Skip to main content

Single nucleotide polymorphism-based analysis of linkage disequilibrium and runs of homozygosity patterns of indigenous sheep in the southern Taklamakan desert

Abstract

Runs of Homozygosity (ROH) are homozygous genomic fragments inherited from parents to offspring. ROH can be used to indicate the level of inbreeding, as well as to identify possible signatures of artificial or natural selection. Indigenous sheep populations on the southern edge of the Taklimakan Desert have evolved unique genetic traits adapted to extreme desert environments. In an attempt to better understand the adaptive mechanisms of these populations under harsh conditions, we used Illumina® Ovine SNP50K BeadChip to perform a genomic characterization of three recognized breeds (Duolang: n = 36, Hetian: n = 84, Qira black: n = 189) and one ecotypic breed (Kunlun: n = 27) in the region. Additionally, we assessed genomic inbreeding coefficients through ROH analysis, revealing insights into the inbreeding history of these populations. Subsequently, we retrieved candidate genes associated with economic traits in sheep from ROH islands in each breed. To better understand the autozygosity and distribution of ROH islands in these indigenous sheep breeds relative to international breeds, we also included three commercial mutton breeds (Poll Dorset: n = 108, Suffolk: n = 163, Texel: n = 150). The study revealed that among seven sheep breeds, Hetian exhibited the shortest linkage disequilibrium (LD) decay distance, while Kunlun demonstrated the highest LD levels. A total of 10,916 ROHs were obtained. The number of ROHs per breed ranged from 34 (Kunlun) to 2,826 (Texel). The length of ROH was mainly 1–5 Mb (63.54%). Furthermore, 991 candidate genes specific to indigenous sheep breeds were identified, including those associated with heat tolerance, adaptability, energy metabolism, reproduction, and immune response. These findings elucidate the genetic adaptation of indigenous sheep in the Taklimakan Desert, uncovering distinctive characteristics of indigenous sheep formation, and advocating for the conservation and genetic enhancement of local sheep populations.

Peer Review reports

Introduction

The Taklamakan Desert, located in the central Tarim Basin in the northwest of China and surrounded by towering mountains [1, 2], is known for its arid conditions, high solar radiation levels, significant day-night temperature fluctuations, and frequent sandstorms. These challenging environmental conditions have driven natural selection in the local sheep population [3, 4] resulting in breeds such as the Duolang [5, 6], Hetian [7, 8], Qira black [9, 10], and Kunlun sheep [11]. Over time, these sheep have developed region-specific adaptive traits that enable them to maintain productivity and reproduce under extreme environmental stress. Despite the remarkable adaptations driven by natural selection, there is a need for further comprehensive research to uncover the genetic mechanisms responsible for phenotypic variations, including coat type, color, adaptability, and disease resistance, resulting from domestication in these populations.

ROH are genome-wide segments of homozygous genotypes resulting from the transmission of identical haplotypes by both parents to their offspring [12]. Throughout biological evolution and the formation of breeds, variations in selection pressure intensity, population size, and structure contribute to diversity in the length, number, distribution, and frequency of ROH in the genome [13]. Examining the unique distribution patterns of ROH within the genomic composition of animal populations proves valuable for deducing the extent and age of inbreeding and unraveling the evolutionary history of livestock [14]. Longer ROH segments typically signify more recent mating events among close relatives, whereas shorter ROH segments indicate inbreeding events occurring in more ancient generations [15, 16]. Dzomba et al. [17] utilized the Illumina® Ovine SNP50K Beadchip to analyze ROH in 14 South African and 17 global sheep breeds. They found that long ROH were more prevalent in breeds with smaller population sizes, accompanied by elevated genomic inbreeding coefficients. Furthermore, the prevalence of short ROH reflected the long-term commercial breeding history of these breeds [18]. Sallam et al. [19] utilized 60 K chip data to analyze ROH in four Egyptian sheep breeds. They discovered that the Damascus breed had the shortest ROH, whereas the Egyptian Nubian breed not only had the longest ROH but also exhibited the highest inbreeding coefficient (FROH= 0.12), as previously reported by Eydivandi et al. under intensive breeding practices [20].

Moreover, selection imparts discernible imprints in particular segments of the animal genome, resulting in an elevated purity of these genomic regions in comparison to others. These selectively favored genomic regions frequently manifest as ROH islands [13, 21, 22], which exhibit a non-random distribution across the genome and are universally shared among individuals within the same species [23]. Consequently, ROH islands are commonly used to evaluate the genetic adaptation of populations in order to identify genes related to livestock adaptability and economic traits. Liu et al. [24] identified candidate genes associated with tail types, reproductive performance, environmental adaptability, and immunity within the high-frequency ROH regions of five distinct Chinese sheep breeds. Mastrangelo et al. [25] discovered candidate genes related to environmental adaptability, immune response, and lipid metabolism based on the high-frequency ROH regions of the Valle del Belice sheep breed. He et al. [26] searched the high-frequency ROH regions of 635 Chinese Merino sheep and retrieved Quantitative Trait Locus (QTLs) associated with horn phenotype, immune traits, and environmental adaptability.

Linkage disequilibrium (LD) can improve the accuracy of genomic association analysis and predict marker regions [27]. LD decay patterns also help in understanding the selection patterns experienced by regions of gene function [28]. Effective population size (Ne), a key population genomics parameter, reflects ideal population size and is crucial for assessing genetic diversity and adaptation potential [29, 30]. Comparing LD and Ne of different species can reveal their inbreeding rate and environmental adaptability.

The Taklamakan region is celebrated for its distinctive geographical and climatic characteristics, presenting unique challenges to the survival and reproduction of native sheep. Our research aims to employ the Illumina® Ovine SNP50K BeadChip to conduct a genomic characterization of the indigenous sheep breeds in the region, revealing their distribution pattern of ROH, genome-level diversity, inbreeding coefficients, and population history. Furthermore, we investigate adaptive mechanisms to extreme environments in these sheep based on identified high-frequency ROH islands in the genome. Additionally, this study can serve as a reference for genetic resource conservation, rational breeding planning, and controlling inbreeding in different sheep populations.

Materials and methods

This study adheres rigorously to the relevant guidelines and regulations of the Ministry of Agriculture of the People’s Republic of China. The animal study underwent thorough review and received approval from the Ethics Committee of the College of Animal Science and Technology of Tarim University (protocol code: 2023039). Prior to their participation in this study, written informed consent was obtained from the owners of the animals.

Animal samples

The experiment involved the selection of 336 healthy adult indigenous ewes of similar age, sourced from the southern edge of the Taklamakan Desert in Xinjiang, China. These ewes represented four distinct populations: Duolang sheep (DUO), Qira Black sheep (QIR), Hetian sheep (HET), and Kunlun sheep (KUN). DUO from the Duolang sheep conservation farm in Maigaiti County, Xinjiang; QIR from the Cele Black sheep conservation farm in Cele County, Xinjiang; HET from Hetian region, Xinjiang; and KUN from the Kunlun sheep conservation farm in Qiemo County, Xinjiang (Fig. 1; Table 1). Additionally, the analysis incorporated 421 genotypes from three sheep breeds of international origin, including Poll Dorset (POL), Suffolk (SUF), and Texel (TEX). Access to the dataset was granted under the permission of the International Sheep Genomics Consortium (ISGC, http://www.sheephapmap.org). To minimize any potential confounding effects stemming from sexual disparities, all 757 samples were deliberately chosen from female specimens.

Fig. 1
figure 1

Geographic distribution of the four indigenous sheep populations from the southern edge of the Taklamakan Desert in Xinjiang, China

Table 1 A brief introduction to experimental animals (Breed name, place of origin, sample size, purpose, reproduction rate, different climatic and geographical conditions, and feeding methods)

Genotyping and quality control

Blood samples were obtained from the jugular vein of sheep, and Deoxyribonucleic acid (DNA) extraction was carried out using a DNA extraction kit (Tiangen Biotech Co., Ltd., Beijing, China). All samples were genotyped using the Illumina® Ovine SNP50 BeadChip (Beijing Compass Biotechnology Co., Ltd, Beijing, China), which includes 46 900 single nucleotide polymorphisms (SNPs).

Quality control for SNPs was executed using PLINK v1.90 (access dates: Jul-18-2023) [31], with exclusion criteria applied to SNP sites having a call rate < 0.95, and Hardy-Weinberg equilibrium (HWE) p-value < 10− 6. Additionally, individuals with more than 5% missing genotypes excluded from the analysis. After quality control, 757 samples with 46,896 SNPs were retained for continued analysis.

Estimation of linkage disequilibrium (LD) and effective population size (Ne)

The LD degree between markers was quantified using the r2 proposed by Hill and Robertson, where r2 represents the square of the correlation coefficient between the alleles of two loci [32, 33]. The values of r2 range from 0 to 1, and the distances between markers were categorized into 10 classes: 0 < r2 ≤ 20 kb; 20 < r2 ≤ 40 kb; 40 < r2 ≤ 60 kb; 60 < r2 ≤ 100 kb; 100 < r2 ≤ 200 kb; 200 < r2 ≤ 500 kb; 500 < r2 ≤ 1000 kb; 1000 < r2 ≤ 2000 kb; 2000 < r2 ≤ 4000 kb; 4000 < r2 ≤ 6000 kb. The calculation formula is as follows:

$$ \:{r^2} = \frac{{{{\left[ {{f_{\left( {AB} \right)}}\cdot{f_{\left( {ab} \right)}} - {f_{\left( {Ab} \right)}}\cdot{f_{\left( {aB} \right)}}} \right]}^2}}}{{{f_{\left( A \right)}}\cdot{f_{\left( a \right)}}\cdot{f_{\left( B \right)}}\cdot{f_{\left( b \right)}}}} $$

Where f(AB), f(ab), f(Ab), and f(aB) represent the frequencies of the four haplotypes formed between genotypes, and f(A), f(a), f(B), and f(b) are the allele frequencies. We utilized PopLDdecay V3.42 (access dates: Jul-25-2023) to compute the r2 of alleles to determine the levels of LD and decay [34]. The obtained results were then visualized using Plot_OnePop.pl (access dates: Jul-25-2023).

We utilized SMC + + v1.15.3 (access dates: Aug-11-2023), which incorporates an innovative spline regularization scheme [35, 36], to infer the population size history and divergence time of the four sheep populations from the southern edge of the Taklamakan Desert in Xinjiang, China. This method effectively minimizing estimation errors. Initially, employing the vcf2smc tool, each population within the VCF dataset was converted into independent SMC + + format output files. Next, the model fitting was executed using the estimate function, with all simulations conducted under the initial condition of a mutation rate of 1.25e-9 [37].

Detection of runs of homozygosity

The distribution of ROH in each sample within seven populations was assessed using PLINK v1.90 (access dates: Aug-31-2023). ROH estimation adhered to the following criteria: (1) Each sliding window contained 50 SNPs; (2) a maximum of one heterozygous genotype per window; (3) a maximum of one missing genotype per window; (4) The scanning window threshold was set at 0.05; (5) a maximum gap between two consecutive SNPs in ROH of 500 kb; (6) The minimum length of ROH was set to 1000 kb to exclude short ROH resulting from LD; (7) To minimize the likelihood of detecting ROH by chance, the minimum number of SNPs constituting a ROH was determined using the method proposed by Lentz et al. [38]. That is:

$$ \:{I_{ROH}} = \frac{{{{\log }_e}\left( {\frac{\alpha }{{{n_S} \times {n_i}}}} \right)}}{{{{\log }_e}\left( {1 - \overline {{N_{het}}} } \right)}} $$

The variable IROH denotes the minimum number of SNPs constituting ROH. α represents the false positive rate of identified ROH (set to 0.05 in this study), nS represents the number of SNPs per individual, ni is the total number of individuals in the population, and \(\:\stackrel{\text{-}}{\text{N}\text{het}\text{}}\) signifies the average heterozygosity across all SNPs.

To assess the distribution of ROH lengths among the seven breeds, ROH lengths were classified into six categories: 1–5 Mb, 5–10 Mb, 10–15 Mb, 15–20 Mb, 20–25 Mb, and > 25 Mb. We calculated the ROH frequency for each length category (Within each breed, count the number of ROH within each length category and calculate the proportion within the total ROHs) and the average ROH length for each category (For each breed, sum up the lengths of all individual ROHs belonging to different categories, and then divide the resulting sums by the number of individuals in that breed). The enumeration of ROH was conducted independently for each chromosome within the seven distinct sheep breeds. Furthermore, the cumulative assessment included determining the total length and aggregate count of ROH for individual sheep specimens.

Estimation of inbreeding coefficient

Utilizing the equation proposed by Sams et al. [15], the ROH-based inbreeding coefficient (FROH) was calculated for each breed as follows:

$$\:{F_{ROH}} = \frac{{\sum \: {L_{ROH}}}}{{{L_{AUTO}}}}$$

We calculated the cumulative length of identified ROH within individuals, denoted as ∑LROH. LAUTO signifies the total length of the autosomal sheep genome, which was 2,452 Mb in our study.

Additionally, for validating the accuracy of FROH, we computed the inbreeding coefficient based on the expected number of homozygous genotypes (FHOM) using PLINK v1.90 with the following parameters: -het (access dates: Sep-17-2023) [31].

$$\:{F_{HOM}} = \frac{{{H_o} - {H_e}}}{{N - {H_e}}}$$

Where Ho represents the observed number of heterozygous genotypes in the population, He represents the theoretical frequency of heterozygous genotypes under Hardy-Weinberg equilibrium, and N represents the total number of SNPs in the individual.

Selection signatures

To identify genomic regions that are most frequently associated with ROH across sheep populations under different environmental conditions, we calculated the percentage of each SNP within ROH by tallying their occurrences. Subsequently, we visualized the distribution of all SNPs along chromosomal positions.

In this study, to identify genomic regions frequently associated with ROH in each breed, we used an SNP frequency threshold of > 45% to screen genomic regions associated with ROH across various breeds [39]. Within these regions, ROH islands were defined as clusters consecutive SNP loci > 10 Mb in length, containing a minimum of 200 SNPs [17]. We annotated the genomic regions associated with these ROH islands using the Oar_v4.0 sheep reference genome. Subsequently, we conducted Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis for genes annotated to ROH islands in different populations using the online tool DAVID (https://david.ncifcrf.gov/) [40]. Additionally, we referred to NCBI (https://www.ncbi.nlm.nih.gov/pmc/) and OMIM (http://www.ncbi.nlm.nih.gov/omim) databases for functional information on genes within the ROH islands.

Results

Estimation of LD and Ne

We observed a gradual decline in r2 values with increasing physical distance between SNPs, accompanied by a slowdown in the rate of decay (Fig. 2a). By comparing LD among seven sheep populations, we observed differences in LD levels and decay rates with the same SNP intervals. Overall, HET exhibited the shortest LD decay distance, while KUN demonstrated the highest LD levels (r2 = 0.081). Additionally, it was evident that among the three commercial meat sheep breeds, SUF had the fastest LD decay rate, followed by TEX and POL.

Fig. 2
figure 2

(a) LD decay map measured by r2 over distance between SNPs in seven sheep populations; (b) Estimated Ne for 10,000 ancestral generations of four Chinese indigenous sheep breeds

Figure 2b illustrates the changing trends of Ne in different sheep populations as the number of generations changes. Between the 1,000th and 2,000th generations, the Ne of all four sheep breeds had reached their maximum, with estimated values of approximately 4,718 (DUO), 1,293 (KUN), 2,765 (QIR), and 2,810(HET). In the most recent generations, the Ne of the four sheep breeds remained relatively stable. QIR and DUO exhibited the largest effective populations, with estimated Ne values of 1,082 (QIR) and 901 (DUO), respectively. In contrast, KUN and HET had the smallest effective populations, with corresponding Ne estimates of 704 (KUN) and 746 (HET), respectively.

Runs of homozygosity patterns

Based on the defined parameters, a total of 10,916 ROH segments were detected. The number of ROHs per breed ranged from 34 (KUN) to 2,826 (TEX). Among the indigenous sheep populations, QIR had the highest total number of ROH (2,255), followed by DUO (346) and HET (294), while KUN had the lowest total number of ROH (34) (Table S1). Importantly, readers should interpret the results presented here with caution, as varying sample sizes may influence the detection and interpretation of ROHs. Therefore, the identified genomic regions and their implications should be considered in light of these differences.

Analyzing ROH lengths in seven sheep breeds revealed predominant clustering within 1–5 Mb (63.54%). Commercial meat breeds showed higher average lengths and proportions of ROHs within this range than local breeds. Among the local breeds with ROHs exceeding 10 Mb in length, QIR had the longest average ROH length and the highest frequency, whereas KUN exhibited the shortest average ROH length and the lowest frequency (Fig. 3, Table S2, Table S3).

Fig. 3
figure 3

The average sum of ROH of each breed in different ROH length classes

Figure 4 illustrates the distribution of ROHs per chromosome for each breed. ROHs did not display a uniform distribution among chromosomes within each breed. Across all breeds, the highest percentage of ROHs was observed on Ovis aries chromosomes (OAR) 2(13.25%) and 3(13.14%). Notably, ROHs were not detected on OAR 17, 21, and 23 in the indigenous sheep (Table S1).

Fig. 4
figure 4

Number of ROH found on each chromosome in seven sheep populations

Figure 5 depicts the relationship between the total number of ROHs per individual and the total genomic length covered by ROHs. There were significant differences among different breeds, with KUN and HET exhibiting a high number of short ROHs. Additionally, extreme individuals with ROH coverage exceeding 400 Mb on autosomes were observed in QIR (Table S4).

Fig. 5
figure 5

The total genomic length covered by ROH per individual (x axis) and total number of ROH per individual (y axis)

Inbreeding coefficient

We computed the inbreeding coefficient (FROH) based on ROH (Table 2). Figure 6A shows the variation and distribution of the average inbreeding coefficient estimates among different sheep populations. Among four indigenous populations from the southern edge of the Taklamakan Desert, QIR had the highest inbreeding coefficient (0.0393 ± 0.0331), while KUN had the lowest (0.0059 ± 0.0083). Moreover, inbreeding coefficient (FHOM) based on the expected E(HOM) and observed O(HOM) number of homozygous genotypes was calculated using Plink V. 1.90 [31] (Table 2). To visualize inbreeding degrees, we plotted FROH and FHOH as line graphs (Fig. 6B). A significant correlation was observed between FROH and FHOM (r = 0.785, p = 0.036).

Fig. 6
figure 6

(A) Distribution of the runs of homozygosity inbreeding coefficient (FROH) within each breed; (B) Trend plot of FROH and FHOM for each sheep breed

Table 2 Mean and standard deviation of ROH and genomic inbreeding coefficients in seven sheep breeds

Gene annotation of common ROH islands

The statistical analysis of ROH islands in seven sheep breeds are presented in Table S4 and Table S5. To compare the differences in ROH islands between the indigenous breeds from the southern edge of the Taklamakan Desert and commercial meat breeds, we categorized the seven sheep breeds into two distinct groups: Chinese indigenous sheep and commercial meat sheep. The results revealed a significant number of ROH islands on OAR 1, 2, 3, 6, 20, and 25 in four indigenous breeds (Fig. 7A); similarly, a significant number of ROH islands were observed on OAR 2, 4, 5, 6, 7, and 10 in three commercial meat breeds (Fig. 7B). A total of 542 ROH islands were identified in the Chinese indigenous sheep group, while 559 ROH islands were found in the commercial meat sheep group.

Fig. 7
figure 7

(a) Manhattan plot of SNPs in ROH islands on different chromosomes of four indigenous sheep breeds; (b) Manhattan plot of SNPs in ROH islands on different chromosomes of three commercial meat sheep breeds; (c) Wayne diagram of candidate genes of four indigenous sheep breeds; (d) Wayne diagram of candidate genes of three commercial meat sheep breeds; (e) Wayne plots of candidate genes in four indigenous sheep breeds versus three commercial meat sheep breeds

The genomic region associated with each of these ROH islands was annotated using the sheep genome Oar_v4.0. In the four indigenous breeds, a total of 10,387 candidate genes were retrieved (DUO: 5,129, HET: 3,621, KUN: 433, QIR: 9,976) (Fig. 7C). In the three commercial meat breeds, a total of 9,979 candidate genes were identified (SUF: 8,593, TEX: 6,411, POL: 6,455) (Fig. 7D). To better understand the adaptive mechanisms of indigenous sheep from the southern Taklamakan Desert under extreme conditions, we excluded intersecting genes with commercial meat breeds, leaving us with 991 genes considered as candidate genes for adaptation to extreme environments (Fig. 7E; Table S6). After functional enrichment analysis of these candidate genes, a total of 44 GO terms and KEGG pathways were enriched at the threshold (p < 0.05). The most significant GO terms were found in coenzyme A biosynthetic process (GO:0015937), regulation of ventricular cardiac muscle cell membrane repolarization (GO:0060307), negative regulation of epithelial cell proliferation involved in prostate gland (GO:0060770), voltage-gated potassium channel complex (GO:0008076), and olfactory receptor activity (GO:0004984). The most enriched KEGG pathways included Olfactory transduction (oas04740), Insulin resistance (oas04931), Adipocytokine signaling pathway (oas04920), and PI3K-Akt signaling pathway (oas04151) (Table S7).

Discussion

This study utilized the Illumina® Ovine SNP50 Beadchip to detect ROH in seven sheep populations, quantified their number, length, distribution, and frequency. Using this ROH information, we inferred the population histories of different sheep breeds, calculated inbreeding coefficients, and explored adaptive mechanisms in indigenous sheep under extreme production conditions by analyzing candidate genes identified in ROH islands. Additionally, we conducted LD analysis on four indigenous sheep breeds and three commercial meat breeds, and estimated Ne for the indigenous breeds.

Linkage disequilibrium and effective population size

Among all sheep breeds, KUN exhibited the highest value of average , which corresponded to the smallest Ne. The smallest values of average were observed in QIR, which had the highest Ne. Notably, the LD decay lines of QIR and HET almost overlap. Our finding is consistent with previous research on HET, which indicated that higher levels correspond to smaller Ne [41].

In this study, the minimal Ne observed in KUN can be primarily attributed to their geographic isolation and the environmental stress of high altitudes. These factors likely contributed to reduced genetic flow and, consequently, a smaller Ne [42, 43]. In contrast, QIR boasted the highest Ne, owing to its status as a prominent wool-producing breed in China. This high Ne level indicates successful conservation efforts for this valuable genetic resource [44].

Furthermore, the study reveals that the Ne of four sheep populations declined over the past approximately 1300 generations. The period coincided with the peak of the Silk Road trade [45], a time of significant cultural and economic exchange in the region. As a crucial junction along the Silk Road, Xinjiang experienced significant genetic admixture, leading to a reduction in the genetic resources of indigenous sheep populations during that era [46, 47]. This historical context provides a plausible explanation for the observed decline in Ne.

ROHs and ROH-based inbreeding coefficient (F ROH)

The length of ROH segments reflects the history of animal inbreeding [14]. As the number of generations increases, the likelihood of ROH segments being disrupted by recombination increases. Consequently, shorter ROH segments primarily result from inbreeding in more distant generations, while longer ROH segments indicate more recent inbreeding [15, 16].

Among the indigenous sheep breeds studied, QIR exhibited the highest proportion of long ROH segments and the highest levels of FROH and FHOH. As an elite sheep breed valued for its high-quality lambskin, QIR is predominantly distributed in Cele County, Hotan Prefecture, Xinjiang [48]. However, recent changes in market demand and the massive introduction of foreign commercial meat sheep breeds have contributed to a significant decline in the number of purebred QIR [49]. Notably, all QIR samples in this study were collected from the Cele Black Sheep Conservation Farm, where the limited population size has led to a high degree of inbreeding. To address this issue, we recommend implementing a regular strategy for exchanging rams within the population to enhance genetic diversity and expand the effective population size of QIR [41].

KUN possesses the smallest proportion of long ROH segments and the lowest levels of inbreeding coefficient. Distributed in higher-altitude mountainous areas and adopting a semi-grazing and semi-confined feeding method, KUN has a relatively small core population size. However, existing breeding plans have successfully avoided excessive inbreeding [41]. This suggests that adaptive breeding strategies can mitigate the negative effects of small population size on genetic diversity.

HET, widely distributed across the Hotan region of Xinjiang, originated from a small ancestral population and has evolved into diverse ecotypes through inbreeding in more distant generations, natural selection, and artificial breeding [41, 50]. The HET samples in this study, originating from various conservation farms, exhibited a higher proportion of short ROH segments due to geographical isolation. This reinforces the significant role of geographical factors in shaping population genetic structures [51, 52].

Selection signatures of candidate genes within ROH islands

Significant genetic differences exist between domestic and foreign sheep in economically important traits, making it a crucial research area in genetic breeding [53, 54, 55]. To deeply explore the adaptive mechanisms of indigenous sheep from the southern edge of the Taklimakan Desert under harsh environmental conditions, this study separately identified ROH islands in indigenous sheep and commercial mutton sheep that may be related to breed-specific traits. After gene annotation, by eliminating intersecting genes, we successfully identified multiple candidate genes associated with the indigenous sheep breeds in extreme environments.

Candidate genes related to olfactory receptor activity and olfactory transduction

The well-developed sense of smell in sheep is a key evolutionary advantage for their survival in the natural environment, affecting critical behaviors like foraging [56], danger detection [57], and reproduction [58]. Olfactory receptors in vertebrates, located in the olfactory epithelium, sense odor molecules in the environment [59]. These receptors are mostly G protein-coupled receptors (GPCRs). When odor molecules bind to these receptors, they activate a biochemical cascade that converts external odors into electrochemical signals, involving G protein activation [60, 61], adenosine triphosphate (ATP) conversion to cyclic adenosine monophosphate (cAMP) [62], ion channel opening, and intracellular Ca2+ increase [63], ultimately activating olfactory neurons [64]. These neurons then transmit the sensory information to the brain, forming the animal’s perception of olfaction [65].

We identified GPCRs, including CELSR2 [66], RXFP3 [67], and VIPR1 [68], which are associated with neuronal development and synapse formation in mammals, potentially influencing the complexity and sensitivity of the olfactory system. Additionally, GRM7 [69] and ADGRL [70] play roles in neurotransmitter release and neuronal signaling, which are crucial for the rapid and accurate processing of olfactory information.

Additionally, we identified several genes within the GPCR family linked to reproduction. GPR18 has been established as a genetic marker associated with phenotypic traits of livestock semen quality [71]. GRM7 has been identified as an important and unique candidate gene associated with the semen volume of Merino sheep [72]. KISS1R has been identified as a major gene affecting sheep reproduction traits [73]. These findings suggest a potential link between olfactory system and reproductive behavior in sheep.

PRKACB, a catalytic subunit of cAMP-dependent protein kinase, mediates olfactory signal transduction by regulating cAMP levels [74]. GNB1, as the β subunit of the G protein family, plays a crucial role in transmembrane signal transduction [75]. MCOLN2 and MCOLN3 belong to the mucolipin family of Ca2+ channel proteins, regulating intracellular Ca2+ levels and neural signal transmission [76, 77]. NALCN specifically mediates sodium leak currents to modulate neuronal excitability [78]. Genes related to mammalian olfactory receptor activity or conduction identified in this study help us understand the unique genetic adaptations evolved by these native sheep breeds under harsh environmental conditions.

Candidate genes related to adipocyte synthesis and energy metabolism

Fat is the main energy reserve in animals, and fat-tailed sheep store excess fat in their tails, helping them adapt to food scarcity and climate changes in arid deserts [79]. In this study, we identified three key pathways specifically linked to fat deposition and energy metabolism: coenzyme A biosynthetic process (GO:0015937), adipocytokine signaling (oas04920), and insulin resistance (oas04931).

Coenzyme A (CoA) is a critical cofactor in fatty acid synthesis, influencing fat deposition. In the CoA biosynthetic pathway, we found that the genes PANK4, PANK3, PPCS, and ACOT7 play pivotal roles. These genes encode enzymes essential for converting pantothenate to 4’-phosphopantothenate and synthesizing 4’-phosphopantothenoylcysteine, a key intermediate in CoA biosynthesis [80, 81, 82]. Additionally, ACOT7 is involved in breaking down acyl-CoA into free fatty acids and CoA [83]. Although these genes collectively play pivotal roles in the CoA biosynthetic process, their specific roles and interactions in fat-tailed sheep have not been previously explored.

The Insulin resistance and adipocytokine signaling pathways involve critical genes such as PRKAA1, SLC2A1, PRKAG2, IRS2, STK11, and IL6. PRKAA1 and PRKAG2 encode subunits of AMP-activated protein kinase (AMPK) [84, 85], which regulates metabolic homeostasis by activating the fatty acid oxidation pathway. SLC2A1 encodes glucose transporter 1 (GLUT1) [86], facilitating glucose transport into cells, essential for fatty acid synthesis and cellular energy production [87]. IRS2 regulates adipocyte metabolism by controlling fat storage and breakdown through its role in insulin signaling [88, 89]. STK11, a key upstream gene in the AMPK cascade, promotes adipogenesis in ruminant preadipocytes [90], while IL6 plays crucial role in insulin resistance and fat metabolism [91]. The candidate genes we identified related to fat deposition and energy metabolism highlight the unique genetic adaptations of indigenous sheep to extreme arid desert environments.

Candidate genes related to desert environment adaptation and immune response

Indigenous sheep from the Taklamakan Desert are exposed to complex physical stressors throughout the year, including significant diurnal temperature fluctuations, drought, and high radiation. We identified some pathways associated with adaptation to desert environments and specific immune responses [92, 93].

In environments with significant temperature variations, animals regulate local circulation to adjust to tissue demands [94]. Vasoconstriction in cold environments increases blood viscosity and reduces flow to minimize heat loss [95], while vasodilation in hot environments decreases blood viscosity and increases flow to aid heat dissipation [96]. The vascular system thus plays a crucial role in maintaining thermal balance. COL4A1 [97], KDR [98], CTNNB1 [99], and SEMA3Ed [100] play a crucial role in blood vessel formation and maintenance of mammalian blood vessels. Notably, KDR facilitates the high-altitude adaptation of Tibetan sheep by regulating angiogenesis [101, 102].

High temperatures enhance hemoglobin oxygenation and dissociation, boosting oxygen transport [103, 104]. To increase tissue oxygen supply, the sympathetic nervous system activates the cardiovascular system, increasing heart rate and cardiac output. We identified two pathways related to cardiac function: regulation of ventricular cardiac muscle cell membrane repolarization and voltage-gated potassium channel complex. Proper repolarization, driven mainly by voltage-gated potassium channels [105], is crucial for maintaining normal cardiac rhythm, contraction, and relaxation.

Blood-fluid shear stress, a biomechanical characteristic [106], exerts force on arterial walls and stimulates endothelial cells to produce nitric oxide (NO) [107]. NO is a crucial signaling molecule that dilates blood vessels, increases blood flow, and inhibits thrombus formation [108]. This explains the involvement of genes like KDR (vascular endothelial growth factor) and NOS3 (encoding nitric oxide synthase) in the pathway of fluid shear stress and atherosclerosis. NOS3 has been reported to regulate the diameter of arterial blood vessels in sheep embryos, causing vasodilation and increasing blood flow, which aids in heat dissipation for indigenous sheep from the southern edge of the Taklamakan Desert in hot environments [109, 110]. Additionally, NO also exhibits anti-inflammatory and anti-platelet aggregation effects, maintaining vascular homeostasis [111, 112].

The immune system is vital for organism health and survival. In desert environments, extreme day-night temperature swings and high variations can disrupt inflammation levels, reducing immune function [113, 114]. Additionally, intense radiation and microorganisms elevate infection risks. To adapt, organisms adjust immune-related gene expression, enhancing responses to specific antigens. The four genes (CSF1, IL6, IL7, IL7R) on the PI3K-Akt signaling pathway are involved in biological processes like immune regulation, inflammatory response, and cell proliferation. While these genes have been previously reported in sheep [115, 116, 117], our study contributes to the understanding of their role in the immune system of desert sheep, emphasizing their novelty and the unique selective pressures they may face. For instance, the role of CSF1 in macrophage differentiation and inflammation regulation may be particularly important in desert environments, where inflammation levels can be disrupted by temperature variations [118, 119]. IL-6 and IL-7, belonging to the interleukin cytokine family, regulate immune cells by mediating the proliferation, differentiation, and activation of T and B cells. Specifically, IL-6 also impacts immune cell activity, participates in inflammatory responses, and regulates metabolic homeostasis [120]. Meanwhile, IL-7 is essential for lymphocyte development and maintaining immune system balance [121], while IL-7R, as the receptor for IL-7, plays a crucial role in IL-7 signal transduction [122]. This is particularly significant when dealing with environmental stressors such as radiation and temperature changes.

Chemotaxis is a crucial process in the immune system, allowing immune cells to sense chemical signals (chemokines) and migrate to inflammatory or lesion sites, participating in pathogen clearance and immune regulation. Our study is the first to report the presence and potential roles of CX3CR1 [123], RARRES2 [124], CCR8 [125], and DOCK2 [126], which are involved in chemotaxis and immune recognition, in desert-dwelling sheep. Similarly, genes involved in the response to gamma radiation, such as FANCD2 [127], XRCC2 [128], and LIG4 [129], have not been previously reported in sheep in desert environments.

Conclusions

This study focused on analyzing LD, Ne, and homozygosity patterns in four distinct indigenous sheep breeds from the southern edge of the Taklamakan Desert in comparison to three foreign commercial meat sheep breeds. Variations in LD and ROH patterns were observed among the four breeds. The QIR genome exhibited the highest inbreeding coefficient and the highest proportion of long ROH segments, indicating relatively recent breeding events. The KUN Ne was smaller, with a low inbreeding coefficient, suggesting appropriate breeding management. Utilizing ROH analysis, we identified genes associated with heat resistance, adaptability to extreme arid environments, energy metabolism, reproduction, and immune response within prominent homozygous islands. These findings significantly impact the adaptability and survival of indigenous sheep populations. These findings will not only enrich our understanding of the genetics of indigenous sheep but will provide essential genetic information for the conservation and rational utilization of this population. Future research can further explore the functional significance of these genes and explore genetic improvement strategies related to indigenous sheep.

Data availability

The datasets analysed during the current study have been submitted to the National Genomics Data Center (NGDC, https://ngdc.cncb.ac.cn/?lang=en) under accession numbers PRJCA027405.

References

  1. Yang F, He Q, Huang J, Mamtimin A, Yang X, Huo W, Zhou C, Liu X, Wei W, Cui C, Wang M, Li H, Yang L, Zhang H, Liu Y, Zheng X, Pan H, Jin L, Zou H, Zhou L, Liu Y, Zhang J, Meng L, Wang Y, Qin X, Yao Y, Liu H, Xue F, Zheng W. Desert environment and climate observation network over the Taklimakan desert. Bull Am Meteorol Soc. 2020;102(6):E1172–91. https://doiorg.publicaciones.saludcastillayleon.es/10.1175/BAMS-D-20-0236.1

    Article  Google Scholar 

  2. Xu X, Wang Y, Wei W, Zhao T, Xu X. Summertime precipitation process and atmospheric water cycle over Tarim basin under the specific large terrain background. Desert Oasis Meteorol. 2014;2014(82):1–11. https://doiorg.publicaciones.saludcastillayleon.es/10.3969/j.issn.1002-0799.2014.02.001

    Article  Google Scholar 

  3. Gui Dongwei Z, Fanjiang L, Zhen Z Bo. Characteristics of the clonal propagation of Alhagi sparsifolia Shap. (Fabaceae) under different groundwater depths in Xinjiang, China. Rangel J. 2013;35:355–62. https://doiorg.publicaciones.saludcastillayleon.es/10.1071/RJ13004

    Article  Google Scholar 

  4. Ma M, Yang X, He Q, Zhou C, Yang F. Characteristics of dust devil and its dust emission in Northern margin of the Taklimakan desert. Aeolian Res. 2020;44:100579. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.aeolia.2020.100579

    Article  Google Scholar 

  5. Wei C, Wang H, Liu G, Wu M, Cao J, Liu Z, Liu R, Zhao F, Zhang L, Lu J, Liu C, Du L. Genome-wide analysis reveals population structure and selection in Chinese indigenous sheep breeds. BMC Genomics. 2015;16(1):194. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12864-015-1384-9

    Article  PubMed  PubMed Central  Google Scholar 

  6. Zhao YX, Yang J, Lv FH, Hu XJ, Xie XL, Zhang M, Li WR, Liu MJ, Wang YT, Li JQ, Liu YG, Ren YL, Wang F, Hehua E, Kantanen J, Arjen Lenstra J, Han JL, Li MH. Genomic reconstruction of the history of native sheep reveals the peopling patterns of nomads and the expansion of early pastoralism in East Asia. Mol Biol Evol. 2017;34(9):2380–95. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/molbev/msx181

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Zhao CH, Wang D, Yang C, Chen Y, Teng J, Zhang XY, Cao Z, Wei XM, Ning C, Yang QE, Lv WF, Zhang Q. Population structure and breed identification of Chinese indigenous sheep breeds using whole genome SNPs and indels. Genet Selection Evolution: GSE. 2024;56(1):60. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12711-024-00927-1

    Article  CAS  PubMed Central  Google Scholar 

  8. Gong S, Liang M, Shi R, Li S. Effects of different concentrations of androgens on KAP24.1 gene expression in Hetian sheep and Karakul sheep. Gene. 2023;878:147591. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.gene.2023.147591

    Article  CAS  PubMed  Google Scholar 

  9. Zhang CL, Zhang J, Tuersuntuoheti M, Chang Q, Liu S. Population structure, genetic diversity and prolificacy in Pishan red sheep under an extreme desert environment. Front Genet. 2023;14:1092066. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fgene.2023.1092066

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Chen HY, Shen H, Jia B, Zhang YS, Wang XH, Zeng XC. Differential gene expression in ovaries of Qira black sheep and Hetian sheep using RNA-Seq technique. PLoS ONE. 2015;10(3):e0120170. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pone.0120170

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Wang J, Suo J, Yang R, Zhang CL, Li X, Han Z, Zhou W, Liu S, Gao Q. Genetic diversity, population structure, and selective signature of sheep in the Northeastern Tarim basin. Front Genet. 2023;14:1281601. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fgene.2023.1281601

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Gibson J, Morton NE, Collins A. Extended tracts of homozygosity in outbred human populations. Hum Mol Genet. 2006;15(5):789–95. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/hmg/ddi493

    Article  CAS  PubMed  Google Scholar 

  13. Smith JM, Haigh J. The hitch-hiking effect of a favourable gene. Genet Res. 2007;89(5–6):391–403. https://doiorg.publicaciones.saludcastillayleon.es/10.1017/S01016672308009579

    Article  PubMed  Google Scholar 

  14. Keller MC, Visscher PM, Goddard ME. Quantification of inbreeding due to distant ancestors and its detection using dense single nucleotide polymorphism data. Genetics. 2011;189(1):237–49. https://doiorg.publicaciones.saludcastillayleon.es/10.1534/genetics.111.130922

    Article  PubMed  PubMed Central  Google Scholar 

  15. Sams AJ, Boyko AR. Fine-scale resolution of runs of homozygosity reveal patterns of inbreeding and substantial overlap with recessive disease genotypes in domestic dogs. G3 (Bethesda). 2019;9(1):117–23. Published 2019 Jan 9. https://doiorg.publicaciones.saludcastillayleon.es/10.1534/g3.118.200836

  16. Szpiech ZA, Xu J, Pemberton TJ, Peng W, Zöllner S, Rosenberg NA, Li JZ. Long runs of homozygosity are enriched for deleterious variation. Am J Hum Genet. 2013;93(1):90–102. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.ajhg.2013.05.003

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Dzomba EF, Chimonyo M, Pierneef R, Muchadeyi FC. Runs of homozygosity analysis of South African sheep breeds from various production systems investigated using OvineSNP50k data. BMC Genomics. 2021;22(1):7. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12864-020-07314-2

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Malesa MT. Population genetics of Swakara sheep inferred using genome-wide SNP genotyping. MSc Thesis: University of KwaZulu-Natal; 2015.

  19. Sallam AM, Reyer H, Wimmers K, Bertolini F, Aboul-Naga A, Braz CU, Rabee AE. Genome-wide landscape of runs of homozygosity and differentiation across Egyptian goat breeds. BMC Genomics. 2023;24(1):573. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12864-023-09679-6

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Eydivandi S, Sahana G, Momen M, Moradi MH, Schönherz AA. Genetic diversity in Iranian indigenous sheep vis-à-vis selected exogenous sheep breeds and wild mouflon. Anim Genet. 2020;51(5):772–87. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/age.12985

    Article  CAS  PubMed  Google Scholar 

  21. Kim ES, Cole JB, Huson H, Wiggans GR, Van Tassell CP, Crooker BA, Liu G, Da Y, Sonstegard TS. Effect of artificial selection on runs of homozygosity in U.s. Holstein cattle. PLoS ONE. 2013;8(11):e80813. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pone.0080813

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Pemberton TJ, Absher D, Feldman MW, Myers RM, Rosenberg NA, Li JZ. Genomic patterns of homozygosity in worldwide human populations. Am J Hum Genet. 2012;91(2):275–92. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.ajhg.2012.06.014

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Zhang Q, Guldbrandtsen B, Bosse M, Lund MS, Sahana G. Runs of homozygosity and distribution of functional variants in the cattle genome. BMC Genomics. 2015;16(1):542. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12864-015-1715-x

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Liu J, Shi L, Li Y, Chen L, Garrick D, Wang L, Zhao F. Estimates of genomic inbreeding and identification of candidate regions that differ between Chinese indigenous sheep breeds. J Anim Sci Biotechnol. 2021;12(1):95. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s40104-021-00608-9

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Mastrangelo S, Tolone M, Sardina MT, Sottile G, Sutera AM, Di Gerlando R, Portolano B. Genome-wide scan for runs of homozygosity identifies potential candidate genes associated with local adaptation in Valle Del Belice sheep. Genet Selection Evolution: GSE. 2017;49(1):84. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12711-017-0360-z

    Article  CAS  PubMed Central  Google Scholar 

  26. He S, Di J, Han B, Chen L, Liu M, Li W. Genome-Wide scan for runs of homozygosity identifies candidate genes related to economically important traits in Chinese Merino. Animals: Open Access J MDPI. 2020;10(3):524. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/ani10030524

    Article  Google Scholar 

  27. Liu S, He S, Chen L, Li W, Di J, Liu M. Estimates of linkage disequilibrium and effective population sizes in Chinese Merino (Xinjiang type) sheep by genome-wide SNPs. Genes Genomics. 2017;39(7):733–45. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s13258-017-0539-2

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Huff CD, Harpending HC, Rogers AR. Detecting positive selection from genome scans of linkage disequilibrium. BMC Genomics. 2010;11:8. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/1471-2164-11-8

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Charlesworth B. Effective population size and patterns of molecular evolution and variation. Nat Rev Genet. 2009;10:195–205. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/nrg2526

    Article  CAS  PubMed  Google Scholar 

  30. Goddard M. Genomic selection: prediction of accuracy and maximisation of long-term response. Genetica. 2009;136(2):245–57. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s10709-008-9308-0

    Article  PubMed  Google Scholar 

  31. Purcell S, Neale B, Todd-Brown K, Thomas L, Ferreira MA, Bender D, Maller J, Sklar P, De Bakker PI, Daly MJ. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am J Hum Genet. 2007;81:559–75. https://doiorg.publicaciones.saludcastillayleon.es/10.1086/519795

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Zhao H, Nettleton D, Dekkers JC. Evaluation of linkage disequilibrium measures between multi-allelic markers as predictors of linkage disequilibrium between single nucleotide polymorphisms. Genet Res. 2007;89(1):1–6. https://doiorg.publicaciones.saludcastillayleon.es/10.1017/S0016672307008634

    Article  CAS  PubMed  Google Scholar 

  33. Hill WG, Robertson A. Linkage disequilibrium in finite populations. TAG. Theoretical and applied genetics. Theoretische Und Angewandte Genetik. 1968;38(6):226–31. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/BF01245622

    Article  CAS  PubMed  Google Scholar 

  34. Zhang C, Dong SS, Xu JY, He WM, Yang TL. PopLDdecay: a fast and effective tool for linkage disequilibrium decay analysis based on variant call format files. Bioinf (Oxford England). 2019;35(10):1786–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/bioinformatics/bty875

  35. Eriksson A, Mahjani B, Mehlig B. Sequential Markov coalescent algorithms for population models with demographic structure. Theor Popul Biol. 2009;76(2):84–91. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.tpb.2009.05.002

    Article  CAS  PubMed  Google Scholar 

  36. Terhorst J, Kamm JA, Song YS. Robust and scalable inference of population history from hundreds of unphased whole genomes. Nat Genet. 2017;49(2):303–9. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/ng.3748

    Article  CAS  PubMed  Google Scholar 

  37. Chen, L., Qiu, Q., Jiang, Y., Wang, K., Lin, Z., Li, Z., Bibi, F., Yang, Y., Wang,J., Nie, W., Su, W., Liu, G., Li, Q., Fu, W., Pan, X., Liu, C., Yang, J., Zhang, C.,Yin, Y., Wang, Y.,… Wang, W. Large-scale ruminant genome sequencing provides insights into their evolution and distinct traits. Science (New York, N.Y.). 2019;364(6446),eaav6202. https://doiorg.publicaciones.saludcastillayleon.es/10.1126/science.aav6202.

  38. Lencz T, Lambert C, DeRosse P, Burdick KE, Morgan TV, Kane JM, Kucherlapati R, Malhotra AK. Runs of homozygosity reveal highly penetrant recessive loci in schizophrenia. Proc Natl Acad Sci USA. 2007;104(50):19942–7. https://doiorg.publicaciones.saludcastillayleon.es/10.1073/pnas.0710021104

    Article  PubMed  PubMed Central  Google Scholar 

  39. Islam R, Li Y, Liu X, Berihulay H, Abied A, Gebreselassie G, Ma Q, Ma Y. Genome-wide runs of homozygosity, effective population size, and detection of positive selection signatures in six Chinese goat breeds. Genes. 2019;10(11):938. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/genes10110938

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Huang daW, Sherman BT, Lempicki RA. Bioinformatics enrichment tools: paths toward the comprehensive functional analysis of large gene lists. Nucleic Acids Res. 2009;37(1):1–13. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/nar/gkn923

    Article  CAS  PubMed  Google Scholar 

  41. Abied A, Xu L, Sahlu BW, Xing F, Ahbara A, Pu Y, Lin J, Berihulay H, Islam R, He X, Mwacharo JM, Zhao Q, Ma Y. Genome-wide analysis revealed homozygosity and demographic history of five Chinese sheep breeds adapted to different environments. Genes. 2020;11(12):1480. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/genes11121480

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Robin VV, Gupta P, Thatte P, Ramakrishnan U. Islands within islands: two montane palaeo-endemic birds impacted by recent anthropogenic fragmentation. Mol Ecol. 2015;24(14):3572–84. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/mec.13266

    Article  CAS  PubMed  Google Scholar 

  43. Carnicero P, Wessely J, Moser D, Font X, Dullinger S, Schönswetter P. Postglacial range expansion of high-elevation plants is restricted by dispersal ability and habitat specialization. J Biogeogr. 2022;49(10):1739–52. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/jbi.14390

    Article  PubMed  PubMed Central  Google Scholar 

  44. Niu ZG, Qin J, Jiang Y, Ding XD, Ding YG, Tang S, Shi HC. The identification of mutation in BMP15 gene associated with litter size in Xinjiang cele black sheep. Animals: Open Access J MDPI. 2021;11(3):668. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/ani11030668

    Article  Google Scholar 

  45. Beaujard P. Tang China and the rise of the silk roads. The worlds of the Indian Ocean: a global history. Cambridge: Cambridge University Press; 2019. pp. 18–41. https://doiorg.publicaciones.saludcastillayleon.es/10.1017/9781108341219.002

    Chapter  Google Scholar 

  46. Qian FQ. Ancient routes, new dream: the silk roads and China’s belt and road initiative. J Cult Herit Manage Sustainable Dev. 2022;12(1):45–57. https://doiorg.publicaciones.saludcastillayleon.es/10.1108/JCHMSD-05-2021-0091

    Article  CAS  Google Scholar 

  47. Chen FH, An CB, Dong GH, et al. Human activities, environmental changes, and rise and decline of silk road civilization in pan-third pole region. Bull Chin Acad Sci. 2017;32(9):967–75.

    Google Scholar 

  48. Wang J, Chen H, Zeng X. Identification of hub genes associated with follicle development in multiple births sheep by WGCNA. Front Veterinary Sci. 2022;9:1057282. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fvets.2022.1057282

    Article  Google Scholar 

  49. Beckwith CI. Empires of the silk road: A history of central Eurasia from the bronze age to the present. USA: Princeton University Press; 2009. http://www.jstor.org/stable/j.ctt7sq0r

    Book  Google Scholar 

  50. Han Z, Zhou W, Zhang L, Wang R, Liu C, Bai X, Liu S. Genetic diversity and runs of homozygosity analysis of Hetian sheep populations revealed by illumina ovine SNP50K BeadChip. Front Ecol Evol. 2023;11:1182966. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fevo.2023.1182966

    Article  Google Scholar 

  51. Kominakis A, Tarsani E, Hager-Theodorides AL, Mastranestasis I, Hadjigeorgiou I. Clustering patterns mirror the geographical distribution and genetic history of Lemnos and Lesvos sheep populations. PLoS ONE. 2021;16(3):e0247787. https://doiorg.publicaciones.saludcastillayleon.es/10.1371/journal.pone.0247787. Published 2021 Mar 3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Eydivandi S, Roudbar MA, Karimi MO, Sahana G. Genomic scans for selective sweeps through haplotype homozygosity and allelic fixation in 14 indigenous sheep breeds from middle East and South Asia. Sci Rep. 2021;11(1):2834. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41598-021-82625-2

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Zhou M, Pan Z, Cao X, Guo X, He X, Sun Q, Di R, Hu W, Wang X, Zhang X, Zhang J, Zhang C, Liu Q, Chu M. Single nucleotide polymorphisms in the HIRA gene affect litter size in small tail Han sheep. Animals: Open Access J MDPI. 2018;8(5):71. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/ani8050071

    Article  Google Scholar 

  54. Pan Z, Li S, Liu Q, Wang Z, Zhou Z, Di R, Miao B, Hu W, Wang X, Hu X, Xu Z, Wei D, He X, Yuan L, Guo X, Liang B, Wang R, Li X, Cao X, Dong X, Li Y. 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(4):giy019. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/gigascience/giy019

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Wang W, Zhang X, Zhou X, Zhang Y, La Y, Zhang Y, Li C, Zhao Y, Li F, Liu B, Jiang Z. 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. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fgene.2019.00300

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Ramos-Lopez O, Riezu-Boj JI, Milagro FI, Zulet MA, Santos JL, Martinez JA, MENA project. Associations between olfactory pathway gene methylation marks, obesity features and dietary intakes. Genes Nutr. 2019;14:11. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12263-019-0635-9

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Clavijo McCormick A, Unsicker SB, Gershenzon J. The specificity of herbivore-induced plant volatiles in attracting herbivore enemies. Trends Plant Sci. 2012;17(5):303–10. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.tplants.2012.03.012

    Article  CAS  PubMed  Google Scholar 

  58. Cann P, Le Danvic C, Porte C, Chesneau D, Keller M, Meillour N-L, P. Variation of Ewe olfactory secretome during a Ram effect. Front Veterinary Sci. 2023;9:1033412. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fvets.2022.1033412

    Article  Google Scholar 

  59. Buck L, Axel R. A novel multigene family may encode odorant receptors: a molecular basis for odor recognition. Cell. 1991;65(1):175–87. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/0092-8674(91)90418-x

    Article  CAS  PubMed  Google Scholar 

  60. Hauser AS, Attwood MM, Rask-Andersen M, Schiöth HB, Gloriam DE. Trends in GPCR drug discovery: new agents, targets and indications. Nat Rev Drug Discov. 2017;16(12):829–42. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/nrd.2017.178

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Gilman AG. G proteins: transducers of receptor-generated signals. Annu Rev Biochem. 1987;56:615–49. https://doiorg.publicaciones.saludcastillayleon.es/10.1146/annurev.bi.56.070187.003151

    Article  CAS  PubMed  Google Scholar 

  62. Boekhoff I, Breer H. Differential stimulation of second messenger pathways by distinct classes of odorants. Neurochem Int. 1990;17(4):553–7. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/0197-0186(90)90043-s

    Article  CAS  PubMed  Google Scholar 

  63. Nakamura T, Gold GH. A cyclic nucleotide-gated conductance in olfactory receptor cilia. Nature. 1987;325(6103):442–4. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/325442a0

    Article  CAS  PubMed  Google Scholar 

  64. Su CY, Menuz K, Carlson JR. Olfactory perception: receptors, cells, and circuits. Cell. 2009;139(1):45–59. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.cell.2009.09.015

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Xu F, Schaefer M, Kida I, Schafer J, Liu N, Rothman DL, Hyder F, Restrepo D, Shepherd GM. Simultaneous activation of mouse main and accessory olfactory bulbs by odors or pheromones. J Comp Neurol. 2005;489(4):491–500. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/cne.20652

    Article  PubMed  Google Scholar 

  66. Chen B, Wang L, Li X, Shi Z, Duan J, Wei JA, Li C, Pang C, Wang D, Zhang K, Chen H, Na W, Zhang L, So KF, Zhou L, Jiang B, Yuan TF, Qu Y. Celsr2 regulates NMDA receptors and dendritic homeostasis in dorsal CA1 to enable social memory. Mol Psychiatry. 2022. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41380-022-01664-x. Advance online publication.

    Article  PubMed  PubMed Central  Google Scholar 

  67. Lee JH, Koh SQ, Guadagna S, Francis PT, Esiri MM, Chen CP, Wong PT, Dawe GS, Lai MK. Altered relaxin family receptors RXFP1 and RXFP3 in the neocortex of depressed Alzheimer’s disease patients. Psychopharmacology. 2016;233(4):591–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s00213-015-4131-7

    Article  CAS  PubMed  Google Scholar 

  68. Blythe EE, von Zastrow M. β-Arrestin-independent endosomal cAMP signaling by a polypeptide hormone GPCR. Nat Chem Biol. 2023. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41589-023-01412-4. Advance online publication.

    Article  PubMed  PubMed Central  Google Scholar 

  69. Freitas GA, Niswender CM. GRM7 gene mutations and consequences for neurodevelopment. Pharmacol Biochem Behav. 2023;225:173546. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.pbb.2023.173546

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Vidal OM, Vélez JI, Arcos-Burgos M. ADGRL3 genomic variation implicated in neurogenesis and ADHD links functional effects to the incretin polypeptide GIP. Sci Rep. 2022;12(1):15922. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41598-022-20343-z

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Khan MZ, Chen W, Naz S, Liu X, Liang H, Chen Y, Kou X, Liu Y, Ashraf I, Han Y, Peng Y, Wang C, Zahoor M. Determinant genetic markers of semen quality in livestock. Front Endocrinol. 2024;15:1456305. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fendo.2024.1456305

    Article  Google Scholar 

  72. Hodge MJ, de Las Heras-Saldana S, Rindfleish SJ, Stephen CP, Pant SD. QTLs and candidate genes associated with semen traits in Merino sheep. Animals: Open Access J MDPI. 2023;13(14):2286. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/ani13142286

    Article  Google Scholar 

  73. Majd SA, Ahmadi A, Talebi R, Koohi PM, Fabre S, Qanbari S. Polymorphism identification in ovine KISS1R / GPR54 gene among pure and crossbreeds of Iranian sheep. Small Ruminant Res. 2019;173:23–9. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.smallrumres.2019.02.005

    Article  Google Scholar 

  74. Taylor SS, Wallbott M, Machal EMF, Søberg K, Ahmed F, Bruystens J, Vu L, Baker B, Wu J, Raimondi F, Ongeri EM, Herberg FW, Skålhegg BS. PKA Cβ: a forgotten catalytic subunit of cAMP-dependent protein kinase opens new windows for PKA signaling and disease pathologies. Biochem J. 2021;478(11):2101–19. https://doiorg.publicaciones.saludcastillayleon.es/10.1042/BCJ20200867

    Article  CAS  PubMed  Google Scholar 

  75. Zheng H, Ma L, Gui R, Lin X, Ke X, Jian X, Ye C, Chen Q. G protein subunit Β1 facilitates influenza A virus replication by promoting the nuclear import of PB2. J Virol. 2022;96(12):e0049422. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/jvi.00494-22

    Article  CAS  PubMed  Google Scholar 

  76. Yu H, Xie M, Meng Z, Lo CY, Chan FL, Jiang L, Meng X, Yao X. Endolysosomal ion channel MCOLN2 (Mucolipin-2) promotes prostate cancer progression via IL-1β/NF-κB pathway. Br J Cancer. 2021;125(10):1420–31. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41416-021-01537-0

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Lei Y, Klionsky DJ. MCOLN3/TRPML3 bridges the regulation of autophagosome biogenesis by PtdIns3P and the calcium channel. Autophagy. 2023;19(2):377–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1080/15548627.2022.2148808

    Article  CAS  PubMed  Google Scholar 

  78. Kang Y, Chen L. Structure and mechanism of NALCN-FAM155A-UNC79-UNC80 channel complex. Nat Commun. 2022;13(1):2639. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41467-022-30403-7

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Bakhtiarizadeh MR, Salami SA. Identification and expression analysis of long noncoding RNAs in fat-tail of sheep breeds. G3 (Bethesda Md). 2019;9(4):1263–76. https://doiorg.publicaciones.saludcastillayleon.es/10.1534/g3.118.201014

    Article  CAS  PubMed  Google Scholar 

  80. Dibble CC, Barritt SA, Perry GE, Lien EC, Geck RC, DuBois-Coyne SE, Bartee D, Zengeya TT, Cohen EB, Yuan M, Hopkins BD, Meier JL, Clohessy JG, Asara JM, Cantley LC, Toker A. PI3K drives the de novo synthesis of coenzyme A from vitamin B5. Nature. 2022;608(7921):192–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41586-022-04984-8

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Leonardi R, Zhang YM, Yun MK, Zhou R, Zeng FY, Lin W, Cui J, Chen T, Rock CO, White SW, Jackowski S. Modulation of pantothenate kinase 3 activity by small molecules that interact with the substrate/allosteric regulatory domain. Chem Biol. 2010;17(8):892–902. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.chembiol.2010.06.006

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Mostert KJ, Sharma N, van der Zwaag M, Staats R, Koekemoer L, Anand R, Sibon OCM, Strauss E. The coenzyme A level modulator hopantenate (HoPan) inhibits phosphopantotenoylcysteine synthetase activity. ACS Chem Biol. 2021;16(11):2401–14. https://doiorg.publicaciones.saludcastillayleon.es/10.1021/acschembio.1c00535

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Jung SH, Lee HC, Hwang HJ, Park HA, Moon YA, Kim BC, Lee HM, Kim KP, Kim YN, Lee BL, Lee JC, Ko YG, Park HJ, Lee JS. Acyl-CoA thioesterase 7 is involved in cell cycle progression via regulation of PKCζ-p53-p21 signaling pathway. Cell Death Dis. 2017;8(5):e2793. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/cddis.2017.202

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Peng HY, Li MQ, Li HP. High glucose suppresses the viability and proliferation of HTR 8/svneo cells through regulation of the MiR 137/PRKAA1/IL 6 axis. Int J Mol Med. 2018;42(2):799–810. https://doiorg.publicaciones.saludcastillayleon.es/10.3892/ijmm.2018.3686

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Lopez-Sainz, A., Dominguez, F., Lopes, L. R., Ochoa, J. P., Barriales-Villa, R., Climent,V., Linschoten, M., Tiron, C., Chiriatti, C., Marques, N., Rasmussen, T. B., Espinosa,M. Á., Beinart, R., Quarta, G., Cesar, S., Field, E., Garcia-Pinilla, J. M., Bilinska,Z., Muir, A. R., Roberts, A. M.,… European Genetic Cardiomyopathies Initiative Investigators. Clinical features and natural history of PRKAG2 variant cardiac glycogenosis. J Am Coll Cardiol. 2020;76(2):186–197. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.jacc.2020.05.029

  86. Cao S, Chen Y, Ren Y, Feng Y, Long S. GLUT1 biological function and inhibition: research advances. Future Med Chem. 2021;13(14):1227–43. https://doiorg.publicaciones.saludcastillayleon.es/10.4155/fmc-2021-0071

    Article  CAS  PubMed  Google Scholar 

  87. Lloyd KP, Ojelabi OA, De Zutter JK, Carruthers A. Reconciling contradictory findings: glucose transporter 1 (GLUT1) functions as an oligomer of allosteric, alternating access transporters. J Biol Chem. 2017;292(51):21035–46. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.M117.815589

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Previs SF, Withers DJ, Ren JM, White MF, Shulman GI. Contrasting effects of IRS-1 versus IRS-2 gene disruption on carbohydrate and lipid metabolism in vivo. J Biol Chem. 2000;275(50):38990–4. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.M006490200

    Article  CAS  PubMed  Google Scholar 

  89. Abel ED, Peroni O, Kim JK, Kim YB, Boss O, Hadro E, Minnemann T, Shulman GI, Kahn BB. Adipose-selective targeting of the GLUT4 gene impairs insulin action in muscle and liver. Nature. 2001;409(6821):729–33. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/35055575

    Article  CAS  PubMed  Google Scholar 

  90. Wang L, Zhang S, Zhang W, Cheng G, Khan R, Junjvlieke Z, Li S, Zan L. miR-424 promotes bovine adipogenesis through an unconventional Post-Transcriptional regulation of STK11. Front Genet. 2020;11:145. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fgene.2020.00145

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Valin A, Del Rey MJ, Municio C, Usategui A, Romero M, Fernández-Felipe J, Cañete JD, Blanco FJ, Ruano Y, Criado G, Pablos JL. IL6/sIL6R regulates TNFα-inflammatory response in synovial fibroblasts through modulation of transcriptional and post-transcriptional mechanisms. BMC Mol Cell Biology. 2020;21(1):74. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12860-020-00317-7

    Article  CAS  Google Scholar 

  92. Fried SK, Bunkin DA, Greenberg AS. Omental and subcutaneous adipose tissues of obese subjects release interleukin-6: depot difference and regulation by glucocorticoid. J Clin Endocrinol Metab. 1998;83(3):847–50. https://doiorg.publicaciones.saludcastillayleon.es/10.1210/jcem.83.3.4660

    Article  CAS  PubMed  Google Scholar 

  93. Díaz-Chamorro S, Garrido-Jiménez S, Barrera-López JF, Mateos-Quirós CM, Cumplido-Laso G, Lorenzo MJ, Román ÁC, Bernardo E, Sabio G, Carvajal-González JM, Centeno F. Title: p38δ regulates IL6 expression modulating ERK phosphorylation in preadipocytes. Front Cell Dev Biology. 2022;9:708844. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fcell.2021.708844

    Article  Google Scholar 

  94. Fernández-Peña C, Reimúndez A, Viana F, Arce VM, Señarís R. Sex differences in thermoregulation in mammals: implications for energy homeostasis. Front Endocrinol. 2023;14:1093376. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fendo.2023.1093376

    Article  Google Scholar 

  95. Señarís R, Ordás P, Reimúndez A, Viana F. Mammalian cold TRP channels: impact on thermoregulation and energy homeostasis. Pflug Arch: Eur J Physiol. 2018;470(5):761–77. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s00424-018-2145-9

    Article  CAS  Google Scholar 

  96. Uchida K, Dezaki K, Yoneshiro T, Watanabe T, Yamazaki J, Saito M, Yada T, Tominaga M, Iwasaki Y. Involvement of thermosensitive TRP channels in energy metabolism. J Physiol Sci. 2017;67(5):549–60. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s12576-017-0552-x

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Angbohang A, Huang L, Li Y, Zhao Y, Gong Y, Fu Y, Mao C, Morales J, Luo P, Ehteramyan M, Gao Y, Margariti A, Gu W, Zhang M, Smith A, Shah AM, Li T, Kong W, Zeng L. X-box binding protein 1-mediated COL4A1s secretion regulates communication between vascular smooth muscle and stem/progenitor cells. J Biol Chem. 2021;296:100541. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.jbc.2021.100541

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Du J, Teng RJ, Guan T, Eis A, Kaul S, Konduri GG, Shi Y. Role of autophagy in angiogenesis in aortic endothelial cells. Am J Physiol Cell Physiol. 2012;302(2):C383–91. https://doiorg.publicaciones.saludcastillayleon.es/10.1152/ajpcell.00164.2011

    Article  CAS  PubMed  Google Scholar 

  99. Wei Y, Jiao Z, Sun T, Lai Z, Wang X. Molecular mechanisms behind vascular mimicry as the target for improved breast cancer management. Int J Women’s Health. 2023;15:1027–38. https://doiorg.publicaciones.saludcastillayleon.es/10.2147/IJWH.S406327

    Article  CAS  Google Scholar 

  100. Casazza A, Finisguerra V, Capparuccia L, Camperi A, Swiercz JM, Rizzolio S, Rolny C, Christensen C, Bertotti A, Sarotto I, Risio M, Trusolino L, Weitz J, Schneider M, Mazzone M, Comoglio PM, Tamagnone L. Sema3E-Plexin D1 signaling drives human cancer cell invasiveness and metastatic spreading in mice. J Clin Investig. 2010;120(8):2684–98. https://doiorg.publicaciones.saludcastillayleon.es/10.1172/JCI42118

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Wu HB, Yang S, Weng HY, Chen Q, Zhao XL, Fu WJ, Niu Q, Ping YF, Wang JM, Zhang X, Yao XH, Bian XW. Autophagy-induced KDR/VEGFR-2 activation promotes the formation of vasculogenic mimicry by glioma stem cells. Autophagy. 2017;13(9):1528–42. https://doiorg.publicaciones.saludcastillayleon.es/10.1080/15548627.2017.1336277

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Jin M, Lu J, Fei X, Lu Z, Quan K, Liu Y, Chu M, Di R, Wei C, Wang H. Selection signatures analysis reveals genes associated with high-altitude adaptation in Tibetan goats from Nagqu, Tibet. Animals: Open Access J MDPI. 2020;10(9):1599. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/ani10091599

    Article  Google Scholar 

  103. Horsman MR. Tissue physiology and the response to heat. International journal of hyperthermia: the official journal of European society for hyperthermic oncology. North Am Hyperth Group. 2006;22(3):197–203. https://doiorg.publicaciones.saludcastillayleon.es/10.1080/02656730600689066

    Article  Google Scholar 

  104. Song CW, Park HJ, Lee CK, Griffin R. Implications of increased tumor blood flow and oxygenation caused by mild temperature hyperthermia in tumor treatment. Int J Hyperthermia: Official J Eur Soc Hyperthermic Oncol North Am Hyperth Group. 2005;21(8):761–7. https://doiorg.publicaciones.saludcastillayleon.es/10.1080/02656730500204487

    Article  CAS  Google Scholar 

  105. Chen L, He Y, Wang X, Ge J, Li H. Ventricular voltage-gated ion channels: detection, characteristics, mechanisms, and drug safety evaluation. Clin Translational Med. 2021;11(10):e530. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/ctm2.530

    Article  CAS  Google Scholar 

  106. Giddens DP, Zarins CK, Glagov S, May 1. (, 1990). Response of arteries to near-wall fluid dynamic behavior. ASME Appl Mech Rev. 1990;43(5S):S98–S102. https://doiorg.publicaciones.saludcastillayleon.es/10.1115/1.3120861

  107. Meza D, Musmacker B, Steadman E, Stransky T, Rubenstein DA, Yin W. Endothelial cell biomechanical responses are dependent on both fluid shear stress and tensile strain. Cell Mol Bioeng. 2019;12(4):311–25. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s12195-019-00585-0

    Article  PubMed  PubMed Central  Google Scholar 

  108. Fang Y, Wu D, Birukov KG. Mechanosensing and mechanoregulation of endothelial cell functions. Compr Physiol. 2019;9(2):873–904. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/cphy.c180020

    Article  PubMed  PubMed Central  Google Scholar 

  109. Grazul-Bilska AT, Johnson ML, Borowicz PP, Bilski JJ, Cymbaluk T, Norberg S, Redmer DA, Reynolds LP. Placental development during early pregnancy in sheep: effects of embryo origin on vascularization. Reprod (Cambridge England). 2014;147(5):639–48. https://doiorg.publicaciones.saludcastillayleon.es/10.1530/REP-13-0663

    Article  CAS  Google Scholar 

  110. Ge, R. L., Cai, Q., Shen, Y. Y., San, A., Ma, L., Zhang, Y., Yi, X., Chen, Y., Yang,L., Huang, Y., He, R., Hui, Y., Hao, M., Li, Y., Wang, B., Ou, X., Xu, J., Zhang,Y., Wu, K., Geng, C.,… Wang, J. Draft genome sequence of the Tibetan antelope.Nature communications. 2013;4:1858. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/ncomms2860.

  111. Garg UC, Hassid A. Nitric oxide-generating vasodilators and 8-bromo-cyclic guanosine monophosphate inhibit mitogenesis and proliferation of cultured rat vascular smooth muscle cells. J Clin Investig. 1989;83(5):1774–7. https://doiorg.publicaciones.saludcastillayleon.es/10.1172/JCI114081

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Ignarro LJ. Biological actions and properties of endothelium-derived nitric oxide formed and released from artery and vein. Circul Res. 1989;65(1):1–21. https://doiorg.publicaciones.saludcastillayleon.es/10.1161/01.res.65.1.1

    Article  CAS  Google Scholar 

  113. Fang M, Lei Z, Ruilin M, Jing W, Leqiang D. High temperature stress induced oxidative stress, gut inflammation and disordered metabolome and microbiome in Tsinling Lenok trout. Ecotoxicol Environ Saf. 2023;266:115607. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.ecoenv.2023.115607

    Article  CAS  PubMed  Google Scholar 

  114. Appenheimer MM, Evans SS. Temperature and adaptive immunity. Handb Clin Neurol. 2018;156:397–415. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/B978-0-444-63912-7.00024-2

  115. Bush SJ, McCulloch MEB, Lisowski ZM, Muriuki C, Clark EL, Young R, Pridans C, Prendergast JGD, Summers KM, Hume DA. Species-specificity of transcriptional regulation and the response to lipopolysaccharide in mammalian macrophages. Front Cell Dev Biology. 2020;8:661. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fcell.2020.00661

    Article  Google Scholar 

  116. Shi X, Zhang Y, Chen S, Du X, Zhang P, Duan X, Fang H, Liu S. Differential gene expression and immune cell infiltration in maedi-visna virus-infected lung tissues. BMC Genomics. 2024;25(1):534. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12864-024-10448-2

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Yan S, Du R, Yao W, Zhang H, Xue Y, Teligun, Li Y, Bao H, Zhao Y, Cao S, Cao G, Li X, Bao S, Song Y. Host-microbe interaction-mediated resistance to DSS-induced inflammatory enteritis in sheep. Microbiome. 2024;12(1):208. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s40168-024-01932-8

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Lin, W., Xu, D., Austin, C. D., Caplazi, P., Senger, K., Sun, Y., Jeet, S., Young,J., Delarosa, D., Suto, E., Huang, Z., Zhang, J., Yan, D., Corzo, C., Barck, K., Rajan,S., Looney, C., Gandham, V., Lesch, J., Liang, W. C.,… Zarrin, A. A. Function of CSF1 and IL34 in macrophage homeostasis, inflammation, and cancer. Front Immunol. 2019;10:2019. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fimmu.2019.02019.

  119. Geissmann F, Manz MG, Jung S, Sieweke MH, Merad M, Ley K. Development of monocytes, macrophages, and dendritic cells. Volume 327. Science (New York; 2010. pp. 656–61. 5966https://doiorg.publicaciones.saludcastillayleon.es/10.1126/science.1178331

  120. Uciechowski P, Dempke WCM. Interleukin-6: a masterplayer in the cytokine network. Oncology. 2020;98(3):131–7. https://doiorg.publicaciones.saludcastillayleon.es/10.1159/000505099

    Article  CAS  PubMed  Google Scholar 

  121. Winer H, Rodrigues GOL, Hixon JA, Aiello FB, Hsu TC, Wachter BT, Li W, Durum SK. IL-7: comprehensive review. Cytokine. 2022;160:156049. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.cyto.2022.156049

    Article  CAS  PubMed  Google Scholar 

  122. Almeida, A. R. M., Neto, J. L., Cachucho, A., Euzébio, M., Meng, X., Kim, R., Fernandes,M. B., Raposo, B., Oliveira, M. L., Ribeiro, D., Fragoso, R., Zenatti, P. P., Soares,T., de Matos, M. R., Corrêa, J. R., Duque, M., Roberts, K. G., Gu, Z., Qu, C., Pereira,C.,… Barata, J. T. Interleukin-7 receptor α mutational activation can initiate precursor B-cell acute lymphoblastic leukemia. Nat Commun. 2021;12(1):7268. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41467-021-27197-5.

  123. Zegarra-Ruiz DF, Kim DV, Norwood K, Kim M, Wu WH, Saldana-Morales FB, Hill AA, Majumdar S, Orozco S, Bell R, Round JL, Longman RS, Egawa T, Bettini ML, Diehl GE. Thymic development of gut-microbiota-specific T cells. Nature. 2021;594(7863):413–7. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41586-021-03531-1

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Liu-Chittenden Y, Jain M, Gaskins K, Wang S, Merino MJ, Kotian S, Kumar Gara S, Davis S, Zhang L, Kebebew E. RARRES2 functions as a tumor suppressor by promoting β-catenin phosphorylation/degradation and inhibiting p38 phosphorylation in adrenocortical carcinoma. Oncogene. 2017;36(25):3541–52. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/onc.2016.497

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Sun, D., Sun, Y., Janezic, E., Zhou, T., Johnson, M., Azumaya, C., Noreng, S., Chiu,C., Seki, A., Arenzana, T. L., Nicoludis, J. M., Shi, Y., Wang, B., Ho, H., Joshi,P., Tam, C., Payandeh, J., Comps-Agrar, L., Wang, J., Rutz, S.,… Masureel, M. Structural basis of antibody inhibition and chemokine activation of the human CC chemokine receptor 8. Nat Commun. 2023;14(1):7940. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41467-023-43601-8.

  126. Ji L, Xu S, Luo H, Zeng F. Insights from DOCK2 in cell function and pathophysiology. Front Mol Biosci. 2022;9:997659. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fmolb.2022.997659

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Nepal M, Che R, Ma C, Zhang J, Fei P. FANCD2 and DNA damage. Int J Mol Sci. 2017;18(8):1804. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/ijms18081804

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Zhao Z, He K, Zhang Y, Hua X, Feng M, Zhao Z, Sun Y, Jiang Y, Xia Q. XRCC2 repairs mitochondrial DNA damage and fuels malignant behavior in hepatocellular carcinoma. Cancer Lett. 2021;512:1–14. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.canlet.2021.04.026

    Article  CAS  PubMed  Google Scholar 

  129. Jun S, Jung YS, Suh HN, Wang W, Kim MJ, Oh YS, Lien EM, Shen X, Matsumoto Y, McCrea PD, Li L, Chen J, Park JI. LIG4 mediates Wnt signalling-induced radioresistance. Nat Commun. 2016;7:10994. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/ncomms10994

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We would like to thank everyone who contributed towards the article.

Funding

This study was funded by grants from the Natural Science Foundation of China (NO:32060743), Bintuan Science and Technology Program (NO: 2022CB001-09), and Autonomous Region Agricultural Area High-efficiency Mutton Sheep Breeding and Promotion Technology System Project (NO: xjnqry-g-2310).

Author information

Authors and Affiliations

Authors

Contributions

Z.-P.H. and S.-D.L. conceived and designed the experiments; Z.-P.H, L.-L.Z., and X.-P.L. did the data analyses; L.-J.Z., X.-C.Z., and W.Z.undertook sampling and discussed and interpreted genetic data in light of breeding evidence; Z.-P.H and S.-D.L.wrote the manuscript with input from the other authors. All authors have read and agreed to the published version of the manuscript.

Corresponding author

Correspondence to Shudong Liu.

Ethics declarations

Ethics approval and consent to participate

This work was conducted in accordance with the standards set by the Ethics Committee of of the College of Animal Science and Technology of Tarim University (SYXK2020-009) and approved by the Ethics Committee of of the College of Animal Science and Technology of Tarim University. Written informed consent was obtained from the owners for the participation of their animals in this study.

All methods are reported in accordance with ARRIVE guidelines (https://arriveguidelines.org/) for the reporting of animal experiments.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Benefit sharing

The authors declare no benefit sharing.

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Han, Zp., Zhang, Ll., Li, Xp. et al. Single nucleotide polymorphism-based analysis of linkage disequilibrium and runs of homozygosity patterns of indigenous sheep in the southern Taklamakan desert. BMC Genomics 26, 267 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12864-025-11445-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12864-025-11445-9

Keywords