Skip to main content

Integration of selective sweeps across the sheep genome: understanding the relationship between production and adaptation traits

Abstract

Background

Livestock populations are under constant selective pressure for higher productivity levels for different selective purposes. This pressure results in the selection of animals with unique adaptive and production traits. The study of genomic regions associated with these unique characteristics has the potential to improve biological knowledge regarding the adaptive process and how it is connected to production levels and resilience, which is the ability of an animal to adapt to stress or an imbalance in homeostasis. Sheep is a species that has been subjected to several natural and artificial selective pressures during its history, resulting in a highly specialized species for production and adaptation to challenging environments. Here, the data from multiple studies that aim at mapping selective sweeps across the sheep genome associated with production and adaptation traits were integrated to identify confirmed selective sweeps (CSS).

Results

In total, 37 studies were used to identify 518 CSS across the sheep genome, which were classified as production (147 prodCSS) and adaptation (219 adapCSS) CSS based on the frequency of each type of associated study. The genes within the CSS were associated with relevant biological processes for adaptation and production. For example, for adapCSS, the associated genes were related to the control of seasonality, circadian rhythm, and thermoregulation. On the other hand, genes associated with prodCSS were related to the control of feeding behaviour, reproduction, and cellular differentiation. In addition, genes harbouring both prodCSS and adapCSS showed an interesting association with lipid metabolism, suggesting a potential role of this process in the regulation of pleiotropic effects between these classes of traits.

Conclusions

The findings of this study contribute to a deeper understanding of the genetic link between productivity and adaptability in sheep breeds. This information may provide insights into the genetic mechanisms that underlie undesirable genetic correlations between these two groups of traits and pave the way for a better understanding of resilience as a positive ability to respond to environmental stressors, where the negative effects on production level are minimized.

Background

The sheep (Ovis aries) is a domestic species known for its diversity and adaptation potential, thriving in extreme agroecological conditions. It was one of the first species to be domesticated in the Middle East approximately 9000–11,000 years before present, originated from wild Asian mouflon species (Ovis orientalis) [1,2,3,4]. Initially, used for meat production, sheep were later selected for milk and wool production approximately 4000–5000 years ago, leading to the existence of over 1400 sheep breeds today. During domestication, genetic phenomena such as the “bottleneck” effect occurred, leading to a decrease in reproductive individuals and in their genetic variability [5]. However, at the same time, the process of splitting the population could have led to an increase in the total variability between populations. Artificial selection acts on the target population to create animals that meet human needs, resulting in different breeds of the same species. This process involves changes in the frequency of loci responsible for controlling phenotypes, impacting not only causal mutations but also nearby genetic markers due to linkage disequilibrium [6,7,8,9].

A signature of selection refers to a genome region with increased frequency of a specific allele in a population due to its functional importance during a selection process [10]. Natural and artificial selection drive the appearance of these signatures, which are characterized by reduced diversity not only directly in the affected mutations but also in nearby regions [11, 12]. Various statistical methods are available to identify signatures of selection, including comparing allelic frequencies between related breeds using FST or other associated statistics, analysing regions with low diversity or haplotypes, and identifying extreme allelic frequency patterns in a population [10, 13,14,15]. Genetic markers, particularly single nucleotide polymorphisms (SNPs), have been used in many studies to identify signatures of selection in domestic species such as cows and sheep [10, 11, 16, 17]. Following the identification of signatures, the next step typically involves evaluating genes within the identified regions to identify functional candidate genes that explain the effects of signatures of selection. Advancements in genomic technologies, such as commercial medium-density SNP chips and the reduced cost of next-generation sequencing (NGS), have facilitated studies on signatures of selection in sheep. Such studies have focused on specialized dairy or meat production breeds, the fat tail phenotype, and disease resistance [18,19,20,21,22]. More recently, studies have emphasized signatures of selection associated with the adaptation of sheep to specific environments such as high altitudes or heat stress [23, 24]. Another important characteristic of the livestock sector that has received increasing attention in recent years is animal resilience [25]. Resilience is the ability of animals to be minimally affected and/or rapidly respond to a disturbance of their health, welfare or productivity status [26]. Understanding the biological mechanisms that are associated with different adaptation responses might be useful in identifying candidate genomic regions and genes to be used for selecting more resilient animals.

Understanding the genetic basis of adaptation and preserving sheep genetic resources are crucial for genetic improvement, conservation, and sustainable livestock production [23, 27,28,29]. In spite of the advancements brought by recent studies on identifying signatures of selection in domestic animals, there are still inherent challenges. These include combining multiple types of analysis and statistical methods to eliminate false-positive results, difficulty in linking signatures of selection to a specific phenotype since the analysis is solely based on genotypic data, and the challenge of identifying the gene and causal mutation responsible for the signature of selection. To address these challenges, it is crucial to select the populations to be included in the study, combine various types of analyses within a single study, and compare results obtained from different studies conducted on different breeds. Studying the biological function of genes in a selection region helps identify genes linked to traits affected by selection. This allows inferring the potential phenotype associated with that selection signature across specific traits.

In light of the above, the objectives of this study were: (1) to perform a systematic review of studies reporting signatures of selection across the ovine genome associated with production and adaptation traits; (2) to combine the information from multiple studies to identify confirmed signatures of selection (CSS); (3) to integrate the information of positional candidate genes within CSS coordinates to identify potential functional profiles associated individually with productivity or adaptability; and (4) to compare functional candidate genes for productivity and adaptability to identify putative candidate genes underlying pleiotropic effects that may exist between these two classes of traits.

Methods

Selection of studies

PubMed (https://pubmed.ncbi.nlm.nih.gov/) was used as search engine, where the following searches were made: “selection signatures sheep”, “selection signatures sheep production”, “selection signatures sheep ecoregions”, and “selective sweep sheep adaptation”. In general, all the articles where the term “adaptation” was present in the title of the description of the population analyzed, were considered as related to adaptation, whereas those that analyzed populations bred for production traits (meat, milk, and wool) were considered as production. Fat tail-related studies, where the classification was unclear, were considered to be “adaptation” related studies.

Identification of confirmed signatures of selection

The remaining articles were examined to retrieve the genomic coordinates for the reported selective sweeps. The NCBI Remap (https://www.ncbi.nlm.nih.gov/genome/tools/remap) tool was used to convert all the coordinates to the version ARS-UI_Ramb_v2.0 of the ovine reference genome. For studies where the selective sweep location was reported as the coordinate of a single marker, a 250-kb interval downstream and upstream from this coordinate was calculated for the remapping step. Those studies that did not report the complete coordinates from the detected selective sweeps were also removed at this step. In addition, those studies where the selective sweeps were mapped based on an old version of the ovine reference genome for which conversion of coordinates to the ARS-UI_Ramb_v2.0 reference genome was not possible were also removed. A summary of the workflow applied to select the articles that composed the final sample is presented in Additional file 1 Figure S1.

The next step was the definition of the CSS that would be studied. For this purpose, the selective sweeps that were reported by three or more different studies and overlapped within the same genomic region were classified as CSS. The region comprising the CSS from all the overlapping studies was defined as the flanking region (smallest and largest genomic coordinate). In addition, for each CSS, the proportion of production and adaptation studies supporting the CSS was calculated. Subsequently, those CSS composed of more than 60% selective sweeps derived from production studies were called production CSS (prodCSS). In comparison, the CSS that comprised more than 70% of adaptation selective sweeps were called adaptation CSS (adapCSS).

Annotation of positional candidate genes and quantitative trait loci (QTL) mapped within the CSS coordinates

The positional candidate genes and quantitative trait loci (QTL) previously reported within the defined CSS were annotated using the GALLO package in R v.4.2.0 [30]. The gtf file used for gene annotation corresponding to the ARS-UI_Ramb_v2.0 of the ovine genome was obtained from NCBI. For the QTL annotation, the gff file from Sheep QTLdb corresponding to the ARS-UI_Ramb_v2.0 of the ovine genome was used. The GALLO package was also used to perform a QTL enrichment analysis for each trait annotated within the CSS flanking interval in the Sheep QTLdb using a genome-wide approach. The enriched QTL were defined based on a false-discovery rate (FDR) < 0.05 and a total number of QTL reported in the Sheep QTLdb larger than 1.

In addition, an enrichment analysis for gene ontology (GO) terms for the three categories available (biological process (BP), molecular function (MF), and cellular component (CC) was performed using the R package gprofiler2 [31]. In order to better understand the functional profile of the genes associated with prodCSS and adapCSS, the enrichment analyses for GO terms were performed individually for the genes annotated exclusively within prodCSS, for the genes annotated exclusively within adapCSS, and for the genes shared between both CSS classes. In addition, the R package rutils [32] was used to reduce the redundancy of GO terms through the go_reduce() function, where the Wang measure was selected to identify similar GO terms with a 0.7 threshold. The child GO terms assigned to the same GO parental terms were grouped into the same class, and the smallest p-value from the child terms was assigned to the parental term. The relationship between the positional candidate genes and the enriched GO terms related to production and adaptation was investigated using a network approach where the R packages igraph [33] and visNetwork [34] were used to identify hub genes of these networks, classified as those genes with the betweenness above the 90% quantile. The betweenness of a node in a network is defined as the number of shortest paths that pass through that node. Consequently, this metric can be used as a signal of the relevance of a gene in an interaction network. Similarly to the QTL enrichment analysis, the enriched GO terms and KEGG pathways were defined based on an FDR < 0.05 threshold. However, only GO terms with less than 1000 genes assigned to them in the gprofiler2 database were considered to avoid broad terms that might not be informative.

Results

Selected studies for the identification of confirmed signatures of selection

In total, 43 articles were retrieved from PubMed. However, five articles were excluded based on different criteria. Detailed information regarding all 43 articles and the exclusion criteria are shown in Additional file 2 Table S1. Among the 43 articles retained for the identification of CSS, 23 articles were classified as adaptation-related studies, and 15 articles were defined as production-related articles (Table 1). The coordinates of each selective sweep reported in the 38 selected studies are available in Additional file 3 Table S2 and were used for identifying CSS considering the Oar_rambouillet_v2.0 sheep reference genome.

Table 1 Studies retained for identification of confirmed selective sweeps associated with adaptation and production across the sheep genome

Confirmed signatures of selection for adaptation and production

The 529 CSS identified are available in Additional file 4 Table S3. Among these CSS, 213 adaptation CSS (adapCSS) and 172 production CSS (prodCSS) were selected based on the percentage of adaptation and production studies validating the CSS (see Additional file 4 Table S3), where a large number of CSS showed a ratio between 60 and 70% for both categories (see Additional file 5 Figure S2). In total, 4318 and 3092 genes were annotated within adapCSS and prodCSS, respectively (see Additional file 6 Table S4). In addition, 1851 genes were shared between adapCSS and prodCSS. However, it is not possible to disregard the potential functionality of these genes for both production- and adaptation-related traits.

Correspondence with QTL effects and functional analysis for positional candidate genes within confirmed signatures of selection

The annotation of QTL for adapCSS and prodCSS resulted in similar patterns. For both, the production QTL class was the most frequent, representing 89.75% and 72.15% of all the QTL annotated for prodCSS and adapCSS, respectively (Fig. 1a, b). In addition, slight increases in the percentages of QTL classes related to the reproduction, wool, health, and exterior classes were observed for adapCSS when compared to prodCSS. In total, 75 and 77 QTL were enriched for prodCSS and adapCSS, respectively (see Additional file 7 Table S5). The majority of these QTL (60 QTL) were shared between prodCSS and adapCSS (Fig. 1c). Among the QTL enriched exclusively for prodCSS (15 QTL), it is relevant to highlight the relatively large number of QTL related to wool (mean fibre diameter and fleece yield) and meat (longissimus muscle depth, longissimus muscle width, forequarter weight, loin yield, soft tissue depth at the GR site, carcass length, meat arachidonic acid content, leg yield and shear force). Regarding the QTL exclusively enriched for adapCSS, some that are related to adaptability, such as haematocrit, platelet count, ovulation rate, horn length, Trichostrongylus colubriformis FEC, and stillbirth, are worth mentioning. In addition, it is relevant to mention the number of milk fatty acid-related QTL observed as exclusively enriched for adapCSS (8 out 17 QTL): linoleic acid, lauric acid, conjugated linoleic acid, capric acid, cis-10 heptadecenoic acid, pentadecylic acid, palmitic acid, and palmitoleic acid. In spite of the identification of milk fatty-acids QTL as enriched exclusively on adapCSS, it is important to reinforce the importance of these QTL for production traits. Following this expectation, the enriched QTL terms shared between adapCSS and prodCSS comprise a combination of all QTL classes (Fig. 1c). However, it is important to mention the enrichment of QTL terms related to morphological traits associated with the selection process to which different sheep breeds were subjected, such as coat colour, horn circumference, horn type, teat number, jaw length, hind leg length, and ear size. In addition, several QTL related to fat deposition in different deposits were identified as enriched for both adapCSS and prodCSS, such as backfat at the third lumbar vertebra, tail fat deposition, internal fat amount, carcass fat percentage, total fat area, and fat weight in the carcass.

Fig. 1
figure 1

Results of the annotation of quantitative trait loci (QTL) for the confirmed selective sweeps (CSS). a Pie plot showing the percentage of each QTL trait type annotated within the coordinates of the CSS composed by more than 60% of production studies (prodCSS); b Pie plot showing the percentage of each QTL trait type annotated within the coordinates of the CSS composed by more than 60% of adaptation studies (adapCSS); and c Venn diagram describing the number of enriched QTL trait terms identified exclusively and shared for the prodCSS (in pink) and adapCSS (in green). The associated enriched QTL trait terms are highlighted in the text boxes

Regarding the enriched GO terms, 51 and 114 terms were identified exclusively for the genes that harboured only prodCSS or adapCSS, respectively (Fig. 2). In addition, 27 enriched GO terms were exclusively enriched for the list of genes harbouring both adapCSS and prodCSS. The complete list of enriched GO terms is available in Additional file 8 Table S6. Among the top 20 most enriched terms for the genes harbouring prodCSS, we highlight GO terms related to reproduction, response to temperature stimulus, response to chemokines and feeding behaviour (Fig. 2). For genes harbouring exclusively adapCSS, among the top 20 most enriched GO terms are terms related to cellular organization, signal transduction, response to light stimulus, organ maturation and growth (Fig. 2). In addition, a relevant number of enriched terms associated with lipid metabolism and adaptative thermogenesis were observed (see Additional file 9 Figure S3). The analysis of the network composed of these terms and the associated genes indicated the connection of these terms by functionally candidate genes, such as ELOVL3, SCD, IP6K1, FLCN, IL18, and FFAR4. Regarding the enrichment results for those genes harbouring both prodCSS and adapCSS, the metabolism of acetyl-CoA and monoacylglycerol, as well as the signalling of purinergic receptors, stand out as candidate processes for adaptation and production in sheep.

Fig. 2
figure 2

Number of enriched gene ontology (GO) terms identified for genes harboring exclusively CSS composed by more than 60% of production studies (prodCSS) in pink, genes harboring exclusively CSS composed by more than 60% of adaptation studies (adapCSS) in green, and harboring both prodCSS and adapCSS in gold. The top 10 enriched GO terms for each group are shown in the bubble plots, where the area of the bubbles corresponds to the number of associated genes and the color indicates the p-value scale (the darker colour, the smaller the false-discoveryrRate adjusted p-value). The richness factor shown in the x-axis is the ratio between the number of genes annotated in the current study associated with a specific GO term divided by the total number of genes associated with this specific term in the database

The betweenness of each gene in the network that was created based on the relationship between the genes harbouring CSS and the enriched GO terms was assessed for the three sets of enrichment results. The hub genes, defined as the genes with a betweenness above the 90% quantile, were selected for the enriched terms of genes harbouring only adapCSS (93 genes, quantile 90% threshold = 1084.64), genes harbouring only prodCSS (37 genes, quantile 90% threshold = 404.96), and genes harbouring both types of CSS (30 genes, quantile 90% threshold = 324.33). The betweenness values of each gene in the three networks are available in Additional file 10 Table S7. Genes harbouring CSS associated with enriched QTL terms were also considered associated with these QTL. Consequently, the relationships between genes and QTL were also represented as networks. The networks generated for the associations between the hub genes selected from the enriched GO term networks and the enriched QTL are shown in Figs. 3, 4 and 5.

Fig. 3
figure 3

Interaction network composed by quantitative trait loci (QTL) and genes mapped in adaptation selective sweeps. a Interaction network for the hub genes (in green) identified in the gene ontology network harboring exclusively confirmed selective sweeps composed by more than 60% of adaptation (adapCSS) studies and the QTL traits (in purples) annotated. The edges between a QTL and a gene indicate that this gene is associated with the respective enriched QTL trait term; and b networks showing the relationship between the hub genes selected for adapCSS (in pink) and the different QTL trait types. The thickness of the edges represents the number of traits annotated for each QTL trait type and associated with the respective gene

Fig. 4
figure 4

Interaction network composed by quantitative trait loci (QTL) and genes mapped in production selective sweeps. a Interaction network for the hub genes (in green) identified in the gene ontology network harboring exclusively confirmed selective sweeps composed by more than 60% of production (prodCSS) studies and the QTL traits (in purples) annotated. The edges between a QTL and a gene indicate that this gene is associated with the respective enriched QTL trait term; and b networks showing the relationship between the hub genes selected for prodCSS (in pink) and the different QTL trait types. The thickness of the edges represents the number of traits annotated for each QTL trait type and associated with the respective gene

Fig. 5
figure 5

Interaction network composed by quantitative trait loci (QTL) and genes for the hub genes identified in the gene ontology network harboring both production (prodCSS) and adaptation (adapCSS) confirmed selective sweeps. a Networks showing the relationship between the hub genes (purple) selected and the different QTL terms (green). The edges between a QTL and a gene indicate that this gene is associated with the respective enriched QTL; bf networks showing the relationship between the hub genes selected and the different QTL types. The thickness of the edges represents the number of traits annotated for each QTL type and associated with the respective gene. b Complete gene-QTL type network. c Network highlighting the direct connection between genes and health-related QTL. d Network highlighting the direct connection between genes and exterior-related QTL. e Network highlighting the direct connection between genes and reproduction-related QTL. f Network highlighting the direct connection between genes and production-related QTL

The qualitative analysis of the network composed of the hub genes harbouring exclusively adapCSS (Fig. 3a) and the enriched QTL terms suggested the presence of a group of genes related to multiple health-related QTL traits (Haemonchus contortus resistance, facial eczema susceptibility, Haemonchus contortus FEC, entropion, haemoglobin, red blood cell count, faecal egg count, and mean corpuscular haemoglobin concentration). The same block of genes is also associated with morphological and reproduction-related QTL phenotypes, such as horn type, teat number, reproductive seasonality, and staple strength and length. A similar analysis of the distribution of the betweenness of these genes in this network suggested ADIPOQ, PCNA, RHOA, TRIM32, TREM2, FFAR4, SORBS1, BRCA2, and KL as potential hub genes (quantile 90% threshold = 153.97). ADIPOQ, which is the gene with the highest betweenness in this network, is associated with QTL for different contents of fatty acids in the meat as well as carcass fat percentage and reproductive seasonality. Interestingly, when the relationship between the selected hub genes for adapCSS and enriched QTL was analysed based on the QTL types (Fig. 3b), 88 out of the 93 genes were directly associated with health-related QTL. In addition, two clusters of genes harbouring adapCSS were observed, one directly associated with meat and carcass and production-related QTL terms (see Additional file 11 Figure S4) and another cluster composed of genes directly linked to reproduction-related QTL (see Additional file 11 Figure S4). In the meat and carcass cluster, it is relevant to highlight the presence of important genes for the control of lipid metabolism, such as SREBF1, NCOR1, ALOX15, TRPV1, and TRPV2.

The network created with the genes harbouring exclusively prodCSS (Fig. 4a, b) suggested that BIN1, GHSR, and FSIP2 are potential hub genes for this network (quantile 90% threshold = 420.70). The BIN1 gene showed a association with multiple carcass and meat-related QTL. The GHSR gene was directly linked to reproduction, wool, health, milk and meat- and carcass-related QTL. The FSIP2 gene was linked to wool, health, milk and meat- and carcass-related QTL. The analysis of the networks composed of QTL term types indicated that almost all the selected genes are directly related to meat- and carcass-related QTL traits (see Additional file 12 Figure S5, excluding NOS1, HNF1A, ACACB, CACNA1C, and PTPN11) and milk-related QTL (see Additional file 12,Figure S5, excluding NTSR1, MUTYH, and EGFR). In addition, genes associated with wool-related QTL phenotypes (see Additional file 12 Figure S5) were also clustered close to health- and production-related QTL in the network analysis (Fig. 4b). In addition, exterior-related QTL were linked with a different cluster of genes compared to wool-, health- and production-related QTL (Fig. 4b and see Additional file 12 Figure S5). Reproduction-related QTL were associated with the MLH1, EGFR, SCN11A, and GHSR genes (Fig. 4b).

Finally, the network composed of hub genes from the enriched GO term networks harbouring both prodCSS and adapCSS and enriched QTL terms suggested that SLC7A5 and P2RX7 are hub genes of this network (quantile 90% threshold = 514.64). Important contributions to the network structure were also observed for the SNCA, GRID2, and PKD2 genes. In the network, a direct connection between these genes and multiple QTL types can be observed, such as health QTL (strongyle FEC, facial eczema susceptibility, faecal egg count, and somatic cell score), meat and carcass QTL (lean meat yield percentage, hot carcass weight, carcass fat percentage, dressing percentage, total fat area, fat weight in carcass, tail fat deposition, and muscle weight in carcass), and milk-related QTL (milk fat percentage, milk fat yield, and milk yield) (Fig. 5a). The QTL type network (Fig. 5b) for these genes indicated a separation between genes linked to health-related QTL (Fig. 5c) and exterior-related QTL (Fig. 5d). In addition, a strong connection between meat and carcass-related QTL terms and the GRID2, PKD2, and SNCA genes was observed (Fig. 5e). The hub genes SLC7A5 and P2RX7 were also connected with production-related QTL that were clustered close to health- and wool-related QTL (Fig. 5f).

The description of enriched GO terms and QTL terms associated with these hub genes is shown in Additional file 13 Table S8. It is important to highlight that in spite of the selection of these potential hub genes, all the other genes included in the abovementioned networks are potential functional candidate genes for production and/or adaptation-related traits.

Discussion

The combination of different sources of selection pressure results in the development of unique adaptative and production traits in livestock animals [6,7,8, 70,71,72]. The intense natural or artificial selection of favourable alleles for production and/or adaptation traits across the genome might reduce the genetic variability near those alleles due to the hitchhiking effect [73]. Selective sweeps are a specific type of genetic hitchhiking observed when directional selection is performed at a specific locus [74]. Here, the integration of studies that aim at identifying selective sweeps for adaptation and production traits across the sheep genome produces functional information about the specificities and similarities between these traits. It is relevant to highlight that among the 37 studies used to identify CSS, only two studies, Estrada-Reyes et al. [46] and McRae et al. [36], were also included in the SheepQTLdb [75]. Therefore, the QTL annotation and enrichment results obtained here can be interpreted as an independent validation of the association of these genomic regions with different production and adaptation traits.

Genomic regions exclusively linked to adaptation CSS

In contrast to almost all livestock animals, a marked seasonality of breeding is observed in sheep, which can be caused by several factors, such as temperature, nutritional status, social interactions, and neuroendocrinal factors [76, 77]. Seasonality can be interpreted as an evolutionary response to environmentally challenging periods. Indeed, environmental factors such as heat and day length are described as affecting milk production in sheep [78,79,80,81,82]. Photoperiodism is one of the most important biological processes associated with the synchronization of the mammalian energy balance with environmental conditions [83]. The response to the light stimulus was present among the most enriched GO terms identified for genes harbouring exclusively adapCSS. Among the genes associated with the control and regulation of the circadian clock, such as TP53 [84], GNAQ [85], DRD2 [86], USP2 [87], and PER1 [88] and photoreceptor function AIPL1 [89] and GNAT1 [90], stand out as relevant candidate genes for signatures of selection associated with adaptation to seasonality in sheep. The other two functionally enriched GO terms observed for the list of genes harbouring exclusively adapCSS were growth and animal organ maturation. The ability of an animal to grow and properly develop in its environment is a crucial characteristic of its adaptation. Three genes were shared between these two terms (EXT1, FGFR3, and RHOA). The EXT1 gene encodes a protein responsible for the elongation step of heparan sulfate biosynthesis, which is associated with the regulation of the development of the brain [91, 92] and bones [93, 94], as well as the gastrulation process [95]. FGFR3 is a receptor tyrosine kinase and acts negatively to regulate bone growth [96]. Finally, RHOA (a hub gene in the current study) is a member of the Rho family of small GTPases and belongs to the RhoA/ROCK pathway responsible for neuronal migration, dendrite development, and axonal extension [97]. In addition, RHOA was associated with chronic hypoxic foetal and adult sheep, suggesting an important role of RhoA pathways in hypertension control in newborns. In sheep, pulmonary hypertension in the newborn is a critical condition in breeds located at high altitudes [98]. The presence of CSS, including the RHOA gene, might suggest the presence of variants in this gene that contribute to resistance to pulmonary hypertension.

Among the enriched GO terms and QTL terms associated with the genes harbouring exclusively adapCSS, an association with lipid metabolism was observed. For example, the main hub gene identified in the networks composed of GO terms and QTL terms was the ADIPOQ gene. This gene encodes the hormone adiponectin, which is responsible for acting in the hypothalamus, stimulating food intake [99]. In addition, adiponectin acts in triglyceride hydrolysis, fatty acid decomposition, fatty acid oxidation and lipid synthesis [100, 101].

In livestock species, variants in the ADIPOQ gene are associated with marbling [102, 103], body measurements [104, 105], and growth and carcass traits [106]. It is crucial to recall that lipid metabolism plays an important role in sheep adaptation to extreme environments [107]. In addition, seasonality and adiposity in the body are associated; for instance, seasonal changes in body weight and fat percentage are observed frequently [108]. Other relevant candidate genes for lipid metabolism that are widely associated with meat and milk production traits in livestock species, such as SCD [109,110,111,112,113] and SREBF1 [114,115,116,117,118], were also identified among the genes harbouring exclusively adapCSS. Genes associated with lipid metabolism and adaptation are often associated with thermoregulation in mammals [119,120,121]. In total, 24 genes harbouring only adapCSS were associated with the GO term adaptive thermogenesis. The abovementioned genes ADIPOQ [122], SCD [123] and SREBF1 [124] are closely related to the control of thermogenesis and energy homeostasis. In addition, other genes involved in important processes associated with brown and white fat physiology were identified as exclusively harbouring adapCSS, such as ELOVL3 [125, 126], FLCN [127], TRPV1 [128], TRPV2 [129], OXT [130], IL18 [131], UCP2 [132, 133], and UCP3 [132, 134]. It is important to highlight the association of several genes harbouring adapCSS with parasite resistance-related enriched QTL (i.e., Salmonella abortusovis susceptibility, Haemonchus contortus FEC, facial eczema susceptibility, and Haemonchus contortus resistance). Interestingly, ADIPOQ, SREBF1, and FLCN were observed among those genes associated with these enriched QTL terms. The formation of lipid droplets which is one of the different functions performed by these genes [135,136,137], is a process that is reported to play an important role in the interaction between hosts and pathogens [105], which suggests a potential role of these genes (and others identified in the current study) in pathogen resistance in sheep. Among the other hub genes identified as harbouring exclusively prodCSS, the FFAR4 gene was previously associated with the regulation of glucose homeostasis, adiposity, gastrointestinal peptide secretion, and taste preference [138].

Another relevant result was obtained by evaluating the most-enriched QTL term observed for adapCSS, i.e. horn type. All the adapCSS associated with the QTL for horn type (and horn circumference) are mapped on chromosome 10. In addition, the majority of these adapCSS are mapped to a region comprising the coordinates of the RXFP2 gene, a gene involved in the development of sexual characteristics in humans and mice [139, 140]. A 1.8-kb insertion in the 3′-UTR of this gene was previously associated with polledness in sheep [141, 142]. In addition, RXFP2 is suggested to act in the development of unique horn phenotypes as a response to semiferalization (the partial animal reversion to life in nature, with human artificial selection no longer dominant) [143].

These results reinforce the association of the adapCSS identified in the current study with relevant biological processes that are directly associated with adaptation to environmental and physiological challenges in sheep breeds. In addition, the faster response and recovery to challenges that disturb animal health, welfare and production are the basis of selection for more resilient animals [20, 25].

Genomic regions harbouring production CSS exclusively

Interestingly, reproduction and feeding behaviour-related terms were among the most enriched terms associated with the genes harbouring exclusively prodCSS [112,113,114]. Important functional candidate genes for reproductive traits, such as ADCY10 [144], B4GALT1 [145], PGR [146], TBX3 [147], SPACA1 [148], SPATA16 [149], and SYCP2 [150], and feeding behaviours, such as CNTFR [151], DMBX1 [152], NTRK2 [153], PYY [154], and RMI1 [155], were observed among those associated with these enriched terms. In addition, a link was observed between reproduction and feeding behaviour by the presence of the genes CNR1 and GHSR (a hub gene in the current study) in both terms [127]. The expression of CNR1 in cortical neurons was shown to be increased after fasting, suggesting a role in feeding [156]. Regarding the association with reproductive traits, the action of CNR1 is associated with important sperm functions and testicular development [157]. It is important to highlight that CNR1 was also classified as a hub gene in the network analysis composed of the genes harbouring exclusively prodCSS and enriched GO terms. Similarly, the GHSR gene, which encodes the ghrelin receptor, acts in both processes. Ghrelin is a gastrointestinal peptide hormone involved in several biological processes, such as gut motility, gastric acid secretion, sleep and wake rhythm, taste sensation, glucose metabolism, and regulation of food intake [158,159,160]. Interestingly, genetic polymorphisms in GHSR were previously associated with body composition and growth in chickens and cattle [161,162,163]. In addition, ghrelin is suggested to be a regulator of the gonadotropic axis with a predominant negative effect through a signal of energy deficit [164]. In sheep, GHSR expression was detected in the adult testis and ovary with a significant effect of season (photoperiod) on its expression in the testis [165].

Heat stress resistance is an important, economically relevant trait in sheep due to its impact on several other traits, such as wool, milk production, immunity, and reproductive performance [166,167,168,169,170]. In total, 17 genes (ADRB2, ARPP21, CPB2, DNAJA1, DNAJA2, EIF2AK4, FBP1, HTR2A, MSTN, NF1, NOS1, NTSR1, POLR2D, SCN11A, SLC52A3, STAC, and TRPV4) identified in regions exclusively harbouring prodCSS were associated with the enriched GO term “response to temperature stimulus”. The search for genetic markers responsible for better resilience to heat stress is a key step for improving small ruminant productivity, welfare and health [171]. The genes BIN1, GHSR, and FSIP2 were above the quantile 90% for the betweenness distribution for all genes included in the network composed by genes and enriched QTL. The functions associated with the GHSR were previously discussed in this subsection. The BIN1 gene encodes the protein amphiphysin, an important regulator of cell muscle differentiation and maturation [172,173,174,175]. Consequently, polymorphisms in BIN1 could be associated with meat and carcass traits in sheep breeds. FSIP2 plays a crucial role in acrosome development and, consequently, in male fertility [176]. Although it was not defined as a hub gene by the criteria defined in the current study for the gene-QTL network, the SLIT2 gene showed an association pattern with milk-related QTL in the analysed networks. The action of SLIT2 in the mammary gland is associated with stem cell self-renewal and the generation of tubular bilayers during ductal morphogenesis [177, 178].

Genomic regions harbouring CSS for production and adaptation

Certain genetic loci can concurrently influence multiple complex traits, a phenomenon known as pleiotropy [179]. Pleiotropy may lead to the inadvertent selection of undesirable hitchhiking effects. The identification of genomic regions and/or variants associated with pleiotropic effects has the potential to improve the development of multivariate trait analysis and, subsequently, improve selection indices, resulting in greater genetic enhancement [180,181,182,183]. However, it is important to highlight that pleiotropy cannot be exclusively the cause of genetic correlation between traits, as linkage and gametic disequilibrium between loci can also strongly contribute to this phenomenon [155].

In the current study, several genes harbouring prodCSS and adapCSS were identified. The two most enriched GO terms for this list of genes were associated with purinergic nucleotide receptor activity. The purinergic receptors can be classified as ionotropic (P2X) and metabotropic (P2Y) and act over a multitude of biological processes; however, their actions over neurotransmitter release, synaptic plasticity, and cellular proliferation, differentiation, degeneration and regeneration stand out [184]. Among the genes associated with purinergic receptor activity identified here, both ionotropic (P2RX2, P2RX4, and P2RX7) and metabotropic (P2RY12, P2RY13, P2RY14, and P2RY4) genes were identified. The P2RX7 gene, which plays a crucial role in the modulation of inflammation and pain [185], was considered a hub gene in the network composed of genes and enriched QTL in the current study. In addition, genes encoding G-protein coupled receptors (GPR34 and GPR87) and a Ca2+-binding protein (NECAB2) were also associated with these enriched terms. The abovementioned genes are involved in, among other processes, the regulation of the immune system and the response to pain [186,187,188,189,190]. The activation of the immune response and sensitivity to stressful stimuli, such as pain, are at the interface between disease resistance and productivity. The improvement in immunological response and resistance is suggested to occur at the cost of productivity caused by the redirection of nutrient use [191]. In addition, genetic variations observed in this trade-off between higher immunity and productivity might be explained by variations in the sensitivity of the stimulus to trigger the immune system and the number of signals generated by its activation [191]. In addition, in cattle, P2RY12, P2RY14, and GPR87 were considered as functional candidate genes for 305-day milk yield [192], reinforcing the potential role of these genes in productivity. Indeed, P2RY12 was associated with all types of QTL, excluding wool-related QTL, in the network composed of hub genes and enriched QTL.

Terms related to acetyl-CoA were also identified as enriched for genes harbouring both prodCSS and adapCSS (ACSS1, MLYCD, MVD, NUDT7, PDHA1, PDHB, PDK3, and TDO2). In sheep, a decrease in fatty acid synthesis in the adipose tissue during lactation is observed due to a decrease in total acetyl-CoA carboxylase activity and the proportion of the enzyme in the active state [193]. The ACSS1 gene, which encodes an acetyl-CoA synthase, was identified among the genes associated with acetyl-CoA metabolism. This gene is associated with a response to metabolic stress [194] through acetate-mediated epigenetic regulation, which can also induce fatty acid synthesis [195]. Furthermore, the ACSS1 gene has been previously associated with lipid metabolism in relation to milk and meat composition in cattle and sheep [196,197,198,199,200]. Another gene associated with acetyl-CoA metabolism in the current study, MVD, was previously reported in selective sweeps for adaptation and productivity in different cattle breeds [201,202,203]. This gene encodes the enzyme mevalonate pyrophosphate decarboxylase, which is responsible for the conversion of mevalonate pyrophosphate into isopentenyl pyrophosphate during one of the first stages of cholesterol biosynthesis [204].

The coat colour QTL term was enriched only for adapCSS in the current study. However, adapCSS and prodCSS were associated with the coat colour QTL trait term during the annotation process. The majority of the CSS associated with these QTL are mapped in the region of chromosome 14 that carries the MC1R gene, which harbours both prodCSS and adapCSS. Mutations in MC1R are associated with the determination of coat colour in different sheep breeds [205, 206]. In addition, a region on chromosome 1, comprising the RUNX1 gene, was associated with coat colour QTL for both adapCSS and prodCSS. RUNX1 encodes a transcription factor associated with the development of hair and other skin appendages [207, 208]. Consequently, RUNX1 emerges as a candidate for coat colour in sheep breeds. In addition, it is important to mention that the region comprising the ASIP1 gene, another gene traditionally associated with coat colour in sheep [209, 210], was observed among the CSS identified with 50 and 60% of production and adaptation studies. These CSS were not included in the downstream functional analyses. However, the link suggested by our integrative analysis between coat colour and adapCSS highlights the known association between coat colour and adaptation to challenging environmental conditions, such as heat stress [211]. In addition, in some breeds, coat colour has an antagonistic effect on size and fitness [212]. Consequently, coat colour is a candidate phenotype to understand the relationship between productivity and adaptability in sheep breeds.

One other gene highlighted in the network between enriched GO terms and QTL trait terms derived from the genes associated with both prod-CSS and adaptCSS was SNCA. This gene encodes alpha-synuclein, a protein highly expressed in the presynaptic terminals of the central nervous system, associated with neurodegenerative disorders [213]. In sheep and goats, alpha-synuclein accumulates in their brains during scrapie infection, suggesting that perturbations in alpha-synuclein metabolism might play a role in prion infection [214]. Scrapie is a relevant health issue due to its neurodegenerative, progressive and lethal characteristics in sheep and goats, resulting in substantial efforts to reduce the disease incidence by selecting more resistant animals [215, 216]. The regulatory process of alpha-synuclein seems to play a crucial role in the brain inflammatory response through the modulation of lipid metabolism in the brain [217]. A link between SNCA and health-, meat- and carcass-related QTL trait terms was observed here. Genetic variants that map to the genomic regions harbouring the SNCA gene in pigs and cattle have been previously associated with backfat thickness [218] and milk somatic cell count [219], respectively. There is no direct evidence of the role of SNCA in production traits in livestock species. However, some studies suggest the action of alpha-synuclein as a glucoregulator in adipose tissue and skeletal muscle [220, 221], which might help explain a potential role in production-related traits in sheep. Another hypothesis is a potential hitchhiking effect observed between the SNCA locus and the NCAPG-LCORL locus. These two genomic regions are 2.31 Mb apart (based on the ARS-UI_Ramb_V2.0 reference genome). The NCAPG-LCORL locus is one of the most relevant loci associated with pleiotropic effects in livestock species, with associations reported for height, body weight, feed intake, gain, age at puberty, and meat and carcass traits [222,223,224,225,226,227].

Conclusions

Production and adaptation are intrinsically related processes in livestock species due to the intensive selective pressures for higher productivity levels to which these animals are subjected, driving the development of unique adaptive and production traits. Based on the identification of CSS associated with production and adaptation in sheep breeds, the present study pinpoints functional candidate genes for productivity and adaptability and candidate genes with the potential to simultaneously regulate both classes of traits. Intriguingly, a relevant role of lipid metabolism arose among the functional candidate genes that were identified in regions exclusively associated with adaptation or production. In addition, on the one hand, for adaptation-related regions, relevant functional candidate genes for the control of seasonality, circadian rhythm, and thermoregulation were observed. On the other hand, for production regions, genes associated with the control of feeding behaviour, reproduction, and cellular differentiation stand out as relevant functional candidates. However, it is important to highlight that the selection signals evaluated here are the result of directional selection and do not reflect signals subjected to balancing selection. In addition, selective sweeps on the X chromosome and interactions between mitogenome-autosomes were not investigated due to the absence of such results in the evaluated manuscripts. Consequently, not all sources and/or signals of selective sweeps were analysed here. The results obtained here help elucidate the genetic relationship between productivity and adaptability in sheep breeds. In addition, a series of functionally-relevant candidate genes are provided, which may help improve the fine-mapping of genomic regions for further studies aiming at the identification of causal variants associated with higher productivity and/or adaptability and resiliency in sheep.

Availability of data and materials

All the data used in the current study is available in the manuscript or supplementary information.

References

  1. Zeder MA. Core questions in domestication research. Proc Natl Acad Sci USA. 2015;112:3191–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Vigne JD. The origins of animal domestication and husbandry: a major change in the history of humanity and the biosphere. C R Biol. 2011;334:171–81.

    Article  PubMed  Google Scholar 

  3. Alberto FJ, Boyer F, Orozco-terWengel P, Streeter I, Servin B, de Villemereuil P, et al. Convergent genomic signatures of domestication in sheep and goats. Nat Commun. 2018;9:813.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Chessa B, Pereira F, Arnaud F, Amorim A, Goyache F, Mainland I, et al. Revealing the history of sheep domestication using retrovirus integrations. Science. 2009;324:532–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Bosse M, Megens HJ, Derks MFL, de Cara ÁMR, Groenen MAM. Deleterious alleles in the context of domestication, inbreeding, and selection. Evol Appl. 2019;12:6–17.

    Article  PubMed  Google Scholar 

  6. Mignon-Grasteau S, Boissy A, Bouix J, Faure JM, Fisher AD, Hinch GN, et al. Genetics of adaptation and domestication in livestock. Livest Prod Sci. 2005;93:3–14.

    Article  Google Scholar 

  7. Canario L, Mignon-Grasteau S, Dupont-Nivet M, Phocas F. Genetics of behavioural adaptation of livestock to farming conditions. Animal. 2013;7:357–77.

    Article  CAS  PubMed  Google Scholar 

  8. Bruford MW, Bradley DG, Luikart G. DNA markers reveal the complexity of livestock domestication. Nat Rev Genet. 2003;4:900–10.

    Article  CAS  PubMed  Google Scholar 

  9. Rauw WM, Johnson AK, Gomez-Raya L, Dekkers JCM. A hypothesis and review of the relationship between selection for improved production efficiency, coping behavior, and domestication. Front Genet. 2017;8:134.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Saravanan KA, Panigrahi M, Kumar H, Bhushan B, Dutt T, Mishra BP. Selection signatures in livestock genome: a review of concepts, approaches and applications. Livest Sci. 2020;241: 104257.

    Article  Google Scholar 

  11. de Simoni Gouveia JJ, da Silva MVGB, Paiva SR, de Oliveira SMP. Identification of selection signatures in livestock species. Genet Mol Biol. 2014;37:330–42.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Kreitman M. Methods to detect selection in populations with applications to the human. Annu Rev Genom Hum Genet. 2000;1:539–59.

    Article  CAS  Google Scholar 

  13. Hancock AM, Di Rienzo A. Detecting the genetic signature of natural selection in human populations: models, methods, and data. Annu Rev Anthropol. 2008;37:197–217.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Vatsiou AI, Bazin E, Gaggiotti OE. Detection of selective sweeps in structured populations: a comparison of recent methods. Mol Ecol. 2016;25:89–103.

    Article  CAS  PubMed  Google Scholar 

  15. Stephan W. Selective sweeps. Genetics. 2019;211:5–13.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Saravanan KA, Panigrahi M, Kumar H, Parida S, Bhushan B, Gaur GK, et al. Genomic scans for selection signatures revealed candidate genes for adaptation and production traits in a variety of cattle breeds. Genomics. 2021;113:955–63.

    Article  CAS  PubMed  Google Scholar 

  17. Fariello MI, Servin B, Tosser-Klopp G, Rupp R, Moreno C, Cristobal MS, et al. Selection signatures in worldwide sheep populations. PLoS ONE. 2014;9: e103813.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Prache S, Schreurs N, Guillier L. Review: Factors affecting sheep carcass and meat quality attributes. Animal. 2022;16: 100330.

    Article  PubMed  Google Scholar 

  19. Yaro M, Munyard KA, Morgan E, Allcock RJN, Stear MJ, Groth DM. Analysis of pooled genome sequences from Djallonke and Sahelian sheep of Ghana reveals co-localisation of regions of reduced heterozygosity with candidate genes for disease resistance and adaptation to a tropical environment. BMC Genom. 2019;20:816.

    Article  CAS  Google Scholar 

  20. Doeschl-Wilson A, Knap PW, Opriessnig T, More SJ. Review: Livestock disease resilience: from individual to herd level. Animal. 2021;15: 100286.

    Article  PubMed  Google Scholar 

  21. Moradi MH, Nejati-Javaremi A, Moradi-Shahrbabak M, Dodds KG, Brauning R, McEwan JC. Hitchhiking mapping of candidate regions associated with fat deposition in Iranian thin and fat tail sheep breeds suggests new insights into molecular aspects of fat tail selection. Animals. 2022;12:1423.

    Article  PubMed  PubMed Central  Google Scholar 

  22. Baazaoui I, Bedhiaf-Romdhani S, Mastrangelo S, Ciani E. Genome-wide analyses reveal population structure and identify candidate genes associated with tail fatness in local sheep from a semi-arid area. Animal. 2021;15: 100193.

    Article  CAS  PubMed  Google Scholar 

  23. Rojas-Downing MM, Nejadhashemi AP, Harrigan T, Woznicki SA. Climate change and livestock: impacts, adaptation, and mitigation. Clim Risk Manag. 2017;16:145–63.

    Article  Google Scholar 

  24. Lamy E, Van Harten S, Sales-Baptista E, Guerra MMM, De Almeida AM. Factors influencing livestock productivity. In: Environmental stress and amelioration in livestock production. Heidelberg: Springer; 2012. p. 19–51.

    Chapter  Google Scholar 

  25. Berghof TVL, Poppe M, Mulder HA. Opportunities to improve resilience in animal breeding programs. Front Genet. 2019;9:692.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Colditz IG, Hine BC. Resilience in farm animals: Biology, management, breeding and implications for animal welfare. Anim Prod Sci. 2016;56:1961–83.

    Article  Google Scholar 

  27. Boettcher PJ, Hoffmann I, Baumung R, Drucker AG, McManus C, Berg P, et al. Genetic resources and genomics for adaptation of livestock to climate change. Front Genet. 2014;5:461.

    PubMed  Google Scholar 

  28. Tedeschi LO, Muir JP, Riley DG, Fox DG. The role of ruminant animals in sustainable livestock intensification programs. Int J Sustain Dev World Ecol. 2015;22:452–65.

    Google Scholar 

  29. Bishop SC, Woolliams JA. Genetic approaches and technologies for improving the sustainability of livestock production. J Sci Food Agric. 2004;84:911–9.

    Article  CAS  Google Scholar 

  30. Fonseca PAS, Suárez-Vega A, Marras G, Cánovas Á. GALLO: an R package for genomic annotation and integration of multiple data sources in livestock for positional candidate loci. Gigascience. 2020;9: giaa149.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Kolberg L, Raudvere U, Kuzmin I, Vilo J, Peterson H. gprofiler2—an R package for gene list functional enrichment analysis and namespace conversion toolset g: Profiler. F1000Res. 2020;9:ELIXIR-709.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Bengtsson H. R. utils: various programming utilities. R package version. 2010. https://henrikbengtsson.github.io/R.utils/index.html. Accessed 16 Apr 2024.

  33. Csardi G, Nepusz T. The igraph software package for complex network research. InterJ Complex Syst. 2006;1695:1–9.

    Google Scholar 

  34. Almende BV, Thieurmel B. visNetwork: network visualization using “vis.js” library. CRAN. 2016. https://cran.r-project.org/web/packages/visNetwork/index.html. Accessed 23 Nov 2023.

  35. Álvarez I, Fernández I, Traoré A, Pérez-Pardal L, Menéndez-Arias NA, Goyache F. Genomic scan of selective sweeps in Djallonké (West African Dwarf) sheep shed light on adaptation to harsh environments. Sci Rep. 2020;10:2824.

    Article  PubMed  PubMed Central  Google Scholar 

  36. McRae KM, McEwan JC, Dodds KG, Gemmell NJ. Signatures of selection in sheep bred for resistance or susceptibility to gastrointestinal nematodes. BMC Genom. 2014;15:1–13.

    Article  Google Scholar 

  37. Moioli B, Pilla F, Ciani E. Signatures of selection identify loci associated with fat tail in sheep. J Anim Sci. 2015;93:4660–9.

    Article  CAS  PubMed  Google Scholar 

  38. Kim ES, Elbeltagy AR, Aboul-Naga AM, Rischkowsky B, Sayre B, Mwacharo JM, et al. Multiple genomic signatures of selection in goats and sheep indigenous to a hot arid environment. Heredity. 2016;116:255–64.

    Article  CAS  PubMed  Google Scholar 

  39. Liu Z, Ji Z, Wang G, Chao T, Hou L, Wang J. Genome-wide analysis reveals signatures of selection for important traits in domestic sheep from different ecoregions. BMC Genom. 2016;17:1–14.

    Article  Google Scholar 

  40. Mwacharo JM, Kim ES, Elbeltagy AR, Aboul-Naga AM, Rischkowsky BA, Rothschild MF. Genomic footprints of dryland stress adaptation in Egyptian fat-Tail sheep and their divergence from East African and western Asia cohorts. Sci Rep. 2017;7:17647.

    Article  PubMed  PubMed Central  Google Scholar 

  41. Manzari Z, Mehrabani-Yeganeh H, Nejati-Javaremi A, Moradi MH, Gholizadeh M. Detecting selection signatures in three Iranian sheep breeds. Anim Genet. 2019;50:298–302.

    Article  CAS  PubMed  Google Scholar 

  42. Wang W, Zhang X, Zhou X, Zhang Y, La Y, Zhang Y, et al. Deep genome resequencing reveals artificial and natural selection for visual deterioration, plateau adaptability and high prolificacy in Chinese domestic sheep. Front Genet. 2019;10:300.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Signer-Hasler H, Burren A, Ammann P, Drögemüller C, Flury C. Runs of homozygosity and signatures of selection: a comparison among eight local Swiss sheep breeds. Anim Genet. 2019;50:512–25.

    Article  CAS  PubMed  Google Scholar 

  44. Edea Z, Dadi H, Dessie T, Kim KS. Genomic signatures of high-altitude adaptation in Ethiopian sheep populations. Genes Genom. 2019;41:973–81.

    Article  Google Scholar 

  45. Yurchenko AA, Deniskova TE, Yudin NS, Dotsev AV, Khamiruev TN, Selionova MI, et al. High-density genotyping reveals signatures of selection related to acclimation and economically important traits in 15 local sheep breeds from Russia. BMC Genom. 2019;20:1–19.

    Article  Google Scholar 

  46. Estrada-Reyes ZM, Tsukahara Y, Amadeu RR, Goetsch AL, Gipson TA, Sahlu T, et al. Signatures of selection for resistance to Haemonchus contortus in sheep and goats. BMC Genom. 2019;20:1–14.

    Article  CAS  Google Scholar 

  47. Liu J, Yuan C, Guo T, Wang F, Zeng Y, Ding X, et al. Genetic signatures of high-altitude adaptation and geographic distribution in Tibetan sheep. Sci Rep. 2020;10:18332.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Mastrangelo S, Bahbahani H, Moioli B, Ahbara A, Al Abri M, Almathen F, et al. Novel and known signals of selection for fat deposition in domestic sheep breeds from Africa and Eurasia. PLoS ONE. 2019;14: e0209632.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Abied A, Bagadi A, Bordbar F, Pu Y, Augustino SMA, Xue X, et al. Genomic diversity, population structure, and signature of selection in five Chinese native sheep breeds adapted to extreme environments. Genes. 2020;11:494.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Abied A, Ahbara AM, Berihulay H, Xu L, Islam R, El-Hag FM, et al. Genome divergence and dynamics in the thin-tailed desert sheep from Sudan. Front Genet. 2021;12: 659507.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Guo T, Zhao H, Yuan C, Huang S, Zhou S, Lu Z, et al. Selective sweeps uncovering the genetic basis of horn and adaptability traits on fine-wool sheep in China. Front Genet. 2021;12: 604235.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Sweet-Jones J, Yurchenko AA, Igoshin AV, Yudin NS, Swain MT, Larkin DM. Resequencing and signatures of selection scan in two Siberian native sheep breeds point to candidate genetic variants for adaptation and economically important traits. Anim Genet. 2021;52:126–31.

    Article  CAS  PubMed  Google Scholar 

  53. Wiener P, Robert C, Ahbara A, Salavati M, Abebe A, Kebede A, et al. Whole-genome sequence data suggest environmental adaptation of Ethiopian sheep populations. Genome Biol Evol. 2021;13: evab014.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Zhang DY, Zhang XX, Di LF, Yuan LF, Li XL, Zhang YK, et al. Whole-genome resequencing reveals molecular imprints of anthropogenic and natural selection in wild and domesticated sheep. Zool Res. 2022;43:695.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Dzomba EF, Van Der Nest MA, Mthembu JNT, Soma P, Snyman MA, Chimonyo M, et al. Selection signature analysis and genome-wide divergence of South African Merino breeds from their founders. Front Genet. 2023;13: 932272.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Moioli B, Scatà MC, Steri R, Napolitano F, Catillo G. Signatures of selection identify loci associated with milk yield in sheep. BMC Genet. 2013;14:1–17.

    Article  Google Scholar 

  57. Gutiérrez-Gil B, Arranz JJ, Pong-Wong R, García-Gámez E, Kijas J, Wiener P. Application of selection mapping to identify genomic regions associated with dairy production in sheep. PLoS ONE. 2014;9: e94623.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Gutiérrez-Gil B, Esteban-Blanco C, Wiener P, Chitneedi PK, Suarez-Vega A, Arranz JJ. High-resolution analysis of selection sweeps identified between fine-wool Merino and coarse-wool Churra sheep breeds. Genet Sel Evol. 2017;49:81.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Purfield DC, McParland S, Wall E, Berry DP. The distribution of runs of homozygosity and selection signatures in six commercial meat sheep breeds. PLoS ONE. 2017;12: e0176780.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Seroussi E, Rosov A, Shirak A, Lam A, Gootwine E. Unveiling genomic regions that underlie differences between Afec-Assaf sheep and its parental Awassi breed. Genet Sel Evol. 2017;49:19.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Ruiz-Larrañaga O, Langa J, Rendo F, Manzano C, Iriondo M, Estonba A. Genomic selection signatures in sheep from the Western Pyrenees. Genet Sel Evol. 2018;50:9.

    Article  PubMed  PubMed Central  Google Scholar 

  62. Megdiche S, Mastrangelo S, Ben Hamouda M, Lenstra JA, Ciani E. A combined multi-cohort approach reveals novel and known genome-wide selection signatures for wool traits in merino and merino-derived sheep breeds. Front Genet. 2019;10: 468521.

    Article  Google Scholar 

  63. Shengwei H, Cunyuan L, Ming L, Xiaoyue L, Wei N, Yueren X, et al. Whole-genome resequencing reveals loci associated with thoracic vertebrae number in sheep. Front Genet. 2019;10:674.

    Article  Google Scholar 

  64. Yao Y, Pan Z, Di R, Liu Q, Hu W, Guo X, et al. Whole genome sequencing reveals the effects of recent artificial selection on litter size of bamei mutton sheep. Animals. 2021;11:157.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Tao L, He XY, Wang FY, Pan LX, Wang XY, Gan SQ, et al. Identification of genes associated with litter size combining genomic approaches in Luzhong mutton sheep. Anim Genet. 2021;52:545–9.

    Article  CAS  PubMed  Google Scholar 

  66. Liu Z, Bai C, Shi L, He Y, Hu M, Sun H, et al. Detection of selection signatures in South African Mutton Merino sheep using whole-genome sequencing data. Anim Genet. 2022;53:224–9.

    Article  CAS  PubMed  Google Scholar 

  67. Rekik E, Ahbara AM, Abate Z, Goshme S, Getachew T, Haile A, et al. Genomic analysis of 10 years of artificial selection in community-based breeding programs in two Ethiopian indigenous sheep breeds. Anim Genet. 2022;53:447–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. 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.

    Article  PubMed  PubMed Central  Google Scholar 

  69. Liu LL, Meng J, Ma HY, Cao H, Liu WJ. Candidate genes for litter size in Xinjiang sheep identified by specific locus amplified fragment (SLAF) sequencing. Anim Biotechnol. 2022;34:3053–62.

    Article  PubMed  Google Scholar 

  70. Fitzhugh HA, Bradford GE. Productivity of hair sheep and opportunities for improvement. In: Hair sheep of western Africa and the Americas. Boca Raton: CRC Press; 2019.

    Chapter  Google Scholar 

  71. Baker RL, Rege JEO. Genetic resistance to diseases and other stresses in improvement of ruminant livestock in the tropics. In: Proceedings of the 5th world congress on genetics applied to livestock production: 7–12 August 1994; Guelph. 1994.

  72. Baker RL, Gray GD. Appropriate breeds and breeding schemes for sheep and goats in the tropics. In: Better worm control for small ruminants in tropical Asia. Canberra: ACIAR Monograph; 2004.

  73. Smith JM, Haigh J. The hitch-hiking effect of a favourable gene. Genet Res. 2007;89:391–403.

    Article  PubMed  Google Scholar 

  74. Berry AJ, Ajioka JW, Kreitman M. Lack of polymorphism on the drosophila fourth chromosome resulting from selection. Genetics. 1991;129:1111–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Hu ZL, Park CA, Reecy JM. Building a livestock genetic and genomic information knowledgebase through integrative developments of animal QTLdb and CorrDB. Nucleic Acids Res. 2019;47:D701–10.

    Article  CAS  PubMed  Google Scholar 

  76. Gündoǧan M, Baki D, Yeni D. Reproductive seasonality in sheep. Acta Agric Scand A Anim Sci. 2003;53:175–9.

    Google Scholar 

  77. Chemineau P, Bodin L, Migaud M, Thiéry JC, Malpaux B. Neuroendocrine and genetic control of seasonal reproduction in sheep and goats. Reprod Domest Anim. 2010;45:42–9.

    Article  PubMed  Google Scholar 

  78. Gootwine E, Pollott GE. Factors affecting milk production in improved Awassi dairy ewes. Anim Sci. 2000;71:607–15.

    Article  Google Scholar 

  79. Pollott GE, Gootwine E. Reproductive performance and milk production of Assaf sheep in an intensive management system. J Dairy Sci. 2004;87:3690–703.

    Article  CAS  PubMed  Google Scholar 

  80. Tsartsianidou V, Kapsona VV, Sánchez-Molano E, Basdagianni Z, Carabaño MJ, Chatziplis D, et al. Understanding the seasonality of performance resilience to climate volatility in Mediterranean dairy sheep. Sci Rep. 2021;11:1889.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Sitzia M, Bonanno A, Todaro M, Cannas A, Atzori AS, Francesconi AHD, et al. Feeding and management techniques to favour summer sheep milk and cheese production in the Mediterranean environment. Small Rumin Res. 2015;126:43–58.

    Article  Google Scholar 

  82. Li S, Delger M, Dave A, Singh H, Ye A. Seasonal variations in the composition and physicochemical characteristics of sheep and goat milks. Foods. 2022;11:1737.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Dardente H, Wood S, Ebling F, Sáenz de Miera C. An integrative view of mammalian seasonal neuroendocrinology. J Neuroendocrinol. 2019;31: e12729.

    Article  PubMed  Google Scholar 

  84. Valafar B, Zaravinos A, Bonavida B. Cross talk between the circadian clock proteins and TP53 in cancer and therapeutic significance. Crit Rev Oncog. 2021;26:19–36.

    Article  PubMed  Google Scholar 

  85. Saini C, Petrenko V, Pulimeno P, Giovannoni L, Berney T, Hebrok M, et al. A functional circadian clock is required for proper insulin secretion by human pancreatic islet cells. Diabetes Obes Metab. 2016;18:355–65.

    Article  CAS  PubMed  Google Scholar 

  86. Wei H, Zapata RC, Lopez-Valencia M, Aslanoglou D, Farino ZJ, Benner V, et al. Dopamine D2 receptor signaling modulates pancreatic beta cell circadian rhythms. Psychoneuroendocrinology. 2020;113: 104551.

    Article  CAS  PubMed  Google Scholar 

  87. Yang Y, Duguay D, Bédard N, Rachalski A, Baquiran G, Na CH, et al. Regulation of behavioral circadian rhythms and clock protein PER1 by the deubiquitinating enzyme USP2. Biol Open. 2012;1:789–801.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Hamada T, Honma S, Honma KI. Light responsiveness of clock genes, Per1 and Per2, in the olfactory bulb of mice. Biochem Biophys Res Commun. 2011;409:727–31.

    Article  CAS  PubMed  Google Scholar 

  89. Kolandaivelu S, Ramamurthy V. AIPL1 protein and its indispensable role in cone photoreceptor function and survival. Adv Exp Med Biol. 2014;801:43–8.

    Article  PubMed  Google Scholar 

  90. Cameron MA, Lucas RJ. Influence of the rod photoresponse on light adaptation and circadian rhythmicity in the cone ERG. Mol Vis. 2009;15:2209–16.

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Yamaguchi Y, Inatani M, Matsumoto Y, Ogawa J, Irie F. Roles of heparan sulfate in mammalian brain development: current views based on the findings from ext1 conditional knockout studies. Prog Mol Biol Transl Sci. 2010;93:133–52.

    Article  CAS  PubMed  Google Scholar 

  92. Inatani M, Yamaguchi Y. Gene expression of EXT1 and EXT2 during mouse brain development. Dev Brain Res. 2003;141:129–36.

    Article  CAS  Google Scholar 

  93. Wuyts W, Van Hul W. Molecular basis of multiple exostoses: mutations in the EXT1 and EXT2 genes. Hum Mutat. 2000;15:220–7.

    Article  CAS  PubMed  Google Scholar 

  94. Hilton MJ, Gutiérrez L, Martinez DA, Wells DE. EXT1 regulates chondrocyte proliferation and differentiation during endochondral bone development. Bone. 2005;36:379–86.

    Article  CAS  PubMed  Google Scholar 

  95. Lin X, Wei G, Shi Z, Dryer L, Esko JD, Wells DE, et al. Disruption of gastrulation and heparan sulfate biosynthesis in EXT1-deficient mice. Dev Biol. 2000;224:299–311.

    Article  CAS  PubMed  Google Scholar 

  96. L’Hôte CGM, Knowles MA. Cell responses to FGFR3 signalling: growth, differentiation and apoptosis. Exp Cell Res. 2005;304:417–31.

    Article  PubMed  Google Scholar 

  97. Fujita Y, Yamashita T. Axon growth inhibition by RhoA/ROCK in the central nervous system. Front Neurosci. 2014;8:338.

    Article  PubMed  PubMed Central  Google Scholar 

  98. Herrera EA, Ebensperger G, Krause BJ, Riquelme RA, Reyes RV, Capetillo M, et al. Sildenafil reverses hypoxic pulmonary hypertension in highland and lowland newborn sheep. Pediatr Res. 2008;63:169–75.

    Article  CAS  PubMed  Google Scholar 

  99. Laursen TL, Zak RB, Shute RJ, Heesch MWS, Dinan NE, Bubak MP, et al. Leptin, adiponectin, and ghrelin responses to endurance exercise in different ambient conditions. Temperature. 2017;4:169–75.

    Article  Google Scholar 

  100. Tang N, Wang S, Qi J, Wu Y, Li Z, et al. Research progress on adiponectin regulating lipid metabolism. Acta Veterinaria et Zootechnica Sinica. 2018;49:2550–7.

    Google Scholar 

  101. Jiang H, Pu Y, Li ZH, Liu W, Deng Y, Liang R, et al. Adiponectin, may be a potential protective factor for obesity-related osteoarthritis. Diabetes Metab Syndr Obes. 2022;15:1305–19.

    Article  PubMed  PubMed Central  Google Scholar 

  102. Choi Y, Davis ME, Chung H. Effects of genetic variants in the promoter region of the bovine adiponectin (ADIPOQ) gene on marbling of Hanwoo beef cattle. Meat Sci. 2015;105:57–62.

    Article  CAS  PubMed  Google Scholar 

  103. Wang L, Xue K, Wang Y, Niu L, Li L, Zhong T, et al. Molecular and functional characterization of the adiponectin (AdipoQ) gene in goat skeletal muscle satellite cells. Asian-Australas J Anim Sci. 2018;31:1088–97.

    Article  PubMed  PubMed Central  Google Scholar 

  104. Al-Jumaili WS, Kadhim AH, Al-Thuwaini TM. Polymorphism of the ADIPOQ gene and its association with productive traits in Awassi Ewes. Mol Biol Rep. 2023;50:913–7.

    Article  CAS  PubMed  Google Scholar 

  105. Zhang L, Yang M, Li C, Xu Y, Sun J, Lei C, et al. Identification and genetic effect of a variable duplication in the promoter region of the cattle ADIPOQ gene. Anim Genet. 2014;45:171–9.

    Article  CAS  PubMed  Google Scholar 

  106. Shin S, Chung E. Novel SNPs in the bovine ADIPOQ and PPARGC1A genes are associated with carcass traits in Hanwoo (Korean cattle). Mol Biol Rep. 2013;40:4651–60.

    Article  CAS  PubMed  Google Scholar 

  107. Xu YX, Wang B, Jing JN, Ma R, Luo YH, Li X, et al. Whole-body adipose tissue multi-omic analyses in sheep reveal molecular mechanisms underlying local adaptation to extreme environments. Commun Biol. 2023;6:159.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Bartness TJ, Demas GE, Kay SC. Seasonal changes in adiposity: the roles of the photoperiod, melatonin and other hormones, and sympathetic nervous system. Exp Biol Med. 2002;227:363–76.

    Article  CAS  Google Scholar 

  109. Rincon G, Islas-Trejo A, Castillo AR, Bauman DE, German BJ, Medrano JF. Polymorphisms in genes in the SREBP1 signalling pathway and SCD are associated with milk fatty acid composition in Holstein cattle. J Dairy Res. 2012;79:66–75.

    Article  CAS  PubMed  Google Scholar 

  110. Matosinho CGR, de Souza Fonseca PA, Peixoto MGCD, Rosse IC, Lopes FCF, Zózimo T, et al. Phenotypic variation in milk fatty acid composition and its association with stearoyl-CoA desaturase 1 (SCD1) gene polymorphisms in Gir cows. J Anim Breed Genet. 2023;40:532–48.

    Article  Google Scholar 

  111. Miari S, Usai MG, Sechi T, Pernisa A, Carta A. One polymorphism at the stearoyl CoA desaturase (SCD) gene is associated to CLA content of sheep milk fat. Ital J Anim Sci. 2009;8:108–10.

    Article  Google Scholar 

  112. Aali M, Moradi-Shahrbabak H, Moradi-Shahrbabak M, Sadeghi M, Kohram H. Polymorphism in the SCD gene is associated with meat quality and fatty acid composition in Iranian fat- and thin-tailed sheep breeds. Livest Sci. 2016;188:81–90.

    Article  Google Scholar 

  113. Piórkowska K, Małopolska M, Ropka-Molik K, Nędza MS, Wiechniak A, Żukowski K, et al. Evaluation of SCD, ACACA and FASN mutations: effects on pork quality and other production traits in pigs selected based on RNA-seq results. Animals. 2020;10:123.

    Article  PubMed  PubMed Central  Google Scholar 

  114. Stachowiak M, Nowacka-Woszuk J, Szydlowski M, Switonski M. The ACACA and SREBF1 genes are promising markers for pig carcass and performance traits, but not for fatty acid content in the longissimus dorsi muscle and adipose tissue. Meat Sci. 2013;95:64–71.

    Article  CAS  PubMed  Google Scholar 

  115. Bhuiyan MSA, Yu SL, Jeon JT, Yoon D, Cho YM, Park EW, et al. DNA polymorphisms in SREBF1 and FASN genes affect fatty acid composition in Korean cattle (Hanwoo). Asian-Australas J Anim Sci. 2009;22:765–73.

    Article  CAS  Google Scholar 

  116. Liang C, Qiao L, Han Y, Liu J, Zhang J, Liu W. Regulatory roles of SREBF1 and SREBF2 in lipid metabolism and deposition in two Chinese representative fat-tailed sheep breeds. Animals. 2020;10:1317.

    Article  PubMed  PubMed Central  Google Scholar 

  117. Neofytou MC, Miltiadou D, Sfakianaki E, Constantinou C, Symeou S, Sparaggis D, et al. The use of ensiled olive cake in the diets of Friesian cows increases beneficial fatty acids in milk and Halloumi cheese and alters the expression of SREBF1 in adipose tissue. J Dairy Sci. 2020;103:8998–9011.

    Article  CAS  PubMed  Google Scholar 

  118. Carreño D, Hervás G, Toral PG, Castro-Carrera T, Frutos P. Fish oil-induced milk fat depression and associated downregulation of mammary lipogenic genes in dairy ewes. J Dairy Sci. 2016;99:7971–81.

    Article  PubMed  Google Scholar 

  119. Carmona MC, Hondares E, Rodriguez De La Concepcion ML, Rodríguez-Sureda V, Peinado-Onsurbe J, Poli V, et al. Defective thermoregulation, impaired lipid metabolism, but preserved adrenergic induction of gene expression in brown fat of mice lacking C/EBPβ. Biochem J. 2005;389:47–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Doi K, Ohno T, Kurahashi M, Kuroshima A. Thermoregulatory nonshivering thermogenesis in men, with special reference to lipid metabolism. Jpn J Physiol. 1979;29:359–72.

    Article  CAS  PubMed  Google Scholar 

  121. Florant GL. Lipid metabolism in hibernators: the importance of essential fatty acids. Am Zool. 1998;38:331–40.

    Article  CAS  Google Scholar 

  122. Qiao L, Yoo HS, Bosco C, Lee B, Feng GS, Schaack J, et al. Adiponectin reduces thermogenesis by inhibiting brown adipose tissue activation in mice. Diabetologia. 2014;57:1027–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Lee SH, Dobrzyn A, Dobrzyn P, Rahman SM, Miyazaki M, Ntambi JM. Lack of stearoyl-CoA desaturase 1 upregulates basal thermogenesis but causes hypothermia in a cold environment. J Lipid Res. 2004;45:1674–82.

    Article  CAS  PubMed  Google Scholar 

  124. Horie T, Nakao T, Miyasaka Y, Nishino T, Matsumura S, Nakazeki F, et al. microRNA-33 maintains adaptive thermogenesis via enhanced sympathetic nerve activity. Nat Commun. 2021;12:843.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Ji L, Gupta M, Feldman BJ. Vitamin D regulates fatty acid composition in subcutaneous adipose tissue through Elovl3. Endocrinology. 2016;157:91–7.

    Article  CAS  PubMed  Google Scholar 

  126. Jakobsson A, Jörgensen JA, Jacobsson A. Differential regulation of fatty acid elongation enzymes in brown adipocytes implies a unique role for Elovl3 during increased fatty acid oxidation. Am J Physiol Endocrinol Metab. 2005;289:E517–26.

    Article  CAS  PubMed  Google Scholar 

  127. Wada S, Neinast M, Jang C, Ibrahim YH, Lee G, Babu A, et al. The tumor suppressor FLCN mediates an alternate mTOR pathway to regulate browning of adipose tissue. Genes Dev. 2016;30:2551–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Mohammed M, Madden CJ, Andresen MC, Morrison SF. Activation of TRPV1 in nucleus tractus solitarius reduces brown adipose tissue thermogenesis, arterial pressure, and heart rate. Am J Physiol Regul Integr Comp Physiol. 2018;315:R134–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Sun W, Uchida K, Tominaga M. TRPV2 regulates BAT thermogenesis and differentiation. Channels. 2017;11:94–6.

    Article  PubMed  Google Scholar 

  130. Xi D, Long C, Lai M, Casella A, O’Lear L, Kublaoui B, et al. Ablation of oxytocin neurons causes a deficit in cold stress response. J Endocr Soc. 2017;1:1041–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Zhang X, Luo S, Wang M, Cao Q, Zhang Z, Huang Q, et al. Differential IL18 signaling via IL18 receptor and Na-Cl co-transporter discriminating thermogenesis and glucose metabolism regulation. Nat Commun. 2022;13:7582.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Brand MD, Esteves TC. Physiological functions of the mitochondrial uncoupling proteins UCP2 and UCP3. Cell Metab. 2005;2:85–93.

    Article  CAS  PubMed  Google Scholar 

  133. Caron A, Labbé SM, Carter S, Roy MC, Lecomte R, Ricquier D, et al. Loss of UCP2 impairs cold-induced non-shivering thermogenesis by promoting a shift toward glucose utilization in brown adipose tissue. Biochimie. 2017;134:118–26.

    Article  CAS  PubMed  Google Scholar 

  134. Hilse KE, Kalinovich AV, Rupprecht A, Smorodchenko A, Zeitz U, Staniek K, et al. The expression of UCP3 directly correlates to UCP1 abundance in brown adipose tissue. Biochim Biophys Acta Bioenergy. 2016;1857:72–8.

    Article  CAS  Google Scholar 

  135. Kessler SM, Laggai S, Van Wonterghem E, Gemperlein K, Müller R, Haybaeck J, et al. Transient hepatic overexpression of insulin-like growth factor 2 induces free cholesterol and lipid droplet formation. Front Physiol. 2016;7:147.

    PubMed  PubMed Central  Google Scholar 

  136. Yan M, Audet-Walsh É, Manteghi S, Dufour CR, Walker B, Baba M, et al. Chronic AMPK activation via loss of FLCN induces functional beige adipose tissue through PGC-1α/ERRα. Genes Dev. 2016;30:1034–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Zhang S, Cui Y, Gao X, Wei C, Wang Q, Yang B, et al. Resveratrol inhibits the formation and accumulation of lipid droplets through AdipoQ signal pathway and lipid metabolism lncRNAs. J Nutr Biochem. 2023;117: 109351.

    Article  CAS  PubMed  Google Scholar 

  138. Oh DY, Walenta E. Omega-3 fatty acids and FFAR4. Front Endocrinol. 2014;5:115.

    Article  Google Scholar 

  139. Feng S, Ferlin A, Truong A, Bathgate R, Wade JD, Corbett S, et al. INSL3/RXFP2 signaling in testicular descent: mice and men. Ann N Y Acad Sci. 2009;1160:197–204.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Yuan FP, Li X, Lin J, Schwabe C, Büllesbach EE, Rao CV, et al. The role of RXFP2 in mediating androgen-induced inguinoscrotal testis descent in LH receptor knockout mice. Reproduction. 2010;139:759–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Wiedemar N, Drögemüller C. A 1.8-kb insertion in the 3′-UTR of RXFP2 is associated with polledness in sheep. Anim Genet. 2015;46:457–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Wang X, Zhou G, Li Q, Zhao D, Chen Y. Discovery of SNPs in RXFP2 related to horn types in sheep. Small Rumin Res. 2014;116:133–6.

    Article  Google Scholar 

  143. Pan Z, Li S, Liu Q, Wang Z, Zhou Z, Di R, et al. Whole-genome sequences of 89 Chinese sheep suggest role of RXFP2 in the development of unique horn phenotype as response to semi-feralization. Gigascience. 2018;7: giy019.

    Article  PubMed  PubMed Central  Google Scholar 

  144. Akbari A, Pipitone GB, Anvar Z, Jaafarinia M, Ferrari M, Carrera P, et al. ADCY10 frameshift variant leading to severe recessive asthenozoospermia and segregating with absorptive hypercalciuria. Hum Reprod. 2019;34:1155–64.

    Article  CAS  PubMed  Google Scholar 

  145. Cipriano VTF, Miranda-Furtado CL, Mioranza A, Paro de Paz CC, Vila RA, Ramos ES, et al. B4GALT1, LHR, FSHR and IGF2 polymorphisms and scrotal circumference in Nellore bulls. Rev Bras Reprod Anim. 2018;42:70–5.

    Google Scholar 

  146. Akison LK, Robker RL. The critical roles of progesterone receptor (PGR) in ovulation, oocyte developmental competence and oviductal transport in mammalian reproduction. Reprod Domest Anim. 2012;47:288–96.

    Article  PubMed  Google Scholar 

  147. Douglas NC, Heng K, Sauer MV, Papaioannou VE. Dynamic expression of Tbx2 subfamily genes in development of the mouse reproductive system. Dev Dyn. 2012;241:365–75.

    Article  CAS  PubMed  Google Scholar 

  148. Chen P, Saiyin H, Shi R, Liu B, Han X, Gao Y, et al. Loss of SPACA1 function causes autosomal recessive globozoospermia by damaging the acrosome-acroplaxome complex. Hum Reprod. 2021;36:2587–96.

    Article  CAS  PubMed  Google Scholar 

  149. Dam AHDM, Koscinski I, Kremer JAM, Moutou C, Jaeger AS, Oudakker AR, et al. Homozygous mutation in SPATA16 is associated with male infertility in human globozoospermia. Am J Hum Genet. 2007;81:813–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Schilit SLP, Menon S, Friedrich C, Kammin T, Wilch E, Hanscom C, et al. SYCP2 translocation-mediated dysregulation and frameshift variants cause human male infertility. Am J Hum Genet. 2020;106:41–57.

    Article  CAS  PubMed  Google Scholar 

  151. Mercader JM, Saus E, Agüera Z, Bayés M, Boni C, Carreras A, et al. Association of NTRK3 and its interaction with NGF suggest an altered cross-regulation of the neurotrophin signaling pathway in eating disorders. Hum Mol Genet. 2008;17:1234–44.

    Article  CAS  PubMed  Google Scholar 

  152. Fujimoto W, Shiuchi T, Miki T, Minokoshi Y, Takahashi Y, Takeuchi A, et al. Dmbx1 is essential in agouti-related protein action. Proc Natl Acad Sci USA. 2007;104:15514–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Ribases M, Gratacos M, Badia A, Jimenez L, Solano R, Vallejo J, et al. Contribution of NTRK2 to the genetic susceptibility to anorexia nervosa, harm avoidance and minimum body mass index. Mol Psychiatry. 2005;10:851–60.

    Article  CAS  PubMed  Google Scholar 

  154. Ueno H, Yamaguchi H, Mizuta M, Nakazato M. The role of PYY in feeding regulation. Regul Pept. 2008;145:12–6.

    Article  CAS  PubMed  Google Scholar 

  155. Suwa A, Yoshino M, Yamazaki C, Naitou M, Fujikawa R, Matsumoto SI, et al. RMI1 deficiency in mice protects from diet and genetic-induced obesity. FEBS J. 2010;277:677–86.

    Article  CAS  PubMed  Google Scholar 

  156. Pérez CA, Stanley SA, Wysocki RW, Havranova J, Ahrens-Nicklas R, Onyimba F, et al. Molecular annotation of integrative feeding neural circuits. Cell Metab. 2011;13:222–32.

    Article  PubMed  PubMed Central  Google Scholar 

  157. Cacciola G, Chioccarelli T, Mackie K, Meccariello R, Ledent C, Fasano S, et al. Expression of type-1 cannabinoid receptor during rat postnatal testicular development: possible involvement in adult leydig cell differentiation. Biol Reprod. 2008;79:758–65.

    Article  CAS  PubMed  Google Scholar 

  158. Klok MD, Jakobsdottir S, Drent ML. The role of leptin and ghrelin in the regulation of food intake and body weight in humans: a review. Obes Rev. 2007;8:21–34.

    Article  CAS  PubMed  Google Scholar 

  159. De Vriese C, Delporte C. Ghrelin: a new peptide regulating growth hormone release and food intake. Int J Biochem Cell Biol. 2008;40:1420–4.

    Article  PubMed  Google Scholar 

  160. Xu G, Wang Z, Li Y, Li Z, Tang H, Zhao J, et al. Ghrelin contributes to derangements of glucose metabolism induced by rapamycin in mice. Diabetologia. 2012;55:1813–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Zhang B, Chen H, Guo Y, Zhang L, Zhao M, Lan X, et al. Associations of polymorphism within the GHSR gene with growth traits in Nanyang cattle. Mol Biol Rep. 2009;36:2259–63.

    Article  CAS  PubMed  Google Scholar 

  162. Fang M, Nie Q, Luo C, Zhang D, Zhang X. Associations of GHSR gene polymorphisms with chicken growth and carcass traits. Mol Biol Rep. 2010;37:423–8.

    Article  CAS  PubMed  Google Scholar 

  163. Darzi Niarami M, Masoudi AA, Vaez TR. Association of single nucleotide polymorphism of GHSR and TGFB2 genes with growth and body composition traits in sire and dam lines of a broiler chicken. Anim Biotechnol. 2014;25:13–22.

    Article  CAS  PubMed  Google Scholar 

  164. Tena-Sempere M. Ghrelin and reproduction: Ghrelin as novel regulator of the gonadotropic axis. Vitam Horm. 2008;77:285–300.

    Article  CAS  PubMed  Google Scholar 

  165. Miller DW, Harrison JL, Brown YA, Doyle U, Lindsay A, Adam CL, et al. Immunohistochemical evidence for an endocrine/paracrine role for ghrelin in the reproductive tissues of sheep apdat 20051031. Reprod Biol Endocrinol. 2005;3:60.

    Article  PubMed  PubMed Central  Google Scholar 

  166. Pérez RV, Cruz UM, Reyes LA, Correa-Calderón A, López Baca MDLA, Lara Rivera AL. Heat stress impacts in hair sheep production. Review. Rev Mex Cienc Pecu. 2020;11:205–22.

    Google Scholar 

  167. Marai IFM, El-Darawany AA, Fadiel A, Abdel-Hafez MAM. Physiological traits as affected by heat stress in sheep—a review. Small Rumin Res. 2007;71:1–12.

    Article  Google Scholar 

  168. Sevi A, Caroprese M. Impact of heat stress on milk production, immunity and udder health in sheep: a critical review. Small Rumin Res. 2012;107:1–7.

    Article  Google Scholar 

  169. van Wettere WHEJ, Kind KL, Gatford KL, Swinbourne AM, Leu ST, Hayman PT, et al. Review of the impact of heat stress on reproductive performance of sheep. J Anim Sci Biotechnol. 2021;12:26.

    Article  PubMed  PubMed Central  Google Scholar 

  170. Finocchiaro R, van Kaam JBCHM, Portolano B, Misztal I. Effect of heat stress on production of mediterranean dairy sheep. J Dairy Sci. 2005;88:1855–64.

    Article  CAS  PubMed  Google Scholar 

  171. Sejian V, Bagath M, Krishnan G, Rashamol VP, Pragna P, Devaraj C, et al. Genes for resilience to heat stress in small ruminants: a review. Small Rumin Res. 2019;173:42–53.

    Article  Google Scholar 

  172. Cowling BS, Prokic I, Tasfaout H, Rabai A, Humbert F, Rinaldi B, et al. Amphiphysin (BIN1) negatively regulates dynamin 2 for normal muscle maturation. J Clin Invest. 2017;127:4477–87.

    Article  PubMed  PubMed Central  Google Scholar 

  173. Lo HP, Lim YW, Xiong Z, Martel N, Ferguson C, Ariotti N, et al. Cavin4 interacts with bin1 to promote t-tubule formation and stability in developing skeletal muscle. J Cell Biol. 2021;220: e201905065.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Silva-Rojas R, Nattarayan V, Jaque-Fernandez F, Gomez-Oca R, Menuet A, Reiss D, et al. Mice with muscle-specific deletion of Bin1 recapitulate centronuclear myopathy and acute downregulation of dynamin 2 improves their phenotypes. Mol Ther. 2022;30:868–80.

    Article  CAS  PubMed  Google Scholar 

  175. Wechsler-Reya RJ, Elliott KJ, Prendergast GC. A role for the putative tumor suppressor Bin1 in muscle cell differentiation. Mol Cell Biol. 1998;18:566–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Fang X, Gamallat Y, Chen Z, Mai H, Zhou P, Sun C, et al. Hypomorphic and hypermorphic mouse models of Fsip2 indicate its dosage-dependent roles in sperm tail and acrosome formation. Development. 2021;148: dev199216.

    Article  CAS  PubMed  Google Scholar 

  177. Ballard MS, Zhu A, Iwai N, Stensrud M, Mapps A, Postiglione MP, et al. Mammary stem cell self-renewal is regulated by Slit2/Robo1 signaling through SNAI1 and mINSC. Cell Rep. 2015;13:290–301.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Strickland P, Shin GC, Plump A, Tessier-Lavigne M, Hinck L. Slit2 and netrin 1 act synergistically as adhesive cues to generate tubular bi-layers during ductal morphogenesis. Development. 2006;133:823–32.

    Article  CAS  PubMed  Google Scholar 

  179. Paaby AB, Rockman MV. The many faces of pleiotropy. Trends Genet. 2013;29:66–72.

    Article  CAS  PubMed  Google Scholar 

  180. Bolormaa S, Pryce JE, Reverter A, Zhang Y, Barendse W, Kemper K, et al. A multi-trait, meta-analysis for detecting pleiotropic polymorphisms for stature, fatness and reproduction in beef cattle. PLoS Genet. 2014;10: e1004198.

    Article  PubMed  PubMed Central  Google Scholar 

  181. Bolormaa S, Swan AA, Brown DJ, Hatcher S, Moghaddar N, Van Der Werf JH, et al. Multiple-trait QTL mapping and genomic prediction for wool traits in sheep. Genet Sel Evol. 2017;49:62.

    Article  PubMed  PubMed Central  Google Scholar 

  182. Scutari M, Howell P, Balding DJ, Mackay I. Multiple quantitative trait analysis using bayesian networks. Genetics. 2014;198:129–37.

    Article  PubMed  PubMed Central  Google Scholar 

  183. Xiang R, MacLeod IM, Daetwyler HD, de Jong G, O’Connor E, Schrooten C, et al. Genome-wide fine-mapping identifies pleiotropic and functional variants that predict many traits across global cattle populations. Nat Commun. 2021;12:860.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Krügel U. Purinergic receptors in psychiatric disorders. Neuropharmacology. 2016;104:212–25.

    Article  PubMed  Google Scholar 

  185. Hu SQ, Hu JL, Zou FL, Liu JP, Luo HL, Hu DX, et al. P2X7 receptor in inflammation and pain. Brain Res Bull. 2022;187:199–209.

    Article  PubMed  Google Scholar 

  186. Zhang M-D, Su J, Adori C, Cinquina V, Malenczyk K, Girach F, et al. Ca 2+-binding protein NECAB2 facilitates inflammatory pain hypersensitivity. J Clin Invest. 2018;128:3757–68.

    Article  PubMed  PubMed Central  Google Scholar 

  187. Cekic C, Linden J. Purinergic regulation of the immune system. Nat Rev Immunol. 2016;16:177–92.

    Article  CAS  PubMed  Google Scholar 

  188. Burnstock G. Purinergic receptors and pain. Curr Pharm Des. 2009;15:1717–35.

    Article  CAS  PubMed  Google Scholar 

  189. Gräler MH, Goetzl EJ. Lysophospholipids and their G protein-coupled receptors in inflammation and immunity. Biochim Biophys Acta Mol Cell Biol Lipids. 2002;1582:168–74.

    Article  Google Scholar 

  190. Pan HL, Wu ZZ, Zhou HY, Chen SR, Zhang HM, Li DP. Modulation of pain transmission by G-protein-coupled receptors. Pharmacol Ther. 2008;117:141–61.

    Article  CAS  PubMed  Google Scholar 

  191. Colditz IG. Effects of the immune system on metabolism: Implications for production and disease resistance in livestock. Livest Prod Sci. 2002;75:257–68.

    Article  Google Scholar 

  192. Paiva JT, Peixoto MGCD, Bruneli FAT, Alvarenga AB, Oliveira HR, Silva AA, et al. Genetic parameters, genome-wide association and gene networks for milk and reproductive traits in Guzerá cattle. Livest Sci. 2020;242: 104273.

    Article  Google Scholar 

  193. Vernon RG, Barber MC, Finley E. Modulation of the activity of acetyl-CoA carboxylase and other lipogenic enzymes by growth hormone, insulin and dexamethasone in sheep adipose tissue and relationship to adaptations to lactation. Biochem J. 1991;274:543–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Zhou W, Nie ZW, Zhou DJ, Cui XS. Acetyl-CoA synthases are essential for maintaining histone acetylation under metabolic stress during zygotic genome activation in pigs. J Cell Physiol. 2021;236:6948–62.

    Article  CAS  PubMed  Google Scholar 

  195. Gao X, Lin SH, Ren F, Li JT, Chen JJ, Yao CB, et al. Acetate functions as an epigenetic metabolite to promote lipid synthesis under hypoxia. Nat Commun. 2016;7:11960.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Berton MP, Fonseca LFS, Gimenez DFJ, Utembergue BL, Cesar ASM, Coutinho LL, et al. Gene expression profile of intramuscular muscle in Nellore cattle with extreme values of fatty acid. BMC Genom. 2016;17:972.

    Article  Google Scholar 

  197. Toral PG, Hervás G, Suárez-Vega A, Arranz JJ, Frutos P. Isolation of RNA from milk somatic cells as an alternative to biopsies of mammary tissue for nutrigenomic studies in dairy ewes. J Dairy Sci. 2016;99:8461–71.

    Article  CAS  PubMed  Google Scholar 

  198. Suárez-Vega A, Toral PG, Gutiérrez-Gil B, Hervás G, Arranz JJ, Frutos P. Elucidating fish oil-induced milk fat depression in dairy sheep: Milk somatic cell transcriptome analysis. Sci Rep. 2017;7:45905.

    Article  PubMed  PubMed Central  Google Scholar 

  199. Dai WT, Zou YX, White RR, Liu JX, Liu HY. Transcriptomic profiles of the bovine mammary gland during lactation and the dry period. Funct Integr Genom. 2018;18:125–40.

    Article  CAS  Google Scholar 

  200. Arora R, Kumar N, Sudarshan S, Fairoze MN, Kaur M, Sharma A, et al. Transcriptome profiling of longissimus thoracis muscles identifies highly connected differentially expressed genes in meat type sheep of India. PLoS ONE. 2019;14: e0217461.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Ben-Jemaa S, Mastrangelo S, Lee SH, Lee JH, Boussaha M. Genome-wide scan for selection signatures reveals novel insights into the adaptive capacity in local North African cattle. Sci Rep. 2020;10:19466.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Sorbolini S, Marras G, Gaspa G, Dimauro C, Cellesi M, Valentini A, et al. 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.

    Article  PubMed  PubMed Central  Google Scholar 

  203. Kooverjee BB, Soma P, Van Der Nest MA, Scholtz MM, Neser FWC. Selection signatures in South African Nguni and Bonsmara cattle populations reveal genes relating to environmental adaptation. Front Genet. 2022;13: 909012.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Krepkiy D, Miziorko HM. Identification of active site residues in mevalonate diphosphate decarboxylase: implications for a family of phosphotransferases. Protein Sci. 2004;13:1875–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Fontanesi L, Beretti F, Riggio V, Dall’Olio S, González EG, Finocchiaro R, et al. Missense and nonsense mutations in melanocortin 1 receptor (MC1R) gene of different goat breeds: association with red and black coat colour phenotypes but with unexpected evidences. BMC Genet. 2009;10:47.

    Article  PubMed  PubMed Central  Google Scholar 

  206. Yang GL, Fu DL, Lang X, Wang YT, Cheng SR, Fang SL, et al. Mutations in MC1R gene determine black coat color phenotype in Chinese sheep. Sci World J. 2013;2013: 675382.

    Article  Google Scholar 

  207. Osorio KM, Lee SE, McDermitt DJ, Waghmare SK, Zhang YV, Woo HN, et al. Runx1 modulates developmental, but not injury-driven, hair follicle stem cell activation. Development. 2008;135:1059–68.

    Article  CAS  PubMed  Google Scholar 

  208. Raveh E, Cohen S, Levanon D, Negreanu V, Groner Y, Gat U. Dynamic expression of Runx1 in skin affects hair structure. Mech Dev. 2006;123:842–50.

    Article  CAS  PubMed  Google Scholar 

  209. Norris BJ, Whan VA. A gene duplication affecting expression of the ovine ASIP gene is responsible for white and black sheep. Genome Res. 2008;18:1282–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Zhang X, Li W, Liu C, Peng X, Lin J, He S, et al. Alteration of sheep coat color pattern by disruption of ASIP gene via CRISPR Cas9. Sci Rep. 2017;7:8149.

    Article  PubMed  PubMed Central  Google Scholar 

  211. Berihulay H, Abied A, He X, Jiang L, Ma Y. Adaptation mechanisms of small ruminants to environmental heat stress. Animals. 2019;9:75.

    Article  PubMed  PubMed Central  Google Scholar 

  212. Gratten J, Wilson AJ, McRae AF, Beraldi D, Visscher PM, Pemberton JM, et al. A localized negative genetic correlation constrains microevolution of coat color in wild sheep. Science. 2008;319:318–20.

    Article  CAS  PubMed  Google Scholar 

  213. Goedert M. Alpha-synuclein and neurodegenerative diseases. Nat Rev Neurosci. 2001;2:492–501.

    Article  CAS  PubMed  Google Scholar 

  214. Adjou KT, Allix S, Ouidja MO, Backer S, Couquet C, Cornuejols MJ, et al. Alpha-synuclein accumulates in the brain of scrapie-affected sheep and goats. J Comp Pathol. 2007;137:78–81.

    Article  CAS  PubMed  Google Scholar 

  215. Baylis M, Goldmann W. The genetics of scrapie in sheep and goats. Curr Mol Med. 2005;4:385–96.

    Article  Google Scholar 

  216. Jeffrey M, González L, Espenes A, Press CM, Martin S, Chaplin M, et al. Transportation of prion protein across the intestinal mucosa of scrapie-susceptible and scrapie-resistant sheep. J Pathol. 2006;209:4–14.

    Article  CAS  PubMed  Google Scholar 

  217. Golovko MY, Barceló-Coblijn G, Castagnet PI, Austin S, Combs CK, Murphy EJ. The role of α-synuclein in brain lipid metabolism: a downstream impact on brain inflammatory response. Mol Cell Biochem. 2009;326:55–66.

    Article  CAS  PubMed  Google Scholar 

  218. Yin S, Song G, Gao N, Gao H, Zeng Q, Lu P, et al. Identifying genetic architecture of carcass and meat quality traits in a Ningxiang indigenous pig population. Genes. 2023;14:1308.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  219. Oliveira HR, Lourenco DAL, Masuda Y, Misztal I, Tsuruta S, Jamrozik J, et al. Single-step genome-wide association for longitudinal traits of Canadian Ayrshire, Holstein, and Jersey dairy cattle. J Dairy Sci. 2019;102:9995–10011.

    Article  CAS  PubMed  Google Scholar 

  220. Rodriguez-Araujo G, Nakagami H, Hayashi H, Mori M, Shiuchi T, Minokoshi Y, et al. Alpha-synuclein elicits glucose uptake and utilization in adipocytes through the Gab1/PI3K/Akt transduction pathway. Cell Mol Life Sci. 2013;70:1123–33.

    Article  CAS  PubMed  Google Scholar 

  221. Khoshi A, Goodarzi G, Mohammadi R, Arezumand R, Moghbeli M, Najariyan M. Reducing effect of insulin resistance on alpha-synuclein gene expression in skeletal muscle. Diabetol Metab Syndr. 2019;11:99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. La Y, Zhang X, Li F, Zhang D, Li C, Mo F, et al. Molecular characterization and expression of SPP1, LAP3 and LCORL and their association with growth traits in sheep. Genes. 2019;10:616.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  223. Takasuga A. PLAG1 and NCAPG-LCORL in livestock. Anim Sci J. 2016;87:159–67.

    Article  CAS  PubMed  Google Scholar 

  224. Lindholm-Perry AK, Sexten AK, Kuehn LA, Smith TPL, King DA, Shackelford SD, et al. Association, effects and validation of polymorphisms within the NCAPG-LCORL locus located on BTA6 with feed intake, gain, meat and carcass traits in beef cattle. BMC Genet. 2011;12:103.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Bouwman AC, Daetwyler HD, Chamberlain AJ, Ponce CH, Sargolzaei M, Schenkel FS, et al. Meta-analysis of genome-wide association studies for cattle stature identifies common genes that regulate body size in mammals. Nat Genet. 2018;50:362–7.

    Article  CAS  PubMed  Google Scholar 

  226. Saif R, Henkel J, Jagannathan V, Drögemüller C, Flury C, Leeb T. The LCORL locus is under selection in large-sized Pakistani goat breeds. Genes. 2020;11:168.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Widmann P, Reverter A, Fortes MRS, Weikard R, Suhre K, Hammon H, et al. A systems biology approach using metabolomic data reveals genes and pathways interacting to modulate divergent growth in cattle. BMC Genom. 2013;14:798.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work has been performed within the framework of the RESILSHEEP project funded by the Spanish Ministry of Science and Innovation, the State Research Agency and by the European Regional Development Fund (Ref PID2022-138676OB-100 funded by MCIN/AEI/10.13039/501100011033/FEDER, UE). PASF is the beneficiary of a Maria Zambrano Grant of the University of Leon funded by the Ministry of Universities (Madrid, Spain) and financed by the European Union-Next Generation EU programme.

Author information

Authors and Affiliations

Authors

Contributions

PASF carried out the formal data analysis, statistical analysis, and drafted the manuscript. ASV: participated in the development of the methodology, evaluation of the results, and in the review of the manuscript. JJA participated in the development of the methodology, evaluation of the results, and in the review of the manuscript. BGG participated in the conceptualization of the study, development of the methodology, validation of the results, and in the review of the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Beatriz Gutiérrez-Gil.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

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.

Supplementary Information

12711_2024_910_MOESM1_ESM.docx

Supplementary Material 1: Figure S1. Workflow for the selection of studies to be included in the pipeline for the identification of confirmed selective sweeps associated with production or adaptation traits.

12711_2024_910_MOESM2_ESM.xlsx

Supplementary Material 2: Table S1. Studies reporting selection signatures for production or adaptation traits across the sheep genome retrieved from Pubmed initially considered for this integrative review.

12711_2024_910_MOESM3_ESM.xlsx

Supplementary Material 3: Table S2. Genomic coordinates of all selective sweeps reported in the 37 studies retained for the identification of confirmed selective sweeps.

12711_2024_910_MOESM4_ESM.xlsx

Supplementary Material 4: Table S3. Confirmed selective sweeps (CSS) identified through the integration of the results obtained in the 37 individual studies evaluated in the present study.

12711_2024_910_MOESM5_ESM.docx

Supplementary Material 5: Figure S2. Distribution of the percentage of production (A) and adaptation (B) studies associated with each confirmed selective sweeps. The bar in red are above the threshold (60%) defined to classify a CSS as production or adaptation CSS.

Supplementary Material 6: Table S4. Genes annotated within the coordinates of the confirmed selective sweeps (CSS).

12711_2024_910_MOESM7_ESM.xlsx

Supplementary Material 7: Table S5. Results of quantitative trait loci (QTL) enrichment analysis performed for the traits annotated within the confirmed selective sweeps (CSS) coordinates for production and adaptation traits in the SheepQTLdb.

12711_2024_910_MOESM8_ESM.xlsx

Supplementary Material 8: Table S6. Results of Gene Ontology (GO) terms enrichment analysis for the genes harboring production and adaptation confirmed selective sweeps.

12711_2024_910_MOESM9_ESM.docx

Supplementary Material 9: Figure S3. Network composed by genes (in purple) and enriched Gene Ontology terms (in green) associated with lipid metabolism identified for the list of genes harboring exclusively confirmed selective sweeps composed by more than 60% of adaptation studies.

12711_2024_910_MOESM10_ESM.xlsx

Supplementary Material 10: Table S7. Betweenness values estimated for each gene present in the enriched gene ontology network network composed by genes harboring exclusively production and adaptation confirmed selective sweeps (CSS) and genes harboring both types of CSS.

12711_2024_910_MOESM11_ESM.docx

Supplementary Material 11: Figure S4. Interaction network composed by Quantitative trait loci (QTL) classes and genes (pink) for the hub genes identified in the gene ontology network harboring exclusively confirmed selective sweeps composed by more than 60% of adaptation (adapCSS) studies. The edges between a QTL and a gene indicate that this gene is associated with the respective QTL class. A) Network highlighting the direct connection between genes and health-related QTL trait terms. B) Network highlighting the direct connection between genes and meat and carcass-related QTL trait terms. C) Network highlighting the direct connection between genes and production-related QTL trait terms. D) Network highlighting the direct connection between genes and reproduction-related QTL trait terms.

12711_2024_910_MOESM12_ESM.docx

Supplementary Material 12: Figure S5. Interaction network composed by Quantitative trait loci (QTL) classes and genes (pink) for the hub genes identified in the gene ontology network harboring exclusively confirmed selective sweeps composed by more than 60% of production (prodCSS) studies. The edges between a QTL and a gene indicate that this gene is associated with the respective QTL class. A) Network highlighting the direct connection between genes and meat and carcass-related QTL. B) Network highlighting the direct connection between genes and milk-related QTL. C) Network highlighting the direct connection between genes and wool-related QTL. D) Network highlighting the direct connection between genes and production-related QTL.

12711_2024_910_MOESM13_ESM.xlsx

Supplementary Material 13: Table S8. Association between the selected hub genes for production and adaptation (or shared) confirmed selective sweeps with enriched gene ontology terms and quantitative trait loci annotated in the SheepQTLdb.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, 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 changes were made. 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/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fonseca, P.A.S., Suárez-Vega, A., Arranz, J.J. et al. Integration of selective sweeps across the sheep genome: understanding the relationship between production and adaptation traits. Genet Sel Evol 56, 40 (2024). https://doi.org/10.1186/s12711-024-00910-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12711-024-00910-w