 vertebrate species. Immunogenetics. 2010;62(9):569–84.
Dohr GA, Motter W, Leitinger S, Desoye G, Urdl W, Winter R, Wilders-Truschnig MM, Uchanska-Ziegler B, Ziegler A. Lack of expression of HLA [corrected] class I and class II molecules on the human oocyte. J Immunol. 1987;138(11):3766–70.
Branscomb A, Seger J, White RL. Evolution of odorant receptors expressed in mammalian testes. Genetics. 2000;156(2):785–97.
Parmentier M, Libert F, Schurmans S, Schiffmann S, Lefort A, Eggerickx D, Ledent C, Mollereau C, Gerard C, Perret J, et al. Expression of members of the putative olfactory receptor gene family in mammalian germ cells. Nature. 1992;355(6359):453–5.
Vanderhaeghen P, Schurmans S, Vassart G, Parmentier M. Specific repertoire of olfactory receptor genes in the male germ cells of several mammalian species. Genomics. 1997;39(3):239–46.
Younger RM, Amadou C, Bethel G, Ehlers A, Lindahl KF, Forbes S, Horton R, Milne S, Mungall AJ, Trowsdale J, et al. Characterization of clustered MHC-linked olfactory receptor genes in human and mouse. Genome Res. 2001;11(4):519–30.
Kyewski B, Derbinski J, Gotter J, Klein L. Promiscuous gene expression and central T-cell tolerance: more than meets the eye. Trends Immunol. 2002;23(7):364–71.
Vidal RL, Ramirez A, Castro M, Concha II, Couve A. Marlin-1 is expressed in testis and associates to the cytoskeleton and GABAB receptors. J Cell Biochem. 2008;103(3):886–95.
Leinders-Zufall T, Brennan P, Widmayer P, Maul-Pavicic SPC, Jager A, Li M, Breer XH, Zufall H, Boehm F. MHC class I peptides as chemosensory signals in the vomeronasal organ. Science. 2004;306(5698):1033–7.
Leinders-Zufall T, Ishii T, Mombaerts P, Zufall F, Boehm T. Structural requirements for the activation of vomeronasal sensory neurons by MHC peptides. Nat Neurosci. 2009;12(12):1551–8.
Milinski M. A review of suggested mechanisms of MHC odor signaling. Biology (Basel). 2022;11(8):1187.
Fallin D, Schork NJ. Accuracy of haplotype frequency estimation for biallelic loci, via the expectation-maximization algorithm for unphased diploid genotype data. Am J Hum Genet. 2000;67(4):947–59.
Choi JW, Chung WH, Lee KT, Cho ES, Lee SW, Choi BH, Lee SH, Lim W, Lim D, Lee YG, et al. Whole-genome resequencing analyses of five pig breeds, including Korean wild and native, and three European origin breeds. DNA Res. 2015;22(4):259–67.
Lee YS, Son S, Lee HK, Lee RH, Shin D. Elucidating breed-specific variants of native pigs in Korea: insights into pig breeds’ genomic characteristics. Anim Cells Syst (Seoul). 2022;26(6):338–47.
Jeon H, Le MT, Ahn B, Cho HS, Le VCQ, Yum J, Hong K, Kim JH, Song H, Park C. Copy number variation of PR-39 cathelicidin, and identification of PR-35, a natural variant of PR-39 with reduced mammalian cytotoxicity. Gene. 2019;692:88–93.
Excoffier L, Lischer HE. Arlequin suite ver 3.5: a new series of programs to perform population genetics analyses under Linux and Windows. Mol Ecol Resour. 2010;10(3):564–7.
Shapiro SS, Wilk MB. An analysis of variance test for normality (complete samples). Biometrika. 1965;52(3–4):591–611.
Brown MB, Forsythe AB. Robust tests for the Equality of variances. J Am Stat Assoc. 1974;69(346):364–7.
Student. The probable error of a Mean. Biometrika. 1908;6(1):1–25.
Okada Y. eLD: entropy-based linkage disequilibrium index between multiallelic sites. Hum Genome Var. 2018;5:29.
Jombart T. Adegenet: a R package for the multivariate analysis of genetic markers. Bioinformatics. 2008;24(11):1403–5.
Kamvar ZN, Tabima JF, Grunwald NJ. Poppr: an R package for genetic analysis of populations with clonal, partially clonal, and/or sexual reproduction. PeerJ. 2014;2:e281.
Dray S, Dufour A-B. Theade4Package: implementing the duality Diagram for ecologists. J Stat Softw. 2007;22(4):1–20.
Tamura K, Stecher G, Kumar S. MEGA11: Molecular Evolutionary Genetics Analysis Version 11. Mol Biol Evol. 2021;38(7):3022–7.
Martin DP, Murrell B, Golden M, Khoosal A, Muhire B. RDP4: detection and analysis of recombination patterns in virus genomes. Virus Evol. 2015;1(1):vev003.
Paradis E, Schliep K. Ape 5.0: an environment for modern phylogenetics and evolutionary analyses in R. Bioinformatics. 2019;35(3):526–8.
Yang Z. PAML 4: phylogenetic analysis by maximum likelihood. Mol Biol Evol. 2007;24(8):1586–91.
Yang Z, Wong WS, Nielsen R. Bayes empirical bayes inference of amino acid sites under positive selection. Mol Biol Evol. 2005;22(4):1107–18.
Barrett JC, Fry B, Maller J, Daly MJ. Haploview: analysis and visualization of LD and haplotype maps. Bioinformatics. 2005;21(2):263–5.
Acknowledgements
The authors have no specific acknowledgements to declare.
Funding
This study was supported by the Cooperative Research Program for Agriculture, Science, and Technology Development (Project No. PJ016221), Rural Development Administration, Republic of Korea.
Author information
Authors and Affiliations
Contributions
Conceptualization and sample collection: M.K., B.A., and J.L. Data curation: M.K., H.C., and C.P. Bioinformatic analysis: M.K., B.A., and J.S. Methodology: M.K., and C.P. Manuscript writing: M.K. and C.P. Comments and discussion: C.P.
Corresponding author
Ethics declarations
Ethical approval
All experiments were approved and performed in accordance with the guidelines and regulations of the Institute of Animal Care and Use Committee and the Center for Research Ethics of Konkuk University.
Consent for publication
Not applicable.
Competing interests
The authors declare no competing interests.
Additional information
Publisher’s note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Electronic supplementary material
Below is the link to the electronic supplementary material.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Kang, M., Ahn, B., Shin, J.Y. et al. Influence of MHC on genetic diversity and testicular expression of linked olfactory receptor genes. BMC Genomics 26, 115 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12864-025-11281-x
Received:
Accepted:
Published:
DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12864-025-11281-x