Next Article in Journal
Phenotypic and Genetic Characterization of Klebsiella pneumoniae Isolates from Wild Animals in Central Italy
Previous Article in Journal
Exploring the Evacuation of Dairy Cattle at Night in Collaboration with the Fire Brigade: How to Prepare Openings for Swift Rescue in Case of Barn Fire
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Within- and between-Breed Selection Signatures in the Original and Improved Valachian Sheep

1
Faculty of Agrobiology and Food Resources, Institute of Nutrition and Genomics, Slovak University of Agriculture in Nitra, Tr. A. Hlinku 2, 94976 Nitra, Slovakia
2
Department of Sustainable Agricultural Systems, Division of Livestock Sciences, University of Natural Resources and Life Sciences, Vienna, Gregor-Mendel-Straße 33, 1180 Vienna, Austria
3
Research Institute of Animal Production—NPPC Slovakia, Hlohovecká 2, 95141 Nitra—Lužianky, Slovakia
4
Faculty of Agrobiology and Food Resources, Institute of Animal Science, Slovak University of Agriculture in Nitra, Tr. A. Hlinku 2, 94976 Nitra, Slovakia
*
Author to whom correspondence should be addressed.
Animals 2022, 12(11), 1346; https://doi.org/10.3390/ani12111346
Submission received: 28 March 2022 / Revised: 13 May 2022 / Accepted: 23 May 2022 / Published: 25 May 2022
(This article belongs to the Section Animal Genetics and Genomics)

Abstract

:

Simple Summary

Each country has their own national breeds that are well adapted to the local climate and other conditions. Our study examines two national sheep breeds from Slovakia, the Original and the Improved Valachian sheep. The main goal of the study was to identify parts of their genome which contain genes enabling such adaptations. These are the so called selection signatures. Within the breed we look for genomic regions that were remarkably similar among a large proportion of genotyped individuals. The results point towards adaptation to high altitude pastures and resistance to parasites. The signatures of selection between the Original and Improved Valachian sheep focus on regions with striking difference in the same region of the genome between the two breeds. These results highlighted genes that improve the milk, meat and wool production, and their quality characteristics.

Abstract

This study explored the genomic diversity and selection signatures in two Slovakian national breeds, the Original Valachian and the Improved Valachian sheep. As they are an important animal genetic resource within the country, but with decreasing population size, our aim is to identify potentially valuable genomic regions. A total of 97 sheep (18 male and 79 female) from the Original Valachian, and 69 sheep (25 male and 44 female) from the Improved Valachian populations were genotyped using the GeneSeek GGP Ovine 50 K chip. The inbreeding levels were assessed with runs of homozygosity (ROH). The selection signatures within breeds were identified based on the top 1% of most homozygous regions within the breed, the so-called ROH islands. The selection signatures between breeds were assessed based on variance in linkage disequilibrium. Overall, we have identified selection signatures with quantitative trait loci (QTL) and genes pointing towards all three production purposes of the Valachian sheep, milk, meat, and wool, including their quality characteristics. Another group with apparent large importance was the various traits related to health and resistance to parasites, which is well in line with the sturdy nature of this breed.

1. Introduction

The domestication of various livestock species is one of the greatest achievements of humankind. This process inherently resulted in permanent changes in the domesticated species, related to their morphological characteristics, behavior, and production levels. Parallel to the changes at the phenotype level, there were changes also occurring on the genomes. These detectable changes are called selection signatures, and they are of great interest to evolutionary biologists and livestock researchers alike.
Sheep was one of the prime targets of domestication efforts, due to the wide variety of useful products, ability to adapt to different climatic conditions, and manageable size [1]. Subsequently, a large variety of breeds were created during the process of artificial selection, with specialized (or sometimes combined) production of wool, meat, and milk. With the widespread availability of the single nucleotide polymorphism (SNP) genotype data, it became possible to characterize the worldwide sheep populations [2]. Such characterization efforts identified the number of sheep breeds somewhere between 1000 [3] and 1400 [4].
The Original Valachian sheep a national, autochthonous breed from Slovakia. The small population of ca. 2900 animals is located in the northern mountainous regions, as it is well adapted to the harsh conditions of high-altitude pastures there. It was traditionally kept for combined triple-use: milk, meat and wool. The production is low, with ca. 80 kg milk per lactation, body weight of 35–55 kg, and average daily gain of 200 g until weaning. The Improved Valachian sheep was created to address these low production characteristics, while keeping the adaptation and climate resistance traits. A targeted and planned crossbreeding scheme was implemented between 1950s and 1980s, using Texel, Hampshire Down, Lincoln Longwool, and other breeds. The Improved Valachian sheep has higher milk production of ca. 100 kg per lactation, much higher body weight of 50 to 75 kg and average daily gain to weaning about 220 g.
Such diversity of breeds could evolve due to the wide variety of natural and artificial selection pressures on the populations. The selection itself leaves detectable changes on the genome caused by the so-called genetic hitchhiking effect [5]. Here, the selection on one gene with beneficial effect inevitably changes the frequencies of nearby neutral alleles. With consistent selection pressure and long enough time, the beneficial allele, along with the surrounding neutral alleles, spread throughout the population, reducing variability (thus increasing homozygosity) and strengthening linkage disequilibrium (LD) in the region. These signals could be then utilized to detect selection signatures to learn about the possible uniqueness of the populations.
One of the most widespread methods to detect the reduced variance in genomic regions of selection signatures is based on so-called runs of homozygosity (ROH) islands. The ROH are fully homozygous genomic segments that occur due to the relatedness of the two parents. Because of this assumption, they are often used to determine the inbreeding coefficient of animals [6]. However, from the evolutionary standpoint, they are often utilized to search for highly conserved regions. When the ROH segments overlap between a sizeable proportion of the population, this could indicate ongoing selection, thus the presence of beneficial genes. The method was used by Mastrangelo et al. [7], who identified the top 1% SNPs most commonly occurring in the homozygous regions as ROH islands in sheep. Based on the genomic coordinates of these regions, they have identified genes related to pigmentation, growth, milk, and carcass traits. In their earlier work Mastrangelo et al. [8] also used the ROH islands to identify genes with local adaptation in the Valle de Belice sheep. Additionally, Liu et al. [9] carried out a study, where they used ROH islands to identify signatures in five indigenous sheep breeds, based on the HD SNP BeadChip. The observed ROH pattern then helped to uncover the different population histories of the breeds. This application also demonstrates the wide scale of applicability of the method on finding interesting genomic regions in small, endangered breeds, which could be conserved and utilized in the future. Similarly, the ROH were utilized as the means of evaluating the diversity, and the ROH islands to identify selection signatures in South African sheep breeds [10].
Another possibility to explore the variation, and therefore the impact of selection on the genome is via patterns of LD. The LD is the population-wide correlation of nearby variants and characterizes the strength of connection between pairs of loci. The so-called LD blocks build up as a result of selection, via the genetic hitchhiking effect. In short, the allele at which the selection is acting is not inherited independently, but together with the loci that are nearby. If the selection pressure is strong enough, the allele is widespread in the population, together with its neighboring neutral alleles. This process results in a strong correlation, thus strong LD within this region. The differences in LD blocks between breeds than indicate differences in selection. This method was used to detect selection signatures in cattle [11,12,13], pigs [14] and horses [15]. Compared to other popular selection signature estimation methods, the studies using the variance in LD are sparse. Even more so in sheep, when even after our best efforts we could not find any suitable mention for this overview.
For our current study, we intend to fill the gap in the knowledge in genomic characterization of small indigenous breeds. We focus in particular on the exploration of diversity and selection signatures of the two Slovakian national sheep breeds, the Original Valachian and the Improved Valachian. They are an important animal genetic resource with diminishing population size; thus, identifying their characteristic and potentially valuable genomic regions is of great importance. Therefore, our aim in this paper is to explore the selection signatures within and between the Original and Improved Valachian sheep populations.

2. Materials and Methods

A total of 97 sheep (18 male and 79 female) from the Original Valachian, and 69 sheep (25 male and 44 female) from the Improved Valachian populations were genotyped using the GeneSeek GGP Ovine 50 K chip, with DNA obtained from hair follicles. The quality control was done using PLINK 1.9 [16], with the same thresholds for removal of SNPs and animals for both populations. Animals with more than 10% missing SNPs were removed. From the SNPs those with more than 10% missingness, below 1% minor allele frequency and not following the expected Hardy–Weinberg equilibrium distributions at p < 10−7 were excluded from the data set for the LD based selection signature analyses. The minor allele frequency threshold was not considered for run of homozygosity analyses. The SNPs from unplaced or sex chromosomes were removed for all analyses. After quality control, 96 animals and 38,236 SNPs remained for the Original Valachian sheep, and 68 animals and 39,240 SNPs were left for the Improved Valachian sheep population.
The inbreeding levels were computed using the software cgaTOH [17], based on ROH segments with minimum length of 1, 2, 4, 8 and 16 Mb. To minimize the occurrence of both false negative and false positive ROH segments, additional quality check was implemented based on Ferenčaković et al. [18]. In short, the number of false negative ROH segments were decreased by allowing certain number of missing, and in case of very long segments even heterozygous SNPs in the run. No missing or heterozygous SNPs were allowed for ROH with minimum length of 1 and 2 Mb, one missing and no heterozygous for ROH with minimum length of 4 Mb, two missing and no heterozygous for ROH with minimum length of 8 Mb, and four missing and one heterozygous for ROH with minimum length of 16 Mb. In addition, all ROH had to have at least 15 consecutive SNPs and no gaps larger than 1 Mb. The inbreeding coefficient for each sheep was computed as:
F R O H = k R O H k L
where k R O H k is the sum of the ROH length for each segment k, and L is the length of the autosomal genome covered by SNPs.
Further, the overlap between the ROH regions with minimum length of 1 Mb was examined for each breed. Such overlaps were called ROH islands, and were considered as selection signatures. The exact cut-off value for the ROH islands to be analyzed in the follow up was defined as the top 1% of most homozygous regions. The results were visualized using R, and the regions were compared to the sheep QTLdb (animalgenome.org, accessed on 20 October 2021) to determine their potential functions.
The linkage disequilibrium was computed as squared correlations using PLINK’s—r2 option, between each SNP within 1000 kb windows. The SNPs were ordered based on the inter marker distance, and the LD decay was visualized as the average within non-overlapping sliding windows of 5 SNPs.
In addition, the varLD software [19] was used to compare the LD patterns between the two breeds, in our case, those of the Original and Improved Valachian sheep. For this purpose, only the 37,788 SNPs were used that fulfilled the quality control criteria and were simultaneously present in both populations. The computation of standardized values for varLD scores, as well as the visualization, was performed in R, using the scripts provided by the varLD authors [19]. The signals above the top 1% of the standardized scores were highlighted as selection signatures between the two breeds, with further fine-tuning of the scales at the top 0.1% and 0.01%.

3. Results and Discussion

3.1. Inbreeding Levels

The inbreeding levels for the two populations are summarized in Table 1. Depending on the minimum length of ROH segments that were considered for the computation, the resulting genomic inbreeding coefficients consider common ancestors from various points in the past. The most recent inbreeding with common ancestors up to 3 generations was described considering the ROH segment length with minimum of 16 Mb. The inbreeding levels for larger time spans in the past were described by shorter segments. Here, the segment length with a minimum of 8, 4, 2 and 1 Mb described inbreeding levels to 6, 12, 25 and 50 generations in the past, respectively. Upon comparison of the two populations, the Improved Valachian sheep was much less inbred compared to the Original Valachian sheep population. These results indicate differences in management of both breeds, especially for the avoidance of recent inbreeding. While the inbreeding up to three generations is fairly low also for the Original Valachian sheep, it starts to be a concern when looking deeper in the pedigree of animals.

3.2. Within-Breed Selection Signatures Based on Run of Homozygosity Islands

The overlap of ROH segments with minimum length of 1 Mb in both populations was assessed and visualized in Figure 1 and Figure 2. The shortest ROH segments were chosen for the assessment of the ROH islands, so even the older selection signatures could be detected. The top 1% of the genomic regions that were covered by ROH the most often were further investigated, first for the occurrence of relevant quantitative trait loci (QTL) for the Original Valachian (Supplementary Table S1) and the Improved Valachian sheep (Supplementary Table S2). Although there was no specific requirement for the top 1% most homozygous SNPs to be neighboring on the genome, the most homozygous SNPs still tended to be near each other. We refer to these homozygous regions for the purposes of identifying QTL content.
In both breeds, there were a number of QTLs related to the main production of the breed, meat, milk and wool. These were represented by QTLs for both the amount of production alone, but also quality parameters, such as fat and protein content of milk, pH and color traits of meat, and similar. Interestingly, in the Original Valachian sheep, there we fewer, but larger regions, also containing a range of adaptation and disease resistance traits, such as response to parasitic infections. In the Improved Valachian sheep the most homozygous regions were more scattered, and not as widespread within the population (the top region was in a ROH for just 8 animals). Such results could be expected due to the history of crossbreeding, with admixture breaking down longer, almost fixed genomic regions into shorter segments. The QTL content of these smaller segments throughout the genome was clearly related to production, mostly milk and meat, also including their quality parameters, but also QTLs related to parasite resistance and health traits. On chromosome 10, around the highest signal, there were a number of QTLs related to horn traits. This might be of significance and could merit a further study, due to the peculiar horn shape of the Valachian sheep. Another interesting region in the Improved Valachian sheep was a small region on chromosome 25 (12.4–12.6 Mb), with a large number of QTL related to wool production. Although wool is currently not of great importance, the previous breeding process placed some emphasis on the improvement of this traits as well, which might have left some detectable signatures behind.
The much lower inbreeding (Table 1) in the Improved Valachian sheep was also visible in the overall homozygosity in the genome. In both cases the ROH islands were de-fined as the top 1% of most homozygous regions in the genome. However, such threshold was clearly different between the Original and Improved Valachian sheep. While in the Original Valachian sheep the 1% threshold was at overlapping segments for 21 individuals, in Improved Valachian it was only for four individuals. Such clear differences were likely caused by recent breeding history. The Improved Valachian sheep was created using the Original Valachian sheep as its base, and the targeted crossbreeding activities were applied between the 1950s and 1980s. The goal was to keep to overall characteristics, the high altitude adaptation, good grazing abilities and disease resistance traits of the autochthonous Original Valachian sheep. At the same time, the production traits of wool, milk and meat were upgraded using other established breeds such as Texel, Hampshire Down, Lincoln Longwool, Leicester Longwool, and East Friesian. Such events brought in many new parental haplotypes, making it less probable even for shorter ROH to occur in the Improved Valachian sheep.

3.3. Linkage Disequilibrium

Similarly to ROH results, the consequences of the applied breeding strategy were also visible for the comparison of LD decay in Figure 3. Additionally, here, the Improved Valachian sheep demonstrated a generally lower LD also in the case of SNPs in immediate vicinity. The LD values decrease much faster, to a considerably lower average values, compared to the Original Valachian sheep population. The LD extended to larger distances in the Original Valachian sheep, as it was undergoing selection for a much longer time period. In the Improved Valachian sheep the introduction of other breeds had a large effect on the LD structure in the population. It seems that not only the ancestral haplotypes in ROH were affected, but haplotypes across the genome were broken up. This breed improvement using other breeds largely affected the LD structure of the Improved Valachian sheep, with much lower LD across the genome, with the exception of SNPs very close to each other.

3.4. Between-Breed Selection Signatures Based on Variation of LD

After a breed-specific analysis of QTL functions in the identified regions, the two breeds were compared to identify regions where they differ from each other. Figure 4 shows the summary of these results. The differences between the correlation matrices were evaluated, and the regions that differ the most were highlighted as selection signatures. We provide visualization of thresholds for the top 1%, 0.1% and 0.01% of the signals in Figure 4, as well as Supplementary Figure S1. The top 1% of the signals included the 29.1 to 30.4 Mb region on chromosome 1, a wide region spanning from 31.7 to 43.4 Mb on chromosome 2, as well as a single SNP at 220.4 Mb of chromosome 2. Additionally, two single SNPs were appearing above the 1% threshold at chromosome 9 at 36.0 and 67.7 Mb, as well as a wider region between 42.7 and 44.9 Mb on chromosome 9. An additional single SNP was above the 1% threshold on chromosome 13, at 64.8 Mb. On chromosome 16, a wider region was identified, between 16.0 and 17.7 Mb. A somewhat wider region was found on chromosome 17, between 30.5 and 33.2 Mb. Other signals above the 1% threshold were found between 37.6 and 37.8 Mb of chromosome 18 and at 15.0 Mb of chromosome 25.
As for the top 0.1% highest signal, there was only one region on chromosome 20, between 29.9 and 32.7 Mb. There seems to be a gap in the signal between 30.9 and 31.9 Mb. This relatively wide signal does not have a pronounced peak, indicating that the whole genomic region could be of interest, rather than any specific point. The SNP indicating the highest signal within this region was located at 30.7 Mb. The top signal for the entire genome was found on chromosome 22. The top 0.1% region was found between 10.3 and 11.6 Mb. The region exceeding the 0.01% threshold was found between 10.8 and 11.0 Mb on chromosome 22.
In general, the more relaxed the thresholds were, the wider the section signatures became. This could be a possible reason why other studies, such as [11] and [14] evaluated only the top 0.1% and 0.01% signals. For both identified regions on chromosome 16 and 17, the top of these were close to the 0.1% threshold, thus their detailed exploration might be justified.
Our top signal directly overlaps with the genes LIPA, IFIT2, IFIT3, IFIT5, SLC16A12 and three pseudogenes. The LIPA gene was found to be involved in the lipid mechanism and composition of lipoproteins [20], response to wounds and inflammations [21], but also in molecular genetic mechanisms affecting fecundity in sheep [22]. The family of IFIT genes was shown to have antiviral functions [23]. Similarly, in sheep, it was shown to influence parasite resistance in Merino sheep populations [24]. The SLC16A12 gene was also shown to have various immune related functions, related to bacterial infections [25,26].
A large number of pseudogenes and protein coding genes were found in the other regions identified by the VarLD approach. Here, they are listed according to the chromosome number for the regions indicated above, and those with possible larger relevance are further described.
For chromosome 1, these protein coding genes were USP24, PHF11, DHCR24, TTC22 and TTC4. The gene PHF11 appears to inhibit replication of certain retroviruses [27] and play a role in activation of the T-cells in the immune system [28]. The DHCR24 appears to have an important function related to the biosynthesis of cholesterol, but also involved in neurological disorders, such as the Alzheimer disease [29].
On chromosome 2, there was a very wide signal harboring a large number of protein coding genes, such as DAPK1, FBP1, FBP2, CTSL, ISCA1, GOLM1, NTRK2, SLC28A3, RMI1, HNRNPK, KIF27, GKAP1, UBQLN1, FRMD3, IL11RA, ARID3C, DCTN3, DNAI1, C9orf24, NUDT2, KIF24, UBAP1, DCAF12, UBAP2, UBE2R2, NOL6, AQP3, AQP7, NFX1, BAG1, B4GALT1, DNAJA1, APTX, EPHX2, CHRNA2, PTK2B, DPYSL2, BNIP3L, PPP2R2A, EBF2, CDCA2, DOCK5, NEFL, NEFM, ADAM7, ADAM28, STC1, LOXL2, ENTPD4, RHOBTB2, PEBP4, EGR3, BIN3, SORBS3, DLIM2, PPP3CC, SLC39A14, PIWIL2. Even though it is a very wide region with a high number of genes, there is a remarkable diversity in gene functions, which is relevant for productive, reproductive and health aspect of animals’ life. These associations were highlighted in a wide range of studies. Unfortunately, there were only a few studies done directly in sheep, but relevant studies were found in humans, but also in livestock, mostly cattle.
The DAPK1 gene had an identified relationship to somatic cell score in cattle [30]. The FBP1 gene plays an important role in the glucosynthesis [31], while the FBP2 gene was shown to be related to milk productions traits [32], activating various metabolic processes in time of increased demands of high milk production [33]. The gene CTSL (cathepsin L) seems to be interesting from multiple angles, as it had been shown to influence carcass traits, production of milk [34] and reproductive traits [35]. The ISCA1 was found to be associated with multiple mitochondrial dysfunctions [36]. According to a very recent study, the NTRK2 was shown to be a novel candidate gene for litter size in sheep [37]. RMI1 was shown to be associated with growth in sheep [38]. The KIF27 gene was involved in fertility [39]. The gene UBQLN1 is involved male fertility, more precisely in ubiquitination during spermatogenesis [40]. The IL11RA gene was also shown to be reproduction related, with its involvement in the placentation in multiple species [41]. ARID3C was involved in immunoglobulin gene transcription [42]. The DNAI1 gene was identified to be relevant for spermatogenesis in cattle [43]. The UBAP2 gene was found to be implicated in hypoxia and heat stress in sheep and goats [44]. In addition, in cattle it was in an LD block involved in wound healing, skin lesions, bone growth and mineralization [45]. The cold acclimatization is likely regulated by the AQP3 and APQ7 gene [46], also in this region. Such gene functions in the selection signatures in the Valachian sheep are well in line with its adaptation of high-altitude pastures. The B4GALT1 gene was found to be associated with milk production in cattle [47]. The DNAJA1 gene was highly expressed in both mammary gland and milk transcriptome, suggesting involvement in milk synthesis and production [48]. At the same time, it was found to be involved in meat tenderness in cattle [49]. Similarly, the nearby APTX gene was found in a study for marbling score in cattle [50]. The PTK2B gene was recently found to be involved in regulation of bovine mastitis [51]. The DPYSL2 gene, also known as CRMP2, is a gene of major interest, as it could be involved in an extensive collection of neurodegenerative, sensory and motor neuron, and central disorders [52], which could be especially relevant for scrapie susceptibility in sheep. The BNIP3L, PPP2R2A, CDCA2 and NEFL genes were related to reproductive traits, namely, to ovarian follicle development and with lambings out of season [53]. The EBF2 was shown to be involved in regulation of body fat [54]. The ADAM7 and ADAM28 genes were previously found in genomic regions related to male fertility. In particular, ADAM28 showed significant non-additive effects for sire conception rate in cattle [55]. The female reproduction via the uterine biology was influenced by the STC1 gene [56]. The LOXL2 gene was found to be related to meat production in cattle [12]. The ENTPD4 gene seems to be influencing both milk production [57] and fertility [58]. Another gene from this region, which seems to be involved in multiple traits, was PEBP4 which was found to simultaneously affect growth and fertility in cattle [59]. The gene SORBS3, PPP3CC and PIWIL2 were located in a region associated with growth and wool traits in [60]. In addition, the PPP3CC gene was linked to heat tolerance in cattle [61], and the PIWIL2 was associated with reproduction traits [62]. The SLC39A14 gene was found to be associated with feed efficiency in cattle [63].
The second signal on chromosome was only a single significant SNP at 220.4 Mb. The genes around this SNP were: TMBIM1 and PNKD, connected to meat color in cattle [64]; USP37, involved in regulation of reproduction in mice [65]; VIL1, related to meat tenderness [66]; CTDSP1, related to milk fatty acid synthesis in goats [67]; SLC11A1, with significant disease resistance related effects, in particular paratuberculosis, tuberculosis and brucellosis in a number of livestock species [68,69].
The next signal was on chromosome 9, around 36Mb, where the genes such as LYN, PLAG1, RPS20 and MOS were located. From these the most important was the PLAG1, with a widely recognized effect on stature and growth traits in multiple species, including sheep [70,71]. The other genes in the region were also related to body size (LYN, MOS) but also fertility and dystocia (RPS20) [72]. The second signal indicated just by a single SNP on chromosome 9 was around 67.7Mb around the SYBU gene, which was related to milk fat synthesis, transport, and metabolism [73,74].
The wider region on chromosome 9 between 42.7 and 44.9 Mb contained protein coding genes such as PDE7A, MTFR1, DNAJC5B, TRIM55, ADHFE1, CSPP1, COPS5, ARFGEF1, and CPA6. From these genes, the most important were the PDE7A related to growth traits, through its significantly upregulation in lambs with muscle hypertrophy, callipyge [75]. The TRIM55 gene was upregulated for beef tenderness, juiciness, and flavor [76]. The COPS5 gene seems to be involved in multiple pathways, as it regulates the endometrial function in goats [77], was shown to be involved in mitochondrial functions [78], and carcass and meat quality traits [79]. The ARFGEF1, and CPA6 were involved in weight and carcass traits in cattle [80].
On chromosome 13 at 64.8 Mb were the following genes: SPAG4, NFS1, and ROMO1. The SPAG4 gene was shown to be connected to reproduction traits and spermatogenesis [81,82].
On the signal on chromosome 16, the KIF2A gene was located, which plays an important role during the meiosis, and its downregulation might interfere with the spermatogenesis [83]. One other gene of importance in this region was the IPO11, which was shown to be upregulated in mastitis [84], but also associated with displaced abomasum in cattle [85].
The signal on chromosome 17 was remarkably sparse, with almost no protein coding genes along the 3 Mb genomic region. The only gene in the region was the FAT4 gene, associated with milk fat composition [86].
The signal on chromosome 18 at 37.8 Mb contains the gene PRKD1, associated with BMI and height in humans [87,88], but also associated with pubertal age in pigs [89].
Chromosome 25 contained a narrow signal indicated by two nearby SNPs at around 15 Mb. This region contained only two protein coding genes, ANK3 and CCDC6. The CCDC6 gene was of particular interest, as it was previously associated with body weight [59], but also in a gene network contributing to hypertrophy in callipyge skeletal muscle [90].

4. Conclusions

In the study, we carried out analysis of selection signatures within the Original and Improved Valachian sheep populations based on homozygosity and LD patterns on their genome. The results could be used in practice to improve the genomic characterization of the two breeds. The unique selection signatures for the respective breeds could serve as to clearly distinguish between the breeds on a genomic level, and aiding the decisions related to the conservation of genetic diversity. Overall, we have identified selection signatures with QTLs and genes pointing towards all three production purposes of the Valachian sheep, milk, meat and wool, including their quality characteristics. Another group with apparent large importance was the various traits related to health traits and resistance to parasites, which is well in line with the sturdy nature of this breed, kept mostly on high altitude pastures. More importantly, the genomic regions related to health and disease resistance, along with others pointing to various production traits, were also apparent in the Improved Valachian sheep. This indicates that the improvement process was a success that brought in increased productive capacity while retaining the adaptation characteristics of this Slovak traditional breed.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ani12111346/s1, Figure S1: Standardized VarLD scores per chromosome, Table S1: QTL content based on the Sheep QTLdb from the top 1% most homozygous regions in the Original Valachian sheep population, Table S2: QTL content based on the Sheep QTLdb from the top 1% most homozygous regions in the Improved Valachian sheep population

Author Contributions

Conceptualization, R.K., N.M. and G.M.; methodology, G.M., N.M. and M.M. (Mária Mészárosová); validation, G.M., N.M. and M.M. (Mária Mészárosová); formal analysis, M.M. (Mária Mészárosová); investigation, M.M. (Mária Mészárosová); resources, M.M. (Milan Margetín). and I.P.; data curation, M.M. (Mária Mészárosová) and N.M.; writing—original draft preparation, M.M. (Mária Mészárosová); writing—review and editing, G.M., R.K., N.M., M.M. (Milan Margetín), I.P.; visualization, M.M. (Mária Mészárosová); supervision, R.K.; project administration, R.K., N.M.; funding acquisition, R.K., N.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Slovak Agency for Research and Development, grants number APVV-17-0060 and APVV-20-0161.

Institutional Review Board Statement

This study did not require ethical approval.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to ongoing research.

Acknowledgments

The authors wish to acknowledge the support of breeders and their willingness to provide samples for genotyping.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study, in the collection, analyses, or interpretation of data, in the writing of the manuscript, or in the decision to publish the results.

References

  1. Zeder, M.A. Domestication and early agriculture in the Mediterranean Basin: Origins, diffusion, and impact. Proc. Natl. Acad. Sci. USA 2008, 105, 11597–11604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Kijas, J.W.; Townley, D.; Dalrymple, B.P.; Heaton, M.P.; Maddox, J.F.; McGrath, A.; Wilson, P.; Ingersoll, R.G.; McCulloch, R.; McWilliam, S.; et al. A Genome Wide Survey of SNP Variation Reveals the Genetic Structure of Sheep Breeds. PLoS ONE 2009, 4, e4668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Kijas, J.W.; Lenstra, J.A.; Hayes, B.; Boitard, S.; Neto, L.R.P.; Cristobal, M.S.; Servin, B.; McCulloch, R.; Whan, V.; Gietzen, K.; et al. Genome-Wide Analysis of the World’s Sheep Breeds Reveals High Levels of Historic Mixture and Strong Recent Selection. PLoS Biol. 2012, 10, e1001258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Scherf, B.D. World Watch List for Domestic Animal Diversity; Food and Agriculture Organization (FAO): Rome, Italy, 2000; Available online: https://www.cabdirect.org/cabdirect/abstract/20013016397 (accessed on 28 November 2021).
  5. Charlesworth, B.; Harvey, P.H.; Barton, N.H. Genetic hitchhiking. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2000, 355, 1553–1562. [Google Scholar] [CrossRef]
  6. Curik, I.; Ferenčaković, M.; Sölkner, J. Inbreeding and runs of homozygosity: A possible solution to an old problem. Livest. Sci. 2014, 166, 26–34. [Google Scholar] [CrossRef]
  7. Mastrangelo, S.; Ciani, E.; Sardina, M.T.; Sottile, G.; Pilla, F.; Portolano, B.; Bi.Ov. Ita Consortium. Runs of homozygosity reveal genome-wide autozygosity in Italian sheep breeds. Anim. Genet. 2018, 49, 71–81. [Google Scholar] [CrossRef] [Green Version]
  8. Mastrangelo, S.; Tolone, M.; Sardina, M.T.; Sottile, G.; Sutera, A.M.; Di Gerlando, R.; Portolano, B. Genome-wide scan for runs of homozygosity identifies potential candidate genes associated with local adaptation in Valle del Belice sheep. Genet. Sel. Evol. 2017, 49, 84. [Google Scholar] [CrossRef] [Green Version]
  9. Liu, J.; Shi, L.; Li, Y.; Chen, L.; Garrick, D.; Wang, L.; Zhao, F. Estimates of genomic inbreeding and identification of candidate regions that differ between Chinese indigenous sheep breeds. J. Anim. Sci. Biotechnol. 2021, 12, 95. [Google Scholar] [CrossRef]
  10. Dzomba, E.F.; Chimonyo, M.; Pierneef, R.; Muchadeyi, F.C. Runs of homozygosity analysis of South African sheep breeds from various production systems investigated using OvineSNP50k data. BMC Genom. 2021, 22, 7. [Google Scholar] [CrossRef]
  11. Pérez O’Brien, A.M.; Utsunomiya, Y.T.; Mészáros, G.; Bickhart, D.M.; Liu, G.E.; Van Tassell, C.P.; Sonstegard, T.S.; Da Silva, M.V.; Garcia, J.F.; Sölkner, J. Assessing signatures of selection through variation in linkage disequilibrium between taurine and indicine cattle. Genet. Sel. Evol. 2014, 46, 19. [Google Scholar] [CrossRef] [Green Version]
  12. Sorbolini, S.; Marras, G.; Gaspa, G.; Dimauro, C.; Cellesi, M.; Valentini, A.; Macciotta, N.P. Detection of selection signatures in Piemontese and Marchigiana cattle, two breeds with similar production aptitudes but different selection histories. Genet. Sel. Evol. 2015, 47, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Kukučková, V.; Moravčíková, N.; Kasarda, R. Evaluating signatures of selection through variation in linkage disequilibrium among different cattle breeds. Acta Fytotech. Zootech. 2016, 19, 105–110. [Google Scholar] [CrossRef] [Green Version]
  14. Li, X.; Yang, S.; Dong, K.; Tang, Z.; Li, K.; Fan, B.; Wang, Z.; Liu, B. Identification of positive selection signatures in pigs by comparing linkage disequilibrium variances. Anim. Genet. 2017, 48, 600–605. [Google Scholar] [CrossRef] [PubMed]
  15. Han, H.; McGivney, B.A.; Farries, G.; Katz, L.M.; MacHugh, D.E.; Randhawa, I.A.S.; Hill, E.W. Selection in Australian Thoroughbred horses acts on a locus associated with early two-year old speed. PLoS ONE 2020, 15, e0227212. [Google Scholar] [CrossRef] [Green Version]
  16. Chang, C.C.; Chow, C.C.; Tellier, L.C.; Vattikuti, S.; Purcell, S.M.; Lee, J.J. Second-generation PLINK: Rising to the challenge of larger and richer datasets. GigaScience 2015, 4, 7. [Google Scholar] [CrossRef]
  17. Zhang, L.; Orloff, M.S.; Reber, S.; Li, S.; Zhao, Y.; Eng, C. cgaTOH: Extended Approach for Identifying Tracts of Homozygosity. PLoS ONE 2013, 8, e57772. [Google Scholar] [CrossRef]
  18. Ferenčaković, M.; Sölkner, J.; Curik, I. Estimating autozygosity from high-throughput information: Effects of SNP density and genotyping errors. Genet. Sel. Evol. 2013, 45, 42. [Google Scholar] [CrossRef] [Green Version]
  19. Ong, R.T.-H.; Teo, Y.-Y. varLD: A program for quantifying variation in linkage disequilibrium patterns between populations. Bioinformatics 2010, 26, 1269–1270. [Google Scholar] [CrossRef]
  20. Arnaboldi, L.; Ossoli, A.; Giorgio, E.; Pisciotta, L.; Lucchi, T.; Grigore, L.; Pavanello, C.; Granata, A.; Pasta, A.; Arosio, B.; et al. LIPA gene mutations affect the composition of lipoproteins: Enrichment in ACAT-derived cholesteryl esters. Atherosclerosis 2020, 297, 8–15. [Google Scholar] [CrossRef]
  21. Deng, X.; Wang, D.; Wang, S.; Wang, H.; Zhou, H. Identification of key genes and pathways involved in response to pain in goat and sheep by transcriptome sequencing. Biol. Res. 2018, 51, 25. [Google Scholar] [CrossRef]
  22. Tang, J.; Hu, W.; Chen, S.; Di, R.; Liu, Q.; Wang, X.; He, X.; Gan, S.; Zhang, X.; Zhang, J.; et al. The genetic mechanism of high prolificacy in small tail han sheep by comparative proteomics of ovaries in the follicular and luteal stages. J. Proteom. 2019, 204, 103394. [Google Scholar] [CrossRef] [PubMed]
  23. Diamond, M.S.; Farzan, M. The broad-spectrum antiviral functions of IFIT and IFITM proteins. Nat. Rev. Immunol. 2013, 13, 46–57. [Google Scholar] [CrossRef]
  24. Zhang, R.; Liu, F.; Hunt, P.; Li, C.; Zhang, L.; Ingham, A.; Li, R.W. Transcriptome analysis unraveled potential mechanisms of resistance to Haemonchus contortus infection in Merino sheep populations bred for parasite resistance. Vet. Res. 2019, 50, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Jonczyk, M.S.; Simon, M.; Kumar, S.; Fernandes, V.E.; Sylvius, N.; Mallon, A.-M.; Denny, P.; Andrew, P.W. Genetic Factors Regulating Lung Vasculature and Immune Cell Functions Associate with Resistance to Pneumococcal Infection. PLoS ONE 2014, 9, e89831. [Google Scholar] [CrossRef]
  26. Mousel, M.R.; White, S.N.; Herndon, M.K.; Herndon, D.R.; Taylor, J.B.; Becker, G.M.; Murdoch, B.M. Genes involved in immune, gene translation and chromatin organization pathways associated with Mycoplasma ovipneumoniae presence in nasal secretions of domestic sheep. PLoS ONE 2021, 16, e0247209. [Google Scholar] [CrossRef] [PubMed]
  27. Kane, M.; Mele, V.; Liberatore, R.A.; Bieniasz, P.D. Inhibition of spumavirus gene expression by PHF11. PLoS Pathog. 2020, 16, e1008644. [Google Scholar] [CrossRef]
  28. Rahman, N.; Stewart, G.; Jones, G. A role for the atopy-associated gene PHF11 in T-cell activation and viability. Immunol. Cell Biol. 2010, 88, 817–824. [Google Scholar] [CrossRef]
  29. Crameri, A.; Biondi, E.; Kuehnle, K.; Lütjohann, D.; Thelen, K.M.; Perga, S.; Dotti, C.G.; Nitsch, R.M.; Ledesma, M.D.; Mohajeri, M.H. The role of seladin-1/DHCR24 in cholesterol biosynthesis, APP processing and Abeta generation in vivo. EMBO J. 2006, 25, 432–443. [Google Scholar] [CrossRef] [Green Version]
  30. Ilie, D.E.; Mizeranschi, A.E.; Mihali, C.V.; Neamț, R.I.; Goilean, G.V.; Georgescu, O.I.; Zaharie, D.; Carabaș, M.; Huțu, I. Genome-Wide Association Studies for Milk Somatic Cell Score in Romanian Dairy Cattle. Genes 2021, 12, 1495. [Google Scholar] [CrossRef]
  31. Granner, D.; Pilkis, S. The genes of hepatic glucose metabolism. J. Biol. Chem. 1990, 265, 10173–10176. [Google Scholar] [CrossRef]
  32. Vineeth, M.R.; Surya, T.; Sivalingam, J.; Kumar, A.; Niranjan, S.K.; Dixit, S.P.; Singh, K.; Tantia, M.S.; Gupta, I.D. Genome-wide discovery of SNPs in candidate genes related to production and fertility traits in Sahiwal cattle. Trop. Anim. Health Prod. 2020, 52, 1707–1715. [Google Scholar] [CrossRef] [PubMed]
  33. Weikard, R.; Goldammer, T.; Brunner, R.M.; Kuehn, C. Tissue-specific mRNA expression patterns reveal a coordinated metabolic response associated with genetic selection for milk production in cows. Physiol. Genom. 2012, 44, 728–739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Medrano, J.F.; Sharrow, L. Milk Protein Typing of Bovine Mammary Gland Tissue Used to Generate a Complementary Deoxyribonucleic Acid Library. J. Dairy Sci. 1989, 72, 3190–3196. [Google Scholar] [CrossRef]
  35. Baba, N.A.; Panigrahi, M.; Verma, A.D.; Sadam, A.; Sulabh, S.; Chhotaray, S.; Parida, S.; Krishnaswamy, N.; Bhushan, B. Endometrial transcript profile of progesterone-regulated genes during early pregnancy of Water Buffalo (Bubalus bubalis). Reprod. Domest. Anim. 2019, 54, 100–107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Shukla, A.; Hebbar, M.; Srivastava, A.; Kadavigere, R.; Upadhyai, P.; Kanthi, A.; Brandau, O.; Bielas, S.; Girisha, K.M. Homozygous p.(Glu87Lys) variant in ISCA1 is associated with a multiple mitochondrial dysfunctions syndrome. J. Hum. Genet. 2017, 62, 723–727. [Google Scholar] [CrossRef] [Green Version]
  37. Esmaeili-Fard, S.M.; Gholizadeh, M.; Hafezian, S.H.; Abdollahi-Arpanahi, R. Genome-wide association study and pathway analysis identify NTRK2 as a novel candidate gene for litter size in sheep. PLoS ONE 2021, 16, e0244408. [Google Scholar] [CrossRef]
  38. Saravanan, K.A.; Panigrahi, M.; Kumar, H.; Bhushan, B.; Dutt, T.; Mishra, B.P. Genome-wide analysis of genetic diversity and selection signatures in three Indian sheep breeds. Livest. Sci. 2021, 243, 104367. [Google Scholar] [CrossRef]
  39. Parker Gaddis, K.L.; Null, D.J.; Cole, J.B. Explorations in genome-wide association studies and network analyses with dairy cattle fertility traits. J. Dairy Sci. 2016, 99, 6420–6435. [Google Scholar] [CrossRef]
  40. Bao, J.; Zhang, J.; Zheng, H.; Xu, C.; Yan, W. UBQLN1 interacts with SPEM1 and participates in spermiogenesis. Mol. Cell. Endocrinol. 2010, 327, 89. [Google Scholar] [CrossRef] [Green Version]
  41. Østrup, E.; Hyttel, P.; Østrup, O.; Østrup, E.; Hyttel, P.; Østrup, O. Embryo–maternal communication: Signalling before and during placentation in cattle and pig. Reprod. Fertil. Dev. 2011, 23, 964–975. [Google Scholar] [CrossRef]
  42. Tidwell, J.A.; Schmidt, C.; Heaton, P.; Wilson, V.; Tucker, P.W. Characterization of a new ARID family transcription factor (Brightlike/ARID3C) that co-activates Bright/ARID3A-mediated immunoglobulin gene transcription. Mol. Immunol. 2011, 49, 260–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Li, G.; Peñagaricano, F.; Weigel, K.A.; Zhang, Y.; Rosa, G.; Khatib, H. Comparative genomics between fly, mouse, and cattle identifies genes associated with sire conception rate. J. Dairy Sci. 2012, 95, 6122–6129. [Google Scholar] [CrossRef] [PubMed]
  44. Serranito, B.; Cavalazzi, M.; Vidal, P.; Taurisson-Mouret, D.; Ciani, E.; Bal, M.; Rouvellac, E.; Servin, B.; Moreno-Romieux, C.; Tosser-Klopp, G.; et al. Local adaptations of Mediterranean sheep and goats through an integrative approach. Sci. Rep. 2021, 11, 21363. [Google Scholar] [CrossRef] [PubMed]
  45. Lai, E.; Danner, A.L.; Famula, T.R.; Oberbauer, A.M. Genome-Wide Association Studies Reveal Susceptibility Loci for Noninfectious Claw Lesions in Holstein Dairy Cattle. Front. Genet. 2021, 12, 657375. [Google Scholar] [CrossRef]
  46. Ghoreishifar, S.M.; Eriksson, S.; Johansson, A.M.; Khansefid, M.; Moghaddaszadeh-Ahrabi, S.; Parna, N.; Davoudi, P.; Javanmard, A. Signatures of selection reveal candidate genes involved in economic traits and cold acclimation in five Swedish cattle breeds. Genet. Sel. Evol. 2020, 52, 52. [Google Scholar] [CrossRef]
  47. Valsalan, J.; Sadan, T.; Venkatachalapathy, T.; Anilkumar, K.; Aravindakshan, T.V. Identification of novel single-nucleotide polymorphism at exon1 and 2 region of B4GALT1 gene and its association with milk production traits in crossbred cattle of Kerala, India. Anim. Biotechnol. 2021, 1–9. [Google Scholar] [CrossRef]
  48. Banos, G.; Clark, E.L.; Bush, S.J.; Dutta, P.; Bramis, G.; Arsenos, G.; Hume, D.A.; Psifidi, A. Genomic analyses underpin the feasibility of concomitant genetic improvement of milk yield and mastitis resistance in dairy sheep. PLoS ONE 2019, 14, e0214346. [Google Scholar] [CrossRef] [Green Version]
  49. Malheiros, J.M.; Enríquez-Valencia, C.E.; da Silva Duran, B.O.; de Paula, T.G.; Curi, R.A.; de Vasconcelos Silva, J.A.I.I.; Dal-Pai-Silva, M.; de Oliveira, H.N.; Chardulo, L.A.L. Association of CAST2, HSP90AA1, DNAJA1 and HSPB1 genes with meat tenderness in Nellore cattle. Meat Sci. 2018, 138, 49–52. [Google Scholar] [CrossRef]
  50. Lim, D.; Lee, S.-H.; Kim, N.-K.; Cho, Y.-M.; Chai, H.-H.; Seong, H.-H.; Kim, H. Gene Co-expression Analysis to Characterize Genes Related to Marbling Trait in Hanwoo (Korean) Cattle. Asian-Aust. J. Anim. Sci. 2013, 26, 19–29. [Google Scholar] [CrossRef]
  51. Yang, F.; Chen, F.; Li, L.; Yan, L.; Badri, T.; Lv, C.; Yu, D.; Zhang, M.; Jang, X.; Li, J.; et al. Three Novel Players: PTK2B, SYK, and TNFRSF21 Were Identified to Be Involved in the Regulation of Bovine Mastitis Susceptibility via GWAS and Post-transcriptional Analysis. Front. Immunol. 2019, 10, 1579. [Google Scholar] [CrossRef] [Green Version]
  52. Khanna, R.; Wilson, S.M.; Brittain, J.M.; Weimer, J.; Sultana, R.; Butterfield, A.; Hensley, K. Opening Pandora’s jar: A primer on the putative roles of CRMP2 in a panoply of neurodegenerative, sensory and motor neuron, and central disorders. Future Neurol. 2012, 7, 749–771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Posbergh, C.J.; Thonney, M.L.; Huson, H.J. Genomic Approaches Identify Novel Gene Associations with Out of Season Lambing in Sheep. J. Hered. 2019, 110, 577–586. [Google Scholar] [CrossRef] [PubMed]
  54. Stine, R.R.; Shapira, S.N.; Lim, H.-W.; Ishibashi, J.; Harms, M.; Won, K.-J.; Seale, P. EBF2 promotes the recruitment of beige adipocytes in white adipose tissue. Mol. Metab. 2016, 5, 57–65. [Google Scholar] [CrossRef] [PubMed]
  55. Nicolini, P.; Amorín, R.; Han, Y.; Peñagaricano, F. Whole-genome scan reveals significant non-additive effects for sire conception rate in Holstein cattle. BMC Genet. 2018, 19, 14. [Google Scholar] [CrossRef] [PubMed]
  56. Bazer, F.W.; Song, G.; Kim, J.; Dunlap, K.A.; Satterfield, M.C.; Johnson, G.A.; Burghardt, R.C.; Wu, G. Uterine biology in pigs and sheep. J. Anim. Sci. Biotechnol. 2012, 3, 23. [Google Scholar] [CrossRef] [Green Version]
  57. Connor, E.E.; Siferd, S.; Elsasser, T.H.; Evock-Clover, C.M.; Van Tassell, C.P.; Sonstegard, T.S.; Fernandes, V.M.; Capuco, A.V. Effects of increased milking frequency on gene expression in the bovine mammary gland. BMC Genom. 2008, 9, 362. [Google Scholar] [CrossRef]
  58. Abdollahi-Arpanahi, R.; Pacheco, H.A.; Peñagaricano, F. Targeted sequencing reveals candidate causal variants for dairy bull subfertility. Anim. Genet. 2021, 52, 509–513. [Google Scholar] [CrossRef]
  59. Kour, A.; Deb, S.M.; Nayee, N.; Niranjan, S.K.; Raina, V.S.; Mukherjee, A.; Gupta, I.D.; Patil, C.S. Novel insights into genome-wide associations in Bos indicus reveal genetic linkages between fertility and growth. Anim. Biotechnol. 2021, 1–17. [Google Scholar] [CrossRef]
  60. Mohammadi, H.; Rafat, S.A.; Moradi Shahrbabak, H.; Shodja, J.; Moradi, M.H. Genome-wide association study and gene ontology for growth and wool characteristics in Zandi sheep. J. Livest. Sci. Technol. 2020, 8, 45–55. [Google Scholar] [CrossRef]
  61. Li, R.; Li, C.; Chen, H.; Li, R.; Chong, Q.; Xiao, H.; Chen, S. Genome-wide scan of selection signatures in Dehong humped cattle for heat tolerance and disease resistance. Anim. Genet. 2020, 51, 292–299. [Google Scholar] [CrossRef]
  62. Fischer, D.; Laiho, A.; Gyenesei, A.; Sironen, A. Identification of Reproduction-Related Gene Polymorphisms Using Whole Transcriptome Sequencing in the Large White Pig Population. G3 Genes Genomes Genet. 2015, 5, 1351–1360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Yang, C.; Zhu, Y.; Ding, Y.; Huang, Z.; Dan, X.; Shi, Y.; Kang, X. Identifying the key genes and functional enrichment pathways associated with feed efficiency in cattle. Gene 2022, 807, 145934. [Google Scholar] [CrossRef] [PubMed]
  64. Marín-Garzón, N.A.; Magalhães, A.F.B.; Mota, L.F.M.; Fonseca, L.F.S.; Chardulo, L.A.L.; Albuquerque, L.G. Genome-wide association study identified genomic regions and putative candidate genes affecting meat color traits in Nellore cattle. Meat Sci. 2021, 171, 108288. [Google Scholar] [CrossRef]
  65. Sakai, K.; Ito, C.; Wakabayashi, M.; Kanzaki, S.; Ito, T.; Takada, S.; Toshimori, K.; Sekita, Y.; Kimura, T. Usp26 mutation in mice leads to defective spermatogenesis depending on genetic background. Sci. Rep. 2019, 9, 13757. [Google Scholar] [CrossRef] [PubMed]
  66. Lee, K.-T.; Chung, W.-H.; Lee, S.-Y.; Choi, J.-W.; Kim, J.; Lim, D.; Lee, S.; Jang, G.-W.; Kim, B.; Choy, Y.H.; et al. Whole-genome resequencing of Hanwoo (Korean cattle) and insight into regions of homozygosity. BMC Genom. 2013, 14, 519. [Google Scholar] [CrossRef] [Green Version]
  67. Wang, H.; Zhu, J.; He, Q.; Loor, J.J.; Luo, J. Association between the expression of miR-26 and goat milk fatty acids. Reprod. Domest. Anim. 2018, 53, 1478–1482. [Google Scholar] [CrossRef] [PubMed]
  68. Kumar, N.; Ganguly, I.; Singh, R.; Deb, S.M.; Kumar, S.; Sharma, A.; Mitra, A. DNA Polymorphism in SLC11A1 Gene and its Association with Brucellosis Resistance in Indian Zebu (Bos indicus) and Crossbred (Bos indicus×Bos taurus) Cattle. Asian-Aust. J. Anim. Sci. 2011, 24, 898–904. [Google Scholar] [CrossRef]
  69. Taka, S.; Liandris, E.; Gazouli, M.; Sotirakoglou, K.; Theodoropoulos, G.; Bountouri, M.; Andreadou, M.; Ikonomopoulos, J. In vitro expression of the SLC11A1 gene in goat monocyte-derived macrophages challenged with Mycobacterium avium subsp paratuberculosis. Infect. Genet. Evol. 2013, 17, 8–15. [Google Scholar] [CrossRef]
  70. Wu, H.; Pan, Y.; Zhang, Q.; Cao, Y.; Li, J.; Chen, H.; Cai, Y.; Sun, X.; Lan, X. Insertion/deletion (InDel) variations in sheep PLAG1 gene locating in growth-related major QTL are associated with adult body weight and morphometric traits. Small Rumin. Res. 2019, 178, 63–69. [Google Scholar] [CrossRef]
  71. Pan, Y.; Wang, M.; Wu, H.; Akhatayeva, Z.; Lan, X.; Fei, P.; Mao, C.; Jiang, F. Indel mutations of sheep PLAG1 gene and their associations with growth traits. Anim. Biotechnol. 2021, 1–7. [Google Scholar] [CrossRef]
  72. Taye, M.; Yoon, J.; Dessie, T.; Cho, S.; Oh, S.J.; Lee, H.-K.; Kim, H. Deciphering signature of selection affecting beef quality traits in Angus cattle. Genes Genom. 2018, 40, 63–75. [Google Scholar] [CrossRef] [PubMed]
  73. Zhou, C.; Shen, D.; Li, C.; Cai, W.; Liu, S.; Yin, H.; Shi, S.; Cao, M.; Zhang, S. Comparative Transcriptomic and Proteomic Analyses Identify Key Genes Associated With Milk Fat Traits in Chinese Holstein Cows. Front. Genet. 2019, 10, 672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Vohra, V.; Chhotaray, S.; Gowane, G.; Alex, R.; Mukherjee, A.; Verma, A.; Deb, S.M. Genome-Wide Association Studies in Indian Buffalo Revealed Genomic Regions for Lactation and Fertility. Front. Genet. 2021, 12, 1668. [Google Scholar] [CrossRef] [PubMed]
  75. Fleming-Waddell, J.N.; Olbricht, G.R.; Taxis, T.M.; White, J.D.; Vuocolo, T.; Craig, B.A.; Tellam, R.L.; Neary, M.K.; Cockett, N.E.; Bidwell, C.A. Effect of DLK1 and RTL1 but Not MEG3 or MEG8 on Muscle Gene Expression in Callipyge Lambs. PLoS ONE 2009, 4, e7399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Bernard, C.; Cassar-Malek, I.; Le Cunff, M.; Dubroeucq, H.; Renand, G.; Hocquette, J.-F. New Indicators of Beef Sensory Quality Revealed by Expression of Specific Genes. J. Agric. Food Chem. 2007, 55, 5229–5237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Yang, D.; Zhang, B.; Wang, Z.; Zhang, L.; Chen, H.; Zhou, D.; Tang, K.; Wang, A.; Lin, P.; Jin, Y. COPS5 negatively regulates goat endometrial function via the ERN1 and mTOR-autophagy pathways during early pregnancy. J. Cell. Physiol. 2019, 234, 18666–18678. [Google Scholar] [CrossRef]
  78. Sun, W.; Hudson, N.J.; Reverter, A.; Waardenberg, A.J.; Tellam, R.L.; Vuocolo, T.; Byrne, K.; Dalrymple, B.P. An Always Correlated gene expression landscape for ovine skeletal muscle, lessons learnt from comparison with an “equivalent” bovine landscape. BMC Res. Notes 2012, 5, 632. [Google Scholar] [CrossRef] [Green Version]
  79. Ramayo-Caldas, Y.; Fortes, M.R.S.; Hudson, N.J.; Porto-Neto, L.R.; Bolormaa, S.; Barendse, W.; Kelly, M.; Moore, S.S.; Goddard, M.E.; Lehnert, S.A.; et al. A marker-derived gene network reveals the regulatory role of PPARGC1A, HNF4G, and FOXP3 in intramuscular fat deposition of beef cattle. J. Anim. Sci. 2014, 92, 2832–2845. [Google Scholar] [CrossRef]
  80. Naserkheil, M.; Bahrami, A.; Lee, D.; Mehrban, H. Integrating Single-Step GWAS and Bipartite Networks Reconstruction Provides Novel Insights into Yearling Weight and Carcass Traits in Hanwoo Beef Cattle. Animals 2020, 10, 1836. [Google Scholar] [CrossRef]
  81. Bai, M.; Sun, L.; Zhao, J.; Xiang, L.; Cheng, X.; Li, J.; Jia, C.; Jiang, H. Histological analysis and identification of spermatogenesis-related genes in 2-, 6-, and 12-month-old sheep testes. Sci. Nat. 2017, 104, 84. [Google Scholar] [CrossRef]
  82. Intergenomics Consortium; Frischknecht, M.; Bapst, B.; Seefried, F.R.; Signer-Hasler, H.; Garrick, D.; Stricker, C.; Fries, R.; Russ, I.; Sölkner, J.; et al. Genome-wide association studies of fertility and calving traits in Brown Swiss cattle using imputed whole-genome sequences. BMC Genom. 2017, 18, 910. [Google Scholar] [CrossRef] [PubMed]
  83. Wu, S.; Mipam, T.; Xu, C.; Zhao, W.; Shah, M.A.; Yi, C.; Luo, H.; Cai, X.; Zhong, J. Testis transcriptome profiling identified genes involved in spermatogenic arrest of cattleyak. PLoS ONE 2020, 15, e0229503. [Google Scholar] [CrossRef] [PubMed]
  84. Sharp, J.A.; Lefèvre, C.; Watt, A.; Nicholas, K.R. Analysis of human breast milk cells: Gene expression profiles during pregnancy, lactation, involution, and mastitic infection. Funct. Integr. Genom. 2016, 16, 297–321. [Google Scholar] [CrossRef] [PubMed]
  85. Mömke, S.; Sickinger, M.; Lichtner, P.; Doll, K.; Rehage, J.; Distl, O. Genome-wide association analysis identifies loci for left-sided displacement of the abomasum in German Holstein cattle. J. Dairy Sci. 2013, 96, 3959–3964. [Google Scholar] [CrossRef] [Green Version]
  86. Buitenhuis, B.; Janss, L.L.; Poulsen, N.A.; Larsen, L.B.; Larsen, M.K.; Sørensen, P. Genome-wide association and biological pathway analysis for milk-fat composition in Danish Holstein and Danish Jersey cattle. BMC Genom. 2014, 15, 1112. [Google Scholar] [CrossRef] [Green Version]
  87. Weedon, M.N.; Lango, H.; Lindgren, C.M.; Wallace, C.; Evans, D.M.; Mangino, M.; Freathy, R.M.; Perry, J.R.B.; Stevens, S.; Hall, A.S.; et al. Genome-wide association analysis identifies 20 loci that influence adult height. Nat. Genet. 2008, 40, 575–583. [Google Scholar] [CrossRef] [Green Version]
  88. Liu, J.Z.; Medland, S.E.; Wright, M.J.; Heath, A.C.; Madden, P.A.F.; Duncan, A.; Montgomery, G.W.; Martin, N.G.; McRae, A.F. Genome-wide association study of height and body mass index in Australian twin families. Twin Res. Hum. Genet. Off. J. Int. Soc. Twin Stud. 2010, 13, 179. [Google Scholar] [CrossRef]
  89. Nonneman, D.J.; Schneider, J.F.; Lents, C.A.; Wiedmann, R.T.; Vallet, J.L.; Rohrer, G.A. Genome-wide association and identification of candidate genes for age at puberty in swine. BMC Genet. 2016, 17, 50. [Google Scholar] [CrossRef] [Green Version]
  90. Vuocolo, T.; Byrne, K.; White, J.; McWilliam, S.; Reverter, A.; Cockett, N.E.; Tellam, R.L. Identification of a gene network contributing to hypertrophy in callipyge skeletal muscle. Physiol. Genom. 2007, 28, 253–272. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Runs of homozygosity islands in the Original Valachian sheep population (horizontal line denotes the top 1% of most homozygous regions, alternating colors distinguish markers on neighboring chromosomes).
Figure 1. Runs of homozygosity islands in the Original Valachian sheep population (horizontal line denotes the top 1% of most homozygous regions, alternating colors distinguish markers on neighboring chromosomes).
Animals 12 01346 g001
Figure 2. Runs of homozygosity islands in the Improved Valachian sheep population (horizontal line denotes the top 1% of most homozygous regions, alternating colors distinguish markers on neighboring chromosomes).
Figure 2. Runs of homozygosity islands in the Improved Valachian sheep population (horizontal line denotes the top 1% of most homozygous regions, alternating colors distinguish markers on neighboring chromosomes).
Animals 12 01346 g002
Figure 3. Linkage disequilibrium decay in the Original Valachian (red line) and Improved Valachian (blue line) sheep population.
Figure 3. Linkage disequilibrium decay in the Original Valachian (red line) and Improved Valachian (blue line) sheep population.
Animals 12 01346 g003
Figure 4. Standardized scores from the varLD indicate selection signatures between the Original and Improved Valachian sheep populations. Horizontal lines indicate highest results in percentage. Alternating colors distinguish markers on neighboring chromosomes.
Figure 4. Standardized scores from the varLD indicate selection signatures between the Original and Improved Valachian sheep populations. Horizontal lines indicate highest results in percentage. Alternating colors distinguish markers on neighboring chromosomes.
Animals 12 01346 g004
Table 1. Summary statistics of the genomic inbreeding coefficients (FROH) in the Original Valachian (OV) and Improved Valachian (IV) sheep.
Table 1. Summary statistics of the genomic inbreeding coefficients (FROH) in the Original Valachian (OV) and Improved Valachian (IV) sheep.
FROH (Up to 3 Generations)FROH (Up to 6 Generations)FROH (Up to 12
Generations)
FROH (Up to 25
Generations)
FROH (Up to 50
Generations)
OV0.020 ± 0.0260.038 ± 0.0400.058 ± 0.0520.068 ± 0.0580.079 ± 0.064
IV0.001 ± 0.0060.003 ± 0.0070.006 ± 0.0100.008 ± 0.0100.013 ± 0.011
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mészárosová, M.; Mészáros, G.; Moravčíková, N.; Pavlík, I.; Margetín, M.; Kasarda, R. Within- and between-Breed Selection Signatures in the Original and Improved Valachian Sheep. Animals 2022, 12, 1346. https://doi.org/10.3390/ani12111346

AMA Style

Mészárosová M, Mészáros G, Moravčíková N, Pavlík I, Margetín M, Kasarda R. Within- and between-Breed Selection Signatures in the Original and Improved Valachian Sheep. Animals. 2022; 12(11):1346. https://doi.org/10.3390/ani12111346

Chicago/Turabian Style

Mészárosová, Mária, Gábor Mészáros, Nina Moravčíková, Ivan Pavlík, Milan Margetín, and Radovan Kasarda. 2022. "Within- and between-Breed Selection Signatures in the Original and Improved Valachian Sheep" Animals 12, no. 11: 1346. https://doi.org/10.3390/ani12111346

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