Skip to main content

Association of a 7.9 kb Endogenous Retrovirus Insertion in Intron 1 of CD36 with Obesity and Fat Measurements in Sheep

Abstract

Background

Endogenous retroviruses (ERVs) enhance genetic diversity in vertebrates, including sheep. This study investigates the role of Ov-ERV-R13-CD36 within CD36 gene and its association with phenotypic traits in sheep. Analyzing 58 sheep genomes revealed that ERVs constitute approximately 6.02% to 10.05% of the genomic content. We identified 31 retroviral insertion polymorphisms (RIPs) from 28 ERV groups. Among these, Ov-ERV-R13-CD36, which is specifically classified as a beta retrovirus, was selected for further analysis due to its location in CD36 gene, known for its role in fat metabolism, obesity (OB), body weight (BW), and body condition score (BCS). We assessed the association of Ov-ERV-R13-CD36 with OB and BCS across six sheep breeds, utilizing data from 1,355 individuals.

Results

Genomic analyses confirmed that Ov-ERV-R13-CD36 is located within CD36 gene on Chromosome 4, with polymorphisms across various sheep genomes. In a subset of 43 genomes, 22 contained the Ov-ERV-R13-CD36 insertion, while 21 exhibited wild-type variants. The studied animals showed variability in BCS and fat content associated with the Ov-ERV-R13-CD36 variant. Notably, Rahmani sheep exhibited a significantly higher BCS (4.62), categorized as obese, while Barki sheep displayed the lowest BCS (2.73), classified as thin to average. The association analysis indicated that sheep with the RIP−/− genotype correlated with higher OB and BCS, particularly in Rahmani and Romanov x Rahmani breeds.

Conclusions

Findings suggest that Ov-ERV-R13-CD36 within CD36 gene correlates with beneficial economic traits associated with OB and BCS, particularly in Rahmani and Romanov x Rahmani breeds. This indicates that Ov-ERV-R13-CD36 could be a valuable genetic marker for breeding programs aimed at enhancing traits like fat deposition and body condition in sheep.

Background

Endogenous retroviruses (ERVs) are a type of long terminal repeat (LTR) retrotransposons that are remnants of ancient retroviral infections in mammalian genomes, making up a significant proportion (8-10%) of these genomes [1,2,3,4,5]. Though often considered genomic parasites, retrotransposons, including ERVs, play crucial roles in genome organization, biological processes, species diversity and evolution [6].

Regions high in retrotransposon density, especially ERVs, are key to identifying genomic interactions and associations [7]. Due to their abundance and activity, they serve as molecular markers, aiding in genetic diversity, phylogenetic studies, and mapping [8]. Retrotransposon insertion polymorphism (RIP) markers have been applied in domestic animal studies like sheep, deer and chicken [9,10,11,12,13,14,15].

ERVs are remnants of ancient viral infections, embedded within mammalian genomes, that offer valuable insights into host-virus evolutionary dynamics. In sheep, various ERV families have been categorized, highlighting their genetic diversity and potential biological roles [16]. Notably, the study of endogenous-Jaagsiekte-sheep-retroviruses (enJSRVs) illustrates how genetic mutations can confer viral resistance, demonstrating the complex interplay between viral evolution and host defense mechanisms [17]. Beyond their biological significance, ERVs also play a critical role in tracing genomic evolution and economic traits across species. For instance, Chessa et al. [13] employed ERVs to trace the migration and distinguish between primitive and modern sheep breeds, shedding light on domestication processes. Similarly, Elleder et al. [14] identified CrERVγ in mule deer, revealing evolutionary links to economically significant livestock such as sheep and pigs.

Also, research has demonstrated the significant role of ERVs in the genetic evolution of different species, highlighting their regulatory functions and impacts on economically important traits [18,19,20]. Furthermore, studies have identified multiple ERV groups and families, suggesting that they continue to influence evolutionary processes and gene expression, thereby shaping traits of economic significance, particularly in mammals like ruminants and rodents [20, 21].

In a study on bovine genetics, research revealed that an LTR retrotransposon insertion in APOB gene affects cholesterol biosynthesis and lipid metabolism, impacting cattle health and productivity [4]. Another study focused on the ERVK [2-1-LTR] clade, identifying loci that influence ERV mobilization rates, which could alter genetic diversity and health traits in the herd [5]. Collectively, these studies underscore the critical role ERVs play in shaping genetic evolution and economic traits in various species.

ERVs can impact gene expression and genetic variation, potentially offering benefits within the host genome [22,23,24]. Despite beneficial aspects, ERV presence/absence (Pres/Abs) within candidate genes can alter economic traits, disease states and health indicators [2, 4, 5, 25,26,27,28].

For instance, Wang et al. [29] identified five RIPs within porcine TLR genes, including a 192 bp ERV insert in TLR6's first intron, enhancing TLR6 and TLR1 expression. This ERV insertion acts as an enhancer, significantly boosting expression of TLR-related genes, indicating its role in influencing the immune response in pigs. In this aspect, variations in genes like SUFU, SYCP2L and GLIPR1L1 have shown correlations with body measurements in goats, useful as molecular markers for body proportions [30]. Moreover, CD36 gene SNPs were tied to fat weight and skin yellowness, offering insights for chicken breeding [31].

Numerous studies have established CD36 gene as a candidate gene associated with traits such as obesity (OB), fat profile, fat deposition, body weight (BW), carcass characteristics, disease resistance and health indices in mammals [31,32,33,34]. CD36 gene, also known as fatty acid translocase (FAT), scavenger receptor class B member 3 (SCARB3) and glycoprotein 88 (GP88), is a significant 88-kD class B scavenger receptor glycoprotein found on the surface of platelets. It is expressed in various mammalian cell types, including platelets, monocytes, macrophages, dendritic cells and subsets of T and B cells [35,36,37,38]. As a receptor, CD36 binds multiple ligands, such as thrombospondin-1, collagen, oxidized phospholipids, oxidized low-density lipoprotein, long-chain fatty acids, erythrocytes parasitized by Plasmodium falciparum and apoptotic cells [39,40,41].

CD36 is part of a family of integral membrane proteins located in many body tissues, playing a crucial role in fatty acyl translocation. It functions as a multi-ligand cell surface receptor for oxidized LDL lipoproteins (ox-LDL), long-chain fatty acids, aged neutrophils, and Plasmodium falciparum-parasitized erythrocytes. These interactions are implicated in various diseases, including insulin resistance, diabetes, atherosclerosis and malaria [42,43,44,45,46,47,48,49,50,51]. Furthermore, CD36 is found on the outer mitochondrial membrane of skeletal muscle, where it facilitates long-chain fatty acid transport and regulates fatty acid oxidation within muscle mitochondria [52].

Our previous research on full-length ERVs in sheep indicated their polymorphism and potential to influence economic traits and disease resistance through interactions with candidate genes [19]. In this aspect, several studies confirmed that CD36 gene is considered as a candidate gene associated with OB/body condition score (BCS), fat profile, fat deposition, carcass traits, disease resistance and health indices in mammals [31,32,33,34].

CD36 gene has been extensively studied in humans. However, CD36 and its implications for production, health and economic traits in livestock have long been neglected by researchers [31,32,33,34, 53].

This study mainly examined the link between the Ov-ERV-R13-CD36 polymorphism and its effects on related phenotypes, including OB/BCS, meat composition and fat measurements, across different sheep breeds. The objectives were: a) to develop a protocol to mine full-length ERVs across 58 sheep genomes, aiming to identify RIP markers for future phylogenetic studies, genetic diversity and QTL mapping, b) to investigate correlations between ERVs located in genes linked to key economic traits in sheep, c) to explore the association between Ov-ERV-R13-CD36 located in CD36 gene, focusing on economic traits, d) to examine the differences in CD36 gene in terms of its location, type, number of transcripts, gene length, predicted amino acid sequences and exon and intron details in sheep and other eight animal species, e) to study the Pres/Abs of Ov-ERV-R13-CD36 within CD36 between reference and 42 non-reference sheep genomes.

Methods

The assembled genomes and gene annotation files employed in the investigation

By utilizing a total of 57 assembled non-reference sheep genomes in addition to the reference genome (Additional File 1, Table S1), our study conducted a genome-wide analysis of ERV-RIPs. These genomes were sourced from the National Centre for Biotechnology Information (NCBI) whole-genome sequencing (WGS) database, accessed on 6 January 2024. The gene annotation file was acquired from the NCBI database (https://ftp.ncbi.nlm.nih.gov/genomes/all/annotation_releases; accessed on 12 January 2024). Specifically, we downloaded a Bed format file containing gene features, which included information on the coordinates of long non-coding RNA (lncRNA) genes, protein-coding genes and other gene annotations.

Genome-wide ERV-RIP screening protocol in sheep

Screening for ERVs across the sheep genome on a genome-wide scale was outlined for mining. This protocol, utilizing 58 sheep genomes, delineated five primary procedural steps, according to Moawad et al. [19], Du et al. [54] and Chen et al. [55], with slight modifications, as follows;

Step 1: Identification of Full-length ERV Insertions

To initiate the process, a custom library containing full-length ERV sequences from our previous work [19], which comprises 28 groups/families (Table S2), was utilized to mask both the 57 non-reference genomes and the reference genome using RepeatMasker (-nolow, -lib custom library) [56]. Furthermore, these 28 groups/families, including the studied ERV-RIP (Ov-ERV-R13-CD36) were examined for similarity using the Repbase browser (https://www.girinst.org/repbase/update/browse.php).

Subsequently, ERVs encapsulating viral proteins such as Group-specific antigen (Gag), Polymerase (Pol) and Envelope protein (Env), flanked by identifiable LTRs and surpassing a length of 5 kb, were classified as full-length ERVs and retained for further scrutiny. The translation of these identified full-length ERVs was executed using GENSCAN (http://hollywood.mit.edu/GENSCAN.html; accessed on 26 January 2024). Following this, the translated sequences were explored against the Pfam database to analyze protein domains and functions [57]. The database utilized the hmmsearch command within the HMMER tool [58] for protein domain determination, enabling the identification of conserved protein families and aiding in the functional annotation of novel proteins across various species. The extraction of 500 bp flanking sequences, both upstream and downstream, surrounding these insertions was facilitated using the bedtools [59] commands flank and getfasta (Version 2.27.1), Additional File 2, Graph S1.

The evolutionary activity was assessed by estimating the insertion times of individual elements using the calcDivergenceFromAlign.pl tool within the RepeatMasker program [60,61,62]. This estimation utilized representative sequences for each element.

Step 2: Alignment with the Reference Genome

The flanking sequences bordering the identified full-length ERV insertions within the non-reference genomes were aligned with the reference genome using Blat [63] (-minIdentity = 90, -minScore = 450). Subsequently, the alignment outcomes were refined based on a length criterion ranging from 450 ~ 550 bp. Additionally, insertions with flanking sequences mapping to multiple genomic positions were omitted. In cases where insertions could not be aligned to the reference genome via the upstream 500 bp flanking sequence, the downstream 500 bp flanking sequences were similarly aligned, and the results from both sets were consolidated. This process facilitated the acquisition of each insertion's corresponding information within the reference genome from every non-reference genome.

Step 3: Detection of Full-length ERV Insertion Polymorphisms

The process of identifying putative full-length ERV insertion polymorphisms between the non-reference and reference genomes involved utilizing bedtools window (-w 50, -v) [59]. Differential insertions, which deviated from the reference genome, were categorized as potential full-length ERV insertion polymorphisms. More specifically, full-length ERV insertions from non-reference genomes that occupied a distinct window (comprising the ERV insertion site and a 400 bp flanking region) compared to the reference genome were regarded as putative full-length ERV insertion polymorphisms (Graph S1).

Step 4: Discovery of ERV Insertions through Data Mining and Validation of Polymorphisms via PCR

Initially, sequences containing the 500 bp flanking regions and the ERV sequences of potential full-length ERV insertions were extracted from all genomes using bedtools getfasta [59]. Subsequently, these sequences were consolidated, and any duplicates were eliminated using bedtools merge (-s, -d 10). Then, 28 ERV groups in sheep (Ovis aries) were discovered (Additional File 3, Data set 1). The conserved domains within these 28 ERV groups were investigated using the NCBI-Conserved-Domains online tool (https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi; accessed on 17 February 2024). After analyzing the masked 58 genomes, 31 full-length ERV-RIP insertions were identified (Data set 2).

For further analysis, 14 ERVs associated with relevant candidate genes (Table S3 and Data set 3) were selected for validation through PCR amplification, using the primers listed in Table S4. Subsequently, five full-length or active ERVs were selected based on their structural characteristics (Graph S2) and polymorphism (Graph S3).

Step 5: Selecting one of the ERV-RIPs for comprehensive exploration

One specific ERV-RIP, designated as Ov-ERV-R13-CD36, was selected for detailed investigation due to its unique association with the CD36 gene and the valuable insights obtained from data mining (Table S3).

Selection of Potential ERV-RIP Intersecting with Candidate Genes

During the comprehensive screening process for ERV-RIP markers, several ERV-RIPs were identified that intersect with candidate genes associated with economically significant traits in sheep. Among these, Ov-ERV-R13-CD36 was selected for further investigation due to its location in CD36 gene.

CD36 is well-documented for its influence on traits such as BCS, BW and fat distribution within the body [31, 39]. As such, the intersection of Ov-ERV-R13-CD36 with CD36 gene presents a promising avenue for understanding the genetic mechanisms underlying these important traits. Identifying and analyzing this interaction could provide valuable insights into breeding strategies aimed at improving these economic traits in sheep [64].

Investigation of CD36 Gene in Sheep and Eight Other Species

CD36 gene was analyzed across the reference genomes of nine species: sheep (Ovis aries), goats (Capra hircus), domestic cattle (Bos taurus), water buffalo (Bubalus bubalis), pigs (Sus scrofa), rabbits (Oryctolagus cuniculus), chicken (Gallus gallus), zebrafish (Danio rerio) and domestic ferrets (Mustela putorius furo). This part of the investigation aimed to determine the gene's location, type, number of transcripts, gene length, predicted amino acid sequences, as well as details on exons and introns. Reference genomes were utilized for each species (Table S5). Subsequently, further research focused on the genomic regions surrounding CD36 gene, including neighbouring genes and exon organization, as detailed in Table S5.

Analyzing Ov-ERV-R13-CD36 Intersection Involving CD36 Gene in the Sheep Reference Genome

The intersect program within bedtools version 2.27.1 [59], was employed to investigate the distribution bias of full-length ERV within sheep genomes and their overlaps with host genes. Through the bedtools intersect functionality, the examination of overlaps between two sets of genomic features was facilitated, offering precise control over the reporting of intersections. Host genes were sourced from the sheep annotation files retrieved from the NCBI database and then compared with the results obtained from RepeatMasker version 4.0.9 [56]. RepeatMasker was employed to annotate all ERVs in the sheep genomes using a custom library of previously identified full-length ERVs.

CD36 gene information, including its exons and introns, was obtained from the NCBI database (https://www.ncbi.nlm.nih.gov/gene/?term=LOC101115115), Ensembl (https://www.ensembl.org/Sheep/Search/Results?q=LOC101115115;site=ensembl;facet_species=Sheep), and UCSC Genome Browser (https://genome.ucsc.edu). These data were organized for figure preparation using the "IBS" (v1.0.3) Illustrator of Biological-Sequences Software, available at https://www.nuget.org/packages/IBS.Data/1.0.3. Predictions of the structures for CD36 gene in a reference genome (GCA_016772045.2) were made using Geneious Prime Software (Augustus Tool, Version 7_ 2024), available at https://www.geneious.com. While, the domains for Ov-ERV-R13-CD36 and CD36 gene were predicted using the NCBI Conserved domains tool (https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi) according to Wang et al. [65], Lu et al. [66] and Marchler-Bauer et al. [67]. Additionally, the amino acid sequences for CD36 gene and Ov-ERV-R13-CD36 were predicted utilizing GENSCAN (http://hollywood.mit.edu/GENSCAN.html).

Exploring Differentiation of Ov-ERV-R13-CD36 within CD36 Gene Among Reference Genome and 42 Non-reference Genomes in Sheep

In this phase of the study, 43 sheep genomes were collected, including one reference genome and 42 non-reference genomes, as detailed in Table S6. These genomes were sourced from the NCBI database (https://ftp.ncbi.nlm.nih.gov/genomes; accessed on 12 January 2024). The investigation focused on identifying Pres/Abs in Ov-ERV-R13-CD36 located within CD36 gene across these genomes. Subsequently, the genomes containing Ov-ERV-R13-CD36 insertion were analyzed for differences in long terminal repeats (LTRs), ERV length and domain structures using Fast-PCR software version 6.8 [68] and the NCBI Conserved Domains tool.

Test Ov-ERV-R13-CD36-Pres/Abs in six sheep breeds

The methodology adopted in this part encompasses a comprehensive analysis of six distinct international sheep breeds, aimed at assessing traits such as OB/BCS, meat chemical composition, fat content and genetic polymorphism. The breeds examined in this study included Barki, Rahmani, Rahmani x Barki crossbred, Awassi, Ossimi and the Romanov x Rahmani crossbred. Barki sheep, primarily found in Egypt, are raised for meat production and are well-adapted to arid environments. Rahmani sheep, native to the Nile Delta region, serve dual purposes, providing both meat and wool. Rahmani x Barki crossbred combines traits from both breeds to enhance meat production and environmental adaptability. Awassi sheep, common in the Middle East, are mainly bred for milk production but also provide meat and wool. Ossimi, also found in Egypt, are primarily raised for meat and wool with good climate adaptability. Romanov sheep, originating in Russia, are known for their prolificacy and are typically used to improve reproductive traits and lamb production. Romanov x Rahmani crossbreed aims to combine the prolificacy of Romanov sheep with the adaptability of Rahmani sheep, suited for enhanced meat production [69, 70]. Through a structured experimental design and rigorous ethical protocols, data collection encompassed physical and chemical assessments, blood sampling for DNA isolation, and PCR verification to investigate genetic variations. This multi-faceted approach provides a robust framework for understanding the phenotypic and genotypic diversity within these sheep breeds, thereby contributing valuable insights into their suitability and performance in various agricultural and breeding applications.

Animals

The present study is part of a large project aimed at evaluating the production traits of six sheep breeds, encompassing a total of 1,355 animals, specifically; Barki (B, n=366), Rahmani (R, n= 198), Rahmani x Barki crossbred (RB, n= 426), Awassi (A, n= 143), Ossimi (O, n= 112) and Romanov x Rahmani (V, n= 110) as illustrated in Additional File 4, Fig. S1. The parents of these animals were sourced from five distinct geographical regions in the northern part of Egypt, namely, Alexandria governorate (GPS: 31.206208, 29.919704), Matrouh Province, El-Hammam district (GPS: 30.833132, 29.397580), Matrouh Province (GPS:31.353910, 27.235560), Salloum plateau city (GPS: 31.573626, 25.15922) and Sakha city (GPS: 31.090560, 30.943543). Although initially raised in the experimental station in northern Egypt, the experimental animals belonging to the six sheep breeds were adult individuals (average age of 36.50±0.75 months) randomly selected from respective contemporary groups born within a timeframe of 1~3 weeks. These animals were all kept under a semi-intensive system, allowing for free-range activities within spacious concrete barns and maintained under uniform conditions of feeding, climate and overall management.

Phenotypic traits

Obesity (OB)/Body Condition Score (BCS)

Three experienced judges independently assessed the body conditions of the studied sheep breeds using OB/BCS, which applied a scoring scale ranging from 1-5 with increments of 0.5 points, as established by Morand-Fehr and Hervieu [71] and Russel et al. [72]. The median BCS recorded was 3, within a range of 1.50 - 5.00. In cases of discrepancies among the judges, a consensus decision was reached to determine the final score. The assessments were conducted at three specific sites: thoracic vertebrae or sterna (BCSa), rib cage (BCSb) and lumbar vertebrae behind the ribs/loin eye (BCSc) [73, 74], as illustrated in Fig. S2A & B. Detailed information on the visual body condition scoring system is provided in Table S7. OB/BCS evaluation was conducted on 1,355 individuals.

Meat chemical composition, physical assessment and fat measurements

Breeds preparation

Six sheep breeds/genotypes (a total n=36); Barki (B; n=3♂ and 3♀), Rahmani (R, n=3♂ and 3), Rahmani x Barki crossbred (RB, n=3♂ and 3♀), Ossimi (O, n=3♂ and 3♀), Awassi (A, n=3♂ and 3♀), Romanov x Rahmani (V, n=3♂ and 3♀), were used to determine the meat chemical composition, assess physical parameters and measure the fat content. These individuals were slaughtered when they reached an age of 38 months.

Slaughtering

The animals were weighed, subjected to an 18 hrs fasting period with ad libitum access to water, and then weighed again just before slaughter. Subsequently, the animals were slaughtered and processed following standard commercial procedures. All experimental protocols and procedures were executed in compliance with the guidelines outlined in the Guide for the Use and Care-of-Agricultural-Animals in Research and Teaching by the Federation-of-Animal-Science-Societies (FASS, 2010) and the Guide for the Care and Use of Agricultural Animals in Research and Teaching, 3rd ed (https://www.aaalac.org/about/Ag_Guide_3rd_ed.pdf). Moreover, all animal experiments were ethically approved by the ethics committee of Alexandria University, Faculty of Agriculture (Al-Shatby), Egypt (No. AU082209203103) and ethics committee of Yangzhou University, College of Animal Science and Technology, China (No. 202403005).

Measuring physical assessment, meat composition and fat measurements

Post-Slaughter Processing

After slaughtering and bleeding, the head was removed at the atlanto-occipital joint, and the feet were detached at the carpal and tarsal joints. Initially, the carcasses were partially skinned while lying on their backs before being suspended by the hind legs for complete skinning. Weights for both carcass and non-carcass components were recorded immediately. Non-carcass components included the lungs, trachea, heart (collectively referred to as the pluck), head, skin, feet, digestive tract, liver, spleen and pancreas. The weight of the digestive contents was calculated by subtracting the weight of the empty digestive tract from that of the full digestive tract, and this was used to determine the empty live weight (ELW).

Physical assessment

Specific fat deposits such as kidney, omental, pelvic and mesenteric fats were isolated and weighed. Physical assessments, including fat, muscle and bone percentages were conducted following methodologies described by Sen et al. [75] and Santos et al. [76]. Carcasses were divided into sections, and the loin eye area was recorded. The trimmed meat, bone and dissected fat were weighed separately, which included initial measurements of bone (kg and %), trimmed meat (kg and %) and dissected fat (kg and %). Ratios such as lean:fat, lean:bone, and carcass:fasted weight (%) were also calculated to provide deeper insights into carcass composition. Composition analysis of chilled cuts was performed manually, calculating the distribution of lean, fat and bone based on the chilled carcass weight.

Fat measurements

For detailed fat measurements, the procedures outlined by Ekiz et al. [77] and Kirton et al. [78] were applied. Weights for specific fat deposits, including heart fat, kidney fat, gut fat, fat tail and gastrointestinal (GI) tract fat, were individually measured. This helped ascertain total fat stores and total body fat, which were calculated by summing all fat depot weights, providing a comprehensive assessment of fat distribution.

Determining the Chemical Composition of Meat

The chemical composition, including moisture, crude protein and ash, was determined according to AOAC, [79] and Madruga et al. [80] on the rib-eye area (longissimus dorsi) (Fig. S3A & B). Specifically, 100 g samples were taken from the left rib-eye area of each lamb. These samples were carefully trimmed to remove connective tissue and external fat. The samples were then freeze-dried, ground to pass through a 1 mm sieve, and stored for later analysis. Ash content was measured by combusting the meat samples at 600 °C for 8 hrs. The nitrogen content was assessed using the Kjeldahl method, with protein content calculated by multiplying the nitrogen value by 6.25. Fat extraction was performed using the Soxhlet apparatus.

Blood sampling, DNA isolation, manipulation and PCR verification

Blood sampling

A venous blood sample of 5 ml was individually collected from each of the 144 randomly selected animals out of a total of 1,355. The blood samples were drawn from the jugular vein using venojects, treated with 0.5 ml of 2.7% EDTA (Pspark, U.K) as an anticoagulant, and immediately transferred in an ice box to the laboratory for further processing.

DNA isolation and processing

Genomic DNA was extracted from individual blood samples collected from randomly chosen males and females sheep of the experimental Barki (B; n=12♂ and 12♀), Rahmani (R, n=12♂ and 12), Rahmani x Barki crossbred (RB, n=12♂ and 12♀), Ossimi (O, n=12♂ and 12♀), Awassi (A, n=12♂ and 12♀) and Romanov x Rahmani (V, n=12♂ and 12♀) groups using a DNA isolation kit (Tiangen Biotech, Beijing, China). The isolated DNA samples were subjected to electrophoresis on a 1.2% agarose gel in 0.5x TBE buffer, as described by Sambrook et al. [81], supplemented with 0.5 μg/ml ethidium bromide. Electrophoresis was conducted using an apparatus with a power supply and visualization was carried out under an ultraviolet transilluminator equipped with a Gel documentation system (Chemi.DocTM XRS+ with Image LabTM Software, BIO-RAD, USA). The purity and integrity of the DNA were assessed using a NanoDrop Spectrophotometer (2000/2000c, Thermo Fisher ScientificTM, USA), with the optical density (OD) ratio monitored at 260/280 nm. DNA samples with a high-quality rating, displaying an OD ratio between 1.80 and 2.00 (average 1.82), were utilized for subsequent analyses, while samples of insufficient quality were re-extracted. The average measurements for concentration, volume and mass were determined as 70.47 ng/μL, 29.5 μL and 2.14 ng, respectively.

PCR verification on narrow scale

Six sheep breeds; Barki, Rahmani, RB, Ossimi, Awassi, and Romanov x Rahmani were chosen for PCR validation of full-length ERV polymorphisms. DNA samples were extracted from blood samples as described previously. Subsequently, ten individual samples from each breed were combined for initial PCR analysis to identify the expected ERV insertion polymorphisms within CD36 gene. Given the typical length of full-length ERVs exceeding 8 kb, the careful design of double PCR primer pairs was implemented for genotype verification. The primer design strategy is elucidated in Fig. S4A. Primer sets (1 and 2) and (1 and 4) for Ov-ERV-R13-CD36 were precisely crafted utilizing Oligo-7 software (V.7/2024) (https://www.oligo.net). The primers were synthesized by Vazyme Biotech Co., Ltd, located in Nanjing, China. Subsequently, the following primers were used: P1_13_F: 5'-AATGGTAAGCTCCCAAACTCA-3', P2_13_R: 5'-ACCAAGGGCAAACTTCCTCGATG-3', and P3_13_R: 5'-TTCACAAGCACTGACGGAT-3'.The amplification procedure used Green SuperMix (TaKaRa, Japan), with 10 pM of primers and 100 ng of genomic DNA per sample. The PCR protocol consisted of 40 cycles with denaturation at 95°C for 1 min, annealing at 58-60°C for 1 min, extension at 72°C for 1 min, followed by a final extension step at 72°C for 2 min. The amplification was performed using a T100 Thermal Cycler from BIO RAD, Singapore. Subsequently, 7-10 µL of PCR products and 4 µL of DL5000 molecular weight markers were subjected to electrophoresis on 1.0 % agarose gels in 1× TAE buffer at a constant voltage of 130 V for 30~45 min. The gels were stained with ethidium bromide and visualized under ultraviolet fluorescence. PCR amplification results showing bands within specific size ranges can determine the presence or absence of homozygous ERV insertion. A band of 1,056 bp with primers 1 and 2 indicates the ERV+/+ insertion genotype, while the absence of a band of 1,159 bp with primers 3 and 4 confirms this genotype. If a band of 1,159 bp is present without the 1,056 bp band, it suggests the ERV-/- genotype. Observing both shorter and longer bands signifies a heterozygous RIP genotype (ERV+/−), as shown in Fig. S4B. This genotyping approach was based on the methodology described by Du et al. [54]'s framework.

PCR verification and Genotyping Investigation for Ov-ERV-R13-CD36 on a Large Scale

This part of the genotyping process involved testing 24 randomly selected individuals from each breed, resulting in a total sample size of 144 animals to assess Ov-ERV-R13-CD36.

Statistical analysis

Analysis of variance

The study collected and analyzed data on OB/BCS, meat chemical composition, physical assessment, and fat measurements of the observed breeds. The normality of the data was assessed using the Shapiro-Wilk test (SAS, 2009), which indicated that all data followed a normal distribution (Shapiro-Wilk test (W) ≥ 0.90). The effects of Ov-ERV-R13-CD36 on the target traits were evaluated using the GLM procedure in SAS, based on the model:

$${\mathrm Y}_{\mathrm{ijkl}}=\mathrm\mu+{\mathrm G}_{\mathrm i}+{\mathrm B}_{\mathrm j}+{\mathrm S}_{\mathrm k}+{\mathrm e}_{\mathrm{ijkl}}$$

Where: µ is the overall mean, Gi signifies the effect of the ith genotype for Ov-ERV-R13-CD36 (+/+, −/− and +/−), Bj represents the fixed effect of the jth breed, Sk indicates the fixed effect of the kth sex and eijkl denotes the residual error. Variations between means within each weight category were analyzed using the least significant difference (LSD0.05).

Genetic Indices and Equilibrium Analysis

The chi-squared test (χ2) was employed to assess whether the six populations conformed to Hardy-Weinberg equilibrium (HWE). HWE analysis was conducted in the context of genetic association studies. Additionally, heterozygosity (Ho), expected heterozygosity (He), effective allele number (Ne) and polymorphic information content (PIC) were determined using Nei's methodology [82, 83].

Results

Analysing the Ov-ERV-R13-CD36 Intersection Involving CD36 Gene in Sheep Genomes

Previous ERV annotation and full-length ERV insertion prediction [19] side by side with the current investigation, revealed that, in the reference genome, there were originally 218 full-length ERVs identified (Table S8). Using our specific methodology and criteria, which are thoroughly explained in the methods section, we identified 28 distinct groups (Table S2) that are distributed across 31 ERV-RIPs (Data set 2) in 58 sheep genomes. Upon further analysis through PCR testing and data mining, we discovered that 14 out of these 31 ERV-RIPs intersect with candidate genes (Table S3 and Data set 3) and exhibit polymorphic characteristics (Graph S3). Finally, one of these 14 ERVs, which is particularly associated with the traits of interest in our study, was selected for more detailed examination. The full-length ERV insertion, designated Ov-ERV-R13-CD36, located on chromosome 4, was selected due to its location within the significant candidate gene CD36 (Table S3 and Fig. 1). Ov-ERV-R13-CD36, belonging to the Cap_ERV_24 family/group, is classified as a Class II (Beta retrovirus). By investigating Repbase data, we found that Cap_ERV_24/Ov-ERV-R13-CD36 is novel.

Fig. 1
figure 1

A Genetic map of Chromosome 4 in sheep (Ovis aries) indicating specific positions and directions for genes, including Cluster of Differentiation 36 (CD36) gene. B Genomic localization and functional implications of the CD36 gene with Endogenous Retrovirus (ERV) insertion (Ov-ERV-R13-CD36) in sheep. C The structure organization of the Ov-ERV-R13-CD36 is as follows: [Gag_p10: Retroviral GAG p10 protein; - Gag_p24: GAG gene protein p24 (core nucleocapsid protein); - RT_RNaseH: Reverse transcriptase RNaseH domain; -RNaseH: Endonuclease domain of reverse transcriptase;- zf-CCHC_5: GAG polyprotein viral zinc-finger; - trimeric_dUTPase: Trimeric dUTP diphosphatase; - RVT_thumb: Reverse transcriptase thumb domain; - Integrase_Zn: Integrase Zinc binding domain; - rve: Integrase core domain;- PBS: Primer binding site; - PPT: Polypurine tract[. D PCR Verification for the selected Ov-ERV-R13-CD36. 10 pooled samples per breed (Total; n=60); 1: Barki, 2: Rahmani, 3: Rahmani x Barki crossbred, 4: Awassi, 5: Ossimi, 6: Romanov x Rahmani. M: DNA Ladder 5 kbp.

Based on the multiple sequence alignment with available genomes, a 7,930 bp full-length ERV insertion on Chromosome 4 within CD36 gene was revealed (Fig. 1A and Data set 4). Further genomic annotation identified that Ov-ERV-R13-CD36 is located in intron 1 of the mutant CD36 gene based on the reference genome (GCA_016772045.2, NC_056057.1:41861053-41968317), specifically within Transcript NCBI: XM_042248492.2, comprising 15 exons and 14 introns (Fig. 1B). Genomic localization and domain analysis of Ov-ERV-R13-CD36 indicated that each LTR region is approximately 445 bp, with domains covering Group-Specific Antigen (Gag; 531 aa), Protease (Pro; 278 aa), Polymerase (Pol; 696 aa), and Envelope (Env; 441 aa) (Fig. 1C). Subsequently, the Ov-ERV-R13-CD36 on chromosome 4 was subjected to initial PCR verification by using 10 pooled samples per breed. PCR amplification performed on the examined sheep breeds resulted in the production of a distinctive short fragment of 1,056 bp and a longer fragment of 1,158 bp (Fig. 1D), indicating the occurrence of a Pres/Abs polymorphism of full-length ERV in the studied sheep breeds. Additionally, domains for CD36 and Ov-ERV-R13-CD36 are presented in Additional File 5.

Comparative genomic analysis revealed that CD36 in several animal genomes (i.e., sheep, goats, cattle, buffaloes, pigs, rabbit, chicken, zebrafish and domestic ferret) has a large first intron, with several nearby genes, including the well-annotated GNAT3 gene (Fig. 2 and Fig. S5). Fig. 2 provides a comprehensive overview of the structural similarities and differences surrounding CD36 across diverse species, which may help elucidate how ERV insertions in this region could impact gene regulation and contribute to phenotypic variation. Furthermore, by identifying conserved genomic features, we can better understand the evolutionary pressures shaping CD36 gene, which may have implications for its functional roles in various traits across different species. Worth mentioning, Ov-ERV-R13-CD36 was identified in the reference genome of sheep within intron 1 of CD36 gene, as shown in Fig. 2A, and is absent from the genomes of other species (Fig. 2B-I). Table S9 highlights significant differences in CD36 gene across various species in terms of gene length, predicted amino acid sequences and transcriptional complexity. Notably, the sheep CD36 gene spans 107,265 bp and has the highest number of transcripts (35), reflecting greater transcriptional complexity than goats and domestic ferrets, which have only 6 and 4 transcripts, respectively. The pig exhibits the largest predicted amino acid sequence, consisting of 1,655 amino acids, contrasting sharply with the shorter sequences found in zebrafish and chicken, each with only 382 amino acids.

Fig. 2
figure 2

Analysis of CD36 gene and its exon organization in the reference genomes of sheep (A), goats (B), cattle (C), buffaloes (D), pigs (E), rabbit (F), chicken (G), zebrafish (H) and domestic ferret (I).

In addition, alignment of the genomic regions of Ov-ERV-R13-CD36 across 43 sheep genomes revealed that 22 genomes contain ERV insertions (mutant type), while 21 genomes exhibit an absence of ERV (wild type), as outlined in Table S10. The aligned sequences for Ov-ERV-R13-CD36 in these 43 genomes are presented in Additional File 6. Furthermore, 22 genomes with the Ov-ERV-R13-CD36 insertion were analyzed for variations in LTRs, ERV length, and domain structures, revealing notable variability in the lengths of the entire ERV, LTR and specific protein-coding domains, as shown in Table 1. The aligned sequences for the Ov-ERV-R13-CD36 insertion within CD36 gene in the 22 selected genomes are provided in Additional File 7. The total lengths of ERV insertion were 7,930 bp for 6 genomes, and 7,943 bp for 15 genomes and 7,756 bp for one genome. The length differences in the ERV domains (Gag, Pol and Env) likely result from accumulated mutations. The Gag domain typically ranges from nucleotide (nt) sequences coding for 531 to 616 amino acids (aa), with outliers at nt sequences coding for 193 and 173 aa in genomes such as GCA_002742125.1 and GCA_022416785.1, respectively. The Pro domain generally measures from nt sequences coding for either 278 or 289 aa, though shorter sequences are present in some genomes. The Pol domain is mostly represented by nt sequences coding for 777 aa long, although some, such as GCA_016772045.2, exhibit a shorter length of 696 aa. The Env domain ranges widely from nt sequences coding for 441 to 618 aa, but it is reduced to a length of 328 aa in GCA_002742125.1.

Table 1 Analysis of the Ov-ERV-R13-CD36 insertion in CD36 gene in the reference genome and 21 selected non-reference sheep genomes

Population genetic analysis of Ov-ERV-R13-CD36 by PCR

Furthermore, to validate the findings on a population level across a broader spectrum, a total of 144 individuals were randomly selected from six sheep breeds, with 24 individuals from each breed, comprising both male and female populations. This selection aimed to assess the occurrence of Pres/Abs in the genomes of the studied sheep breeds for this Ov-ERV-R13-CD36. This RIP is polymorphic in diverse breeds, and the representative PCR detection results are illustrated in Fig. 3. The analysis of Ov-ERV-R13-CD36 genotyping in the studied sheep breeds revealed notable variations. Among the examined breeds, Barki displayed a diverse distribution with RIP+/+ at 16.67%, RIP−/− at 29.17% and RIP+/− at 54.16%. Rahmani and Romanov x Rahmani exhibited a distinct pattern with 100% RIP−/− genotype, RB took the second place for RIP−/− genotype at 83.33 after Rahmani and Romanov x Rahmani, while Awassi showed 100% RIP+/− (Table 2 and Fig. S6). These results underscore the genetic diversity present among the sheep populations, as reflected in Ho, PIC and Ne, emphasizing the genetic complexity within these sheep breeds (Table S11).

Fig. 3
figure 3

PCR Verification for the selected Ov-ERV-R13-CD36. 24 individuals (n=12♂ and 12♀) per breed, (Totally n=144); RxB: Rahmani x Barki crossbred and RxR: Romanov x Rahmani. M: DNA Ladder 5 kbp.

Table 2 Ov-ERV-R13-CD36 genotypes and their association with Obesity (OB)/ Body Condition Score (BCS) in the studied sheep breeds.

The observed absence in Ov-ERV-R13-CD36 could potentially impact CD36 gene`s function in regulating OB/BCS and fat storage within the body. Absence in critical regions of CD36 gene (Figs. 1 & 3) may disrupt its normal physiological activities related to fat metabolism and storage, especially for Rahmani, Romanov x Rahmani and RB breeds (Tables 2 and 3). Understanding the implications of these absences on CD36 function is essential for elucidating their potential role in obesity development and fat accumulation processes.

Genotypes of Ov-ERV-R13-CD36 and their association with economic traits in sheep

The current study identified potential associations between Ov-ERV-R13-CD36 genotypes and both BCS and OB measurements (Table 2), as well as fat measurements (Table 3) in the studied sheep breeds. The analysis of Ov-ERV-R13-CD36 genotypes across various breeds revealed noteworthy correlations with OB/BCS. Among the studied breeds, Rahmani and Romanov x Rahmani exhibited the highest OB/BCS (p < 0.001), both displaying a 100% RIP−/− genotype. In contrast, Barki breed had the highest frequency of the RIP+/+ genotype (16.67%) and the lowest OB/BCS (p < 0.001), indicating a range in body condition from thin to average. Also, RB breed demonstrated a significant presence of the RIP−/− genotype, accounting for 83.33% of the population, and achieved the second-highest OB/BCS after Rahmani and Romanov x Rahmani breeds (Table 2).

Table 3 The physical assessment, chemical composition and fat measurements of the studied sheep breeds

Additionally, the analysis of dissected fat in various sheep breeds uncovered potential genetic associations with Ov-ERV-R13-CD36 genotypes. Rahmani sheep, with a 100% RIP−/− genotype, exhibited the highest dissected fat weight (3.682 ± 0.13 kg, p = 0.002), suggesting a predisposition to fat accumulation. Conversely, Barki sheep had the lowest fat weight (2.257 ± 0.181 kg), indicating leaner body composition. Also, Romanov x Rahmani sheep showed relatively high fat weight (2.876 ± 0.18 kg, p = 0.002), closely followed by BR breed, which exhibited a fat weight of 2.837 ± 0.231 kg. These findings may emphasize the complex relationship between genotype and fat deposition in sheep (Table 3).

The variations in fat levels and genetics among Rahmani, Romanov x Rahmani, and BR sheep breeds may correspond to their Ov-ERV-R13-CD36 genotypes. The introduction of hyper-breed genetics may enhance total fat content in RB crosses [70], with the exception of Rahmani and Romanov x Rahmani breeds. This potential enhancement is likely associated with an increase in the frequency of ERV−/− individuals, which may positively influence genetic traits related to fat content.

Discussion

ERVs exhibit a significant antisense bias in gene introns, indicating strong selection against elements aligned with gene transcription. This suggests that such biases might influence economic traits by potentially affecting gene regulation [53, 84, 85]. The findings from Klymiuk et al. [16] and Cumer et al. [17] provide an overview of the ERVs present in the ovine genome and their evolutionary implications. Klymiuk et al. [16] identified multiple OERV families and discovered novel open-reading frames, suggesting that these sequences could have active biological roles or implications in sheep biology. On the other hand, Cumer et al. [17] challenged previous notions regarding the domestication-associated fixation of protective enJSRV mutations by proposing an alternative model based on their evolutionary timeline. This model posits that such protective mutations, which arose after key evolutionary divergences in the Ovis lineage, may have provided a selective advantage due to natural selection, rather than domestication alone.

In this aspect, the integration of ERVs can have a significant impact on gene activity and ultimately influence phenotype by modulating regulatory elements such as introns, exons, promoters and enhancers of genes [28, 86,87,88,89,90]. Phenotypic variations resulting from ERV-Pres/Abs have been documented in diverse domesticated species, including dogs [91], cats [92], sheep[19] and pigs [54]. For instance, a study on pigs found a correlation between ERV-D14-RIP within STAB2L target gene and traits related to BW and growth. This correlation affects variations in BW and growth rate, particularly in Large White pigs [54]. The identification of polymorphisms and differentiations arising from ERV-Pres/Abs in domesticated animals may significantly contribute to revealing underlying structural variations responsible for phenotypic diversity and could potentially impact breeding practices [10].

In the present study, the analysis of ERV distribution in 58 sheep genomes, including one reference genome (ARS-UI_Ramb_v2.0), revealed varying genomic coverage of ERVs, ranging from approximately 6.02% to 10.05% (Table S8). Using a specialized dataset of 28 ERV groups (Data set 1 and Table S2) across 58 sheep genomes, bioinformatic analysis identified 31 ERV-RIPs (Data set 2). Table S2 provides a detailed classification of the analyzed ERV families and highlights their relevance to Ov-ERV-R13-CD36. Our investigation of Repbase data revealed that 22 out of the 28 families, including the tested family (Cap_ERV_24/Ov-ERV-R13-CD36), are novel, while the remaining 6 have been previously identified, including Cap_ERV_1 (OviAri-5.324_LTR), Cap_ERV_4 (OviAri-5.2557_int), Cap_ERV_10 (OviAri-1.272_LTR), Cap_ERV_12 (OviAri-6.2056), Cap_ERV_13 (OviAri-1.306), and Cap_ERV_14 (OviAri-3.284_LTR). Further validation revealed that out of the 31 full-length ERV-RIPs, 14 ERVs (45.16%) were located within candidate genes (Table S3) and subsequently confirmed through PCR detection (Data set 3 and Graph 3). Moreover, the selected ERV insertion fragments analyzed were over 7 kb in length, indicating a predominance of truncated copies within the sheep genome, specifically the selected RIP named Ov-ERV-R13-CD36 spanning approximately 7.93 kb. In this regard, Du et al. [54] reported that the presence of full-length ERV insertion polymorphisms, characterized by substantial structural changes exceeding 5 kb, pronounced genetic influence. This suggests a potential impact for some ERVs on gene expression and function due to significant overlap with coding and non-coding regions [85, 93,94,95,96,97].

In-depth investigation of the selected Ov-ERV-R13-CD36 revealed significant structural variations in CD36 gene, attributed to the absence of ERV insertions, particularly in Rahmani and Romanov x Rahmani breeds (Fig. 1D). PCR verification on a large scale further validated the occurrence of Pres/Abs in the genomes of the studied sheep breeds for this specific Ov-ERV-R13-CD36 (Fig. 3). The 100% Ov-ERV-R13-CD36 −/− genotype (absence /wild type) observed in Rahmani and Romanov x Rahmani breeds, verified using PCR, aligns with our investigation of 43 sheep genomes (Additional File 6). Among these, 21 were found to have absences (Table S10) across several international sheep breeds: Rambouillet, East Friesian, Polled Dorset, Waggir, Charollais, Texel, Romanov, White Dorper, Hu, Yunnan, Kermani, Ujimqin, Romney, Suffolk and Charollais.

The associations between Ov-ERV-R13-CD36 genotypes and economic traits, including OB/BCS (Table 2), and meat chemical composition and fat measurements (Table 3), displayed notable variations among different sheep breeds. Rahmani and Romanov x Rahmani breeds, with a unique 100% RIP−/− genotype, followed by RB breed (83.33%) demonstrated the highest mean for OB/BCS and fat content among all the studied breeds. These findings imply that these particular breeds may have lost the protective effect of CD36 gene against OB/BCS due to the absence of Ov-ERV-R13-CD36 within CD36 gene. In contrast, Barki breed, with the highest distribution of the RIP+/+ genotype (16.67%) among different breeds, and comprising 54.16% RIP−/+ and 29.17% RIP−/− genotypes, displayed the lowest OB/BCS and mean dissected fat weight. This finding indicates that Barki breed may have a protective effect due to the function of CD36 gene against OB/BCS through insertion. The varying distribution of RIP genotypes within the studied breed highlights the genetic diversity present in this population and its potential impact on fat metabolism.

The current results align closely with Bokor et al. [98] research findings in humans indicating that SNPs in CD36 were significantly associated with higher body mass index and body fat. Also, the current result agrees with Clop et al. [99] who identified three CD36 variants associated with growth and fat deposition by analyzing the genome of pigs. In this aspect, Zhao et al. [31] found 11 SNPs in CD36 gene in chicken, with specific mutations detected at different locations; 2 SNPs in the 5′ flanking regions, 8 SNPs in the intron region and 1 SNP in the exon region, which belonged to a synonymous mutation. These mutations were associated with specific traits, such as abdominal fat weight, full-bore weight rate and skin yellowness. Haplotypes of these SNPs were correlated with heart weight, stomach weight, wing weight, leg skin yellowness and shin skin yellowness before slaughter. In birds, the SNP (g. 476593 T > C) of CD36 gene was significantly associated with total cholesterol and lipoprotein cholesterol levels, the result demonstrated that CD36 might be an important genetic marker for the selection of lipid metabolism and meat quality traits in ducks [100]. Additionally, Rać et al. [101] reported that CD36 expression in skeletal muscles increases in response to elevated levels of triacyl glycerides (TG) and fatty acids (FA) in the plasma. This regulatory mechanism is influenced by the energy demands of the tissue.

CD36 not only transports fatty acids into the cell for lipid synthesis and metabolism but also affects cholesterol uptake and inflammatory responses [102, 103]. In CD36 knock-down mice, CD36 deficiency suppressed fat deposition in the viscera, subcutis and gonads caused by high-fat feeding and decreased BW [104]. Researchers injected CD36 knockout precursor adipocytes into the back of mice and then fed them at high-fat levels, then found that the weight of new fat and the size and number of adipocytes in the experimental group were lower than those in the control group compared to the mice injected with no knockout cells [105]. These results suggest that CD36 plays a prominent role in the differentiation and development of adipose tissue [106]. Long-chain fatty acids (LCFAs) are an important substrate for ATP production within the skeletal muscle. As a fatty acid transporter protein, CD36 locates in skeletal muscle cell membranes and mitochondrial membranes, and plays a major role in the uptake and transport of LCFA, regulating energy sources and lipid metabolism in skeletal muscle [107]. Also, it serves as a ligand-receptor of thrombospondin, long-chain fatty acids, oxidized low-density lipoproteins (LDLs), fatty acid profile and oxidized lipid uptake [51].

The current findings align with the research conducted by Fagerberg et al. [108], which revealed a ubiquitous expression of CD36 in various organs, with the most pronounced expression observed in fat tissues. The authors proposed that CD36 protein might function intracellularly to influence adipose tissues and organ size, thereby regulating adipogenesis and other related processes. Furthermore, CD36-RNA is found in several tissues. The researchers confirmed that CD36 gene expression was positively correlated with other genes influencing obesity, BCS, fat measurements and body size, as determined through quantitative transcriptomics analysis (such as RNA-Seq) which measured normalized mRNA abundance and verified using statistical correlation methods like Spearman correlation coefficient. The current findings suggest that the homozygous Ov-ERV-R13-CD36-/- variant in the Rahmani, Romanov x Rahmani, and RB breeds may be associated with a larger body size and extreme OB/BCS.

Therefore, exploring the molecular mechanisms through which CD36 affects OB/BCS or fat measurements in various domesticated animals, notably sheep, could offer valuable insights into potential therapeutic targets for OB and metabolic disorders, especially with the newly generated and organized data concerning CD36 gene across various species.

While this study provides valuable insights into the role of Ov-ERV-R13-CD36 in economic traits associated with CD36 in sheep, there are limitations that warrant consideration. Future research should explore how environmental factors may influence the expression and effects of ERV insertions. Additionally, employing advanced techniques, such as quantitative real-time PCR (qRT-PCR) and RNA sequencing, could offer a more detailed understanding of CD36 gene expression and its regulatory mechanisms. Future investigations will show whether Ov-ERV-R13-CD36 represents a causal variant statistically associated with the observed trait, or simply a tag variant in linkage disequilibrium with a causal variant

Conclusions

In conclusion, our extensive examination of Ov-ERV-R13-CD36 within CD36 gene across multiple sheep breeds has revealed potential genetic correlations with OB and BCS, as well as fat deposition. Through detailed genomic investigations, including PCR analysis and data mining, we initially identified 31 ERV-RIPs among 58 sheep genomes. From these, 14 displayed polymorphic characteristics that intersect with candidate genes, with Ov-ERV-R13-CD36 being particularly associated with economically beneficial traits. Our findings emphasize Ov-ERV-R13-CD36 as a potential genetic marker, specifically its full-length 7,930 bp insertion within intron 1 of the mutant CD36 gene on chromosome 4. This presence, validated by PCR, revealed distinct variation in ERV-RIP-/-&+/+ genotypes across different breeds. Notably, Rahmani and Romanov x Rahmani exhibited a uniform RIP−/− genotype that correlated with significantly higher (p < 0.001) OB/BCS and increased fat accumulation. While these insights offer a basis for strategically targeting breeding programs aimed at enhancing these traits, it is essential to acknowledge the limitations of our study. A more comprehensive understanding of how the ERV insertion affects CD36 expression and its causal relationship with OB is needed. Future research could benefit from broader and more diverse sample populations, exploring environmental factors and further elucidating the function of CD36 variations in fat metabolism. Additionally, it would be valuable to investigate the potential impacts of Ov-ERV-R13-CD36 absences on CD36 functionality and their role in OB and fat accumulation processes more deeply. Employing qRT-PCR in future studies would enhance our understanding of gene expression dynamics related to these traits. Overall, this research lays a foundational step towards comprehending the genetic basis of these economically significant traits in sheep, opening promising avenues for subsequent breeding programs and genetic research.

Data availability

All data generated or analyzed during this study are included in this manuscript, its information files, and additional files; A. Additional File 1. B. Additional File 2. C. Additional File 3. D. Additional File 4. F. Additional File 6. G. Additional File 7.

Abbreviations

LTRs:

Long Terminal Repeats

ERVs:

Endogenous Retroviruses

RIPs:

Retroviral Insertion Polymorphisms

LINEs:

Long Interspersed Nuclear Elements

SINEs:

Short Interspersed Nuclear Elements

CD36 :

Cluster of differentiation 36

NCBI:

National Centre for Biotechnology Information

PCR:

Polymerase Chain Reaction

QTL:

Quantitative Trait Locus

WGS:

Whole-Genome Sequencing

lncRNA:

Long Non-Coding RNA Genes

Gag:

Group-Specific Antigen

Pol:

Polymerase

Pro:

Protease

Env:

Envelope

HWE :

Hardy-Weinberg Equilibrium

Ho :

Heterozygosity

He :

expected heterozygosity

Ne :

Effective Allele Number

PIC :

Polymorphic Information Content

References

  1. Platt RN, Vandewege MW, Ray DA. Mammalian transposable elements and their impacts on genome evolution. Chromosome Res. 2018;26(1–2):25–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Mager DL, Stoye JP. Mammalian Endogenous Retroviruses. Microbiol. Spectr. 2015;3(1):Mdna3-0009-2014.

  3. Dopkins N, Nixon DF. Activation of human endogenous retroviruses and its physiological consequences. Nat Rev Mol Cell Biol. 2024;25(3):212–22.

    Article  CAS  PubMed  Google Scholar 

  4. Becker D, Weikard R, Heimes A, Hadlich F, Hammon HM, Meyerholz MM, Petzl W, Zerbe H, Schuberth H-J, Hoedemaker M, et al. Allele-biased expression of the bovine APOB gene associated with the cholesterol deficiency defect suggests cis-regulatory enhancer effects of the LTR retrotransposon insertion. Sci Rep. 2022;12(1):13469.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Tang L, Swedlund B, Dupont S, Harland C, Costa Monteiro Moreira G, Durkin K, Artesi M, Mullaart E, Sartelet A, Karim L, et al. GWAS reveals determinants of mobilization rate and dynamics of an active endogenous retrovirus of cattle. Nat. Commun. 2024;15(1):2154.

  6. Eickbush TH, Malik HS. 2002. Origins and evolution of retrotransposons. Biology. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/9781555817954.CH49. Available from: https://api.semanticscholar.org/CorpusID:82631936.

  7. Kalendar R, Aizharkyn K, Khapilina O, Amenov A, Tagimanova D. Plant diversity and transcriptional variability assessed by retrotransposon-based molecular markers. Vavilov J Genet Breed. 2017;21(1):128–34.

    Article  Google Scholar 

  8. Kalendar R, Flavell AJ, Ellis TH, Sjakste T, Moisy C, Schulman AH. Analysis of plant diversity with retrotransposon-based molecular markers. Heredity. 2011;106(4):520–30.

    Article  CAS  PubMed  Google Scholar 

  9. Flavell AJ, Knox MR, Pearce SR, Ellis TH. Retrotransposon-based insertion polymorphisms (RBIP) for high throughput marker analysis. Plant J. 1998;16(5):643–50.

    Article  CAS  PubMed  Google Scholar 

  10. Lee J, Mun S, Kim DH, Cho CS, Oh DY, Han K. Chicken (Gallus gallus) endogenous retrovirus generates genomic variations in the chicken genome. Mob DNA. 2017;8:2.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Kalendar R, Schulman AH. IRAP and REMAP for retrotransposon-based genotyping and fingerprinting. Nat Protoc. 2006;1(5):2478–84.

    Article  CAS  PubMed  Google Scholar 

  12. Tam SM, Mhiri C, Vogelaar A, Kerkveld M, Pearce SR, Grandbastien MA. Comparative analyses of genetic diversities within tomato and pepper collections detected by retrotransposon-based SSAP. AFLP and SSR Theor Appl Genet. 2005;110(5):819–31.

    Article  CAS  PubMed  Google Scholar 

  13. Chessa B, Pereira F, Arnaud F, Amorim A, Goyache F, Mainland I, Kao RR, Pemberton JM, Beraldi D, Stear MJ, et al. Revealing the history of sheep domestication using retrovirus integrations. Science. 2009;324(5926):532–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Elleder D, Kim O, Padhi A, Bankert JG, Simeonov I, Schuster SC, Wittekindt NE, Motameny S, Poss M. Polymorphic integrations of an endogenous gammaretrovirus in the mule deer genome. J Virol. 2012;86(5):2787–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Hron T, Fabryova H, Elleder D. Insight into the epigenetic landscape of a currently endogenizing gammaretrovirus in mule deer (Odocoileus hemionus). Genomics. 2020;112(1):886–96.

    Article  CAS  PubMed  Google Scholar 

  16. Klymiuk N, Müller M, Brem G, Aigner B. Characterization of endogenous retroviruses in sheep. J Virol. 2003;77(20):11268–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Cumer T, Pompanon F, Boyer F. Old origin of a protective endogenous retrovirus (enJSRV) in the Ovis genus. Heredity. 2019;122(2):187–94.

    Article  CAS  PubMed  Google Scholar 

  18. Mager DL, Stoye JP. Studies of endogenous retroviruses reveal a continuing evolutionary saga. Nat Rev Microbiol. 2012;10(6):395–406. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/nrmicro2783.

    Article  CAS  Google Scholar 

  19. Moawad AS, Wang F, Zheng Y, Chen C, Saleh AA, Hou J, Song C. Evolution of Endogenous Retroviruses in the Subfamily of Caprinae. Viruses. 2024;16(3):398.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Verneret M, Leroux C, Faraut T, Navratil V, Lerat E, Turpin J. A genome-wide study of ruminants reveals two endogenous retrovirus families still active in goats. Mob DNA. 2025;16(1):4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Sakashita A, Maezawa S, Takahashi K, Alavattam KG, Yukawa M, Hu YC, Kojima S, Parrish NF, Barski A, Pavlicev M, et al. Endogenous retroviruses drive species-specific germline transcriptomes in mammals. Nat Struct Mol Biol. 2020;27(10):967–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Konkel MK, Walker JA, Batzer MA. LINEs and SINEs of primate evolution. Evol Anthropol. 2010;19(6):236–49.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Kyriakou E, Magiorkinis G. Interplay between endogenous and exogenous human retroviruses. Trends Microbiol. 2023;31(9):933–46.

    Article  CAS  PubMed  Google Scholar 

  24. Gozashti L, Feschotte C, Hoekstra HE. Transposable Element Interactions Shape the Ecology of the Deer Mouse Genome. Mol Biol Evol. 2023;40(4):msad069.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Wang J, Lu X, Zhang W, Liu G-H. Endogenous retroviruses in development and health. Trends Microbiol. 2024;32(4):342–54.

    Article  CAS  PubMed  Google Scholar 

  26. Dhillon P, Mulholland KA, Hu H, Park J, Sheng X, Abedini A, Liu H, Vassalotti A, Wu J, Susztak K. Increased levels of endogenous retroviruses trigger fibroinflammation and play a role in kidney disease development. Nat Commun. 2023;14(1):559.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Yu J, Qiu P, Ai J, Liu B, Han GZ, Zhu F, Zhang W, Cui J. Endogenous retrovirus activation: potential for immunology and clinical applications. Natl Sci Rev. 2024;11(4):nwae034.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Ueda MT, Kryukov K, Mitsuhashi S, Mitsuhashi H, Imanishi T, Nakagawa S. Comprehensive genomic analysis reveals dynamic evolution of endogenous retroviruses that code for retroviral-like protein domains. Mob DNA. 2020;11(1):29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Wang X, Chen Z, Murani E, D’Alessandro E, An Y, Chen C, Li K, Galeano G, Wimmers K, Song C. A 192 bp ERV fragment insertion in the first intron of porcine TLR6 may act as an enhancer associated with the increased expressions of TLR6 and TLR1. Mob DNA. 2021;12(1):20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Saleh AA, Xue L, Zhao Y. Screening Indels from the whole genome to identify the candidates and their association with economic traits in several goat breeds. Funct Integr Genomics. 2023;23(1):58.

    Article  CAS  PubMed  Google Scholar 

  31. Zhao C, Hu B, Zhang Z, Luo Q, Nie Q, Zhang X, Li H. Detection of CD36 gene polymorphism associated with chicken carcass traits and skin yellowness. Poult Sci. 2023;102(7):102691.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Glatz JF, Heather LC, Luiken JJ. CD36 as a gatekeeper of myocardial lipid metabolism and therapeutic target for metabolic disease. Physiol Rev. 2024;104(2):727–64.

    Article  CAS  PubMed  Google Scholar 

  33. Zhang N, Huang D, Li X, Yan J, Yan Q, Ge W, Zhou J. Identification and validation of oxidative stress-related genes in sepsis-induced myopathy. Medicine. 2024;103(18):e37933.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Hishida A, Nakatochi M, Sutoh Y, Nakano S, Momozawa Y, Narita A, Tanno K, Shimizu A, Hozawa A, Kinoshita K, Yamaji T, Goto A, Noda M, Sawada N, Ikezaki H, Nagayoshi M, Hara M, Suzuki S, Koyama T, Koriyama C, Katsuura-Kamano S, Kadota A, Kuriki K, Yamamoto M, Sasaki M, Iwasaki M, Matsuo K, Wakai K. GWAS meta-analysis of kidney function traits in Japanese populations. J Epidemiol. 2024;34(11):526–34. https://doiorg.publicaciones.saludcastillayleon.es/10.2188/jea.JE20230281.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Oquendo P, Hundt E, Lawler J, Seed B. CD36 directly mediates cytoadherence of Plasmodium falciparum parasitized erythrocytes. Cell. 1989;58(1):95–101.

    Article  CAS  PubMed  Google Scholar 

  36. Lyu Q, Lin Y, Pan Y, Guan X, Ji X, Peng M, Li Q, Wang Z, Zhang Z, Luo Z. The polymorphism analysis for CD36 among platelet donors. Sci Rep. 2024;14(1):8534.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Agarwal S, Saha S, Ghosh R, Sarmadhikari D, Asthana S, Maiti TK, Khadgawat R, Guchhait P. Elevated glycosylation of CD36 in platelets is a risk factor for oxLDL-mediated platelet activation in type 2 diabetes. FEBS J. 2024;291(2):376–91.

    Article  CAS  PubMed  Google Scholar 

  38. He H, Tang L, Jin Y, Wang Y, Wang H, Ding S, Chen Y, Tian J, Wang M, Duan S. High-Throughput CD36 Phenotyping on Human Platelets Based on Sandwich ELISA and Mutant Gene Analysis. Transfus Med Hemother. 2023;51(1):32–40. https://doiorg.publicaciones.saludcastillayleon.es/10.1159/000530039.PMID:38314240;PMCID:PMC10836864.

    Article  PubMed  PubMed Central  Google Scholar 

  39. Ashaq MS, Zhang S, Xu M, Li Y, Zhao B. The regulatory role of CD36 in hematopoiesis beyond fatty acid uptake. Life Sci. 2024;339:122442. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.lfs.2024.122442.

  40. Alattar AG, Storry JR, Olsson ML. Evidence that CD36 is expressed on red blood cells and constitutes a novel blood group system of clinical importance. Vox Sanguinis. 2024;119(5):496-504. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/vox.13595.

  41. Sun A, Wang Y, Wu Y, Ge F, Ye X, Wang H, Chen Y, Ding S, Tian J, Li Q, Liao Y, Duan S. Analysis of CD36 expression and genetic variation in blood donors in Zhongshan City, China. SSRN Electron J. 2022. Available from: https://api.semanticscholar.org/CorpusID:250162056.

  42. Asch AS, Silbiger S, Heimer E, Nachman RL. Thrombospondin sequence motif (CSVTCG) is responsible for CD36 binding. Biochem Biophys Res Commun. 1992;182(3):1208–17.

    Article  CAS  PubMed  Google Scholar 

  43. Navazo MDP, Daviet L, Savill J, Ren Y, Leung LL, McGregor JL. Identification of a domain (155–183) on CD36 implicated in the phagocytosis of apoptotic neutrophils. J Biol Chem. 1996;271(26):15381–5.

    Article  CAS  PubMed  Google Scholar 

  44. Serghides L, Crandall I, Hull E, Kain KC. The Plasmodium falciparum–CD36 interaction is modified by a single amino acid substitution in CD36. Blood. 1998;92(5):1814–9.

    Article  CAS  PubMed  Google Scholar 

  45. Simantov R, Silverstein RL. CD36: a critical anti-angiogenic receptor. Front Biosci. 2003;8(1–3):s874-882.

    Article  CAS  PubMed  Google Scholar 

  46. Adachi H, Tsujimoto M. Endothelial scavenger receptors. Prog Lipid Res. 2006;45(5):379–404.

    Article  CAS  PubMed  Google Scholar 

  47. Collot-Teixeira S, Martin J, McDermott-Roe C, Poston R, McGregor JL. CD36 and macrophages in atherosclerosis. Cardiovasc Res. 2007;75(3):468–77.

    Article  CAS  PubMed  Google Scholar 

  48. Martin CA, Longman E, Wooding C, Hoosdally SJ, Ali S, Aitman TJ, Gutmann DA, Freemont PS, Byrne B, Linton KJ. CD36, a class B scavenger receptor, functions as a monomer to bind acetylated and oxidized low-density lipoproteins. Protein Sci. 2007;16(11):2531–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Gautam S, Banerjee M. The macrophage Ox-LDL receptor, CD36 and its association with type II diabetes mellitus. Mol Genet Metab. 2011;102(4):389–98.

    Article  CAS  PubMed  Google Scholar 

  50. Ren Y. Peroxisome Proliferator-Activator Receptor: A Link between Macrophage CD36 and Inflammation in Malaria Infection. PPAR Res. 2012;640769. https://doiorg.publicaciones.saludcastillayleon.es/10.1155/2012/640769.

  51. Holmes RS. Comparative studies of vertebrate platelet glycoprotein 4 (CD36). Biomolecules. 2012;2(3):389–414.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Smith BK, Jain SS, Rimbaud S, Dam A, Quadrilatero J, Ventura-Clapier R, Bonen A, Holloway GP. FAT/CD36 is located on the outer mitochondrial membrane, upstream of long-chain acyl-CoA synthetase, and regulates palmitate oxidation. Biochem J. 2011;437(1):125–34.

    Article  CAS  PubMed  Google Scholar 

  53. Bionaz M, Vargas-Bello-Pérez E, Busato S. Advances in fatty acids nutrition in dairy cows: from gut to cells and effects on performance. J Anim Sci Biotechnol. 2020;11(1):110.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Du Z, Chen C, Zheng Y, Wang X, Song C. Retroviral Insertion Polymorphism (RIP) of Porcine Endogenous Retroviruses (PERVs) in Pig Genomes. Animals (Basel). 2024;14(4):621. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/ani14040621.

  55. Chen X, Li D. ERVcaller: identifying polymorphic endogenous retrovirus and other transposable element insertions using whole-genome sequencing data. Bioinformatics. 2019;35(20):3913–22.

    Article  CAS  PubMed  Google Scholar 

  56. Chen N. Using RepeatMasker to identify repetitive elements in genomic sequences. Curr Protoc Bioinformatics. 2004;Chapter 4:Unit 4.10.

  57. El-Gebali S, Mistry J, Bateman A, Eddy SR, Luciani A, Potter SC, Qureshi M, Richardson LJ, Salazar GA, Smart A, et al. The Pfam protein families database in 2019. Nucleic Acids Res. 2019;47(D1):D427–32.

    Article  CAS  PubMed  Google Scholar 

  58. Potter SC, Luciani A, Eddy SR, Park Y, Lopez R, Finn RD. HMMER web server: 2018 update. Nucleic Acids Res. 2018;46(W1):W200–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Quinlan AR, Hall IM. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics. 2010;26(6):841–2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Tarailo-Graovac M, Chen N. Using RepeatMasker to identify repetitive elements in genomic sequences. Curr Protoc Bioinformatics. 2009;25(1):1–14.

    Article  Google Scholar 

  61. Wang B, Saleh AA, Yang N, Asare E, Chen H, Wang Q, Chen C, Song C, Gao B. High Diversity of Long Terminal Repeat Retrotransposons in Compact Vertebrate Genomes: Insights from Genomes of Tetraodontiformes. Animals. 2024;14(10):1425.

    Article  PubMed  PubMed Central  Google Scholar 

  62. Guo M, Addy GA, Yang N, Asare E, Wu H, Saleh AA, Shi S, Gao B, Song C. PiggyBac Transposon Mining in the Small Genomes of Animals. Biology. 2023;13(1):24.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Kent WJ. BLAT–the BLAST-like alignment tool. Genome Res. 2002;12(4):656–64.

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Krishnan G, Bagath M, Devaraj C, Soren NM, Veeranna RK. Chemosensing of fat digestion by the expression pattern of GPR40, GPR120, CD36 and enteroendocrine profile in sheep. Res Vet Sci. 2022;150:89–97.

    Article  CAS  PubMed  Google Scholar 

  65. Wang J, Chitsaz F, Derbyshire MK, Gonzales NR, Gwadz M, Lu S, Marchler GH, Song JS, Thanki N, Yamashita RA, et al. The conserved domain database in 2023. Nucleic Acids Res. 2023;51(D1):D384–8.

    Article  CAS  PubMed  Google Scholar 

  66. Lu S, Wang J, Chitsaz F, Derbyshire MK, Geer RC, Gonzales NR, Gwadz M, Hurwitz DI, Marchler GH, Song JS, et al. CDD/SPARCLE: the conserved domain database in 2020. Nucleic Acids Res. 2020;48(D1):D265–8.

    Article  CAS  PubMed  Google Scholar 

  67. Marchler-Bauer A, Bo Y, Han L, He J, Lanczycki CJ, Lu S, Chitsaz F, Derbyshire MK, Geer RC, Gonzales NR, et al. CDD/SPARCLE: functional classification of proteins via subfamily domain architectures. Nucleic Acids Res. 2017;45(D1):D200–3.

    Article  CAS  PubMed  Google Scholar 

  68. Kalendar R, Tselykh TV, Khassenov B, Ramanculov EM. Introduction on Using the FastPCR Software and the Related Java Web Tools for PCR and Oligonucleotide Assembly and Analysis. Methods Mol Biol. 2017;1620:33–64.

    Article  CAS  PubMed  Google Scholar 

  69. Saleh AA, Hammoud MH, Dabour NA, Hafez EE, Sharaby MA. BMPR-1B, BMP-15 and GDF-9 genes structure and their relationship with litter size in six sheep breeds reared in Egypt. BMC Res Notes. 2020;13(1):215.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Saleh A, Hassanine N. Egyptian Sheep Breeds and genetic tools to improve. DOI. 2019;10: m9.

    Google Scholar 

  71. Morand-Fehr J, Hervieu P. Assessing the Body Condition of Goats: Importance and Method. HAL Open Science. 1999;231:22-34. Available from: https://hal.archives-ouvertes.fr/hal-02699194.

  72. Russel A, Doney J, Gunn R. Subjective assessment of body fat in live sheep. J Agric Sci. 1969;72(3):451–4.

    Article  Google Scholar 

  73. Saleh AA, Rashad AMA, Hassanine N, Sharaby MA, Zhao Y. Morphological body measurements, body indices, and their genetic background for several Chinese goat breeds. Trop Anim Health Prod. 2022;54(4):204.

    Article  PubMed  Google Scholar 

  74. Saleh AA, Rashad AMA, Hassanine N, Sharaby MA, Zhao Y. Evaluation of morphological traits and physiological variables of several Chinese goat breeds and their crosses. Trop Anim Health Prod. 2021;53(1):74.

    Article  PubMed  Google Scholar 

  75. Sen AR, Santra A, Karim SA. Carcass yield, composition and meat quality attributes of sheep and goat under semiarid conditions. Meat Sci. 2004;66(4):757–63.

    Article  CAS  PubMed  Google Scholar 

  76. Santos VA, Silva SR, Azevedo JM. Carcass composition and meat quality of equally mature kids and lambs. J Anim Sci. 2008;86(8):1943–50.

    Article  CAS  PubMed  Google Scholar 

  77. Ekiz B, Kecici PD, Ograk YZ, Yalcintan H, Yilmaz A. Evaluation of the functionality of EUROP carcass classification system in thin-tailed and fat-tailed lambs. Meat Sci. 2021;181: 108603.

    Article  PubMed  Google Scholar 

  78. Kirton AH, Woods EG, Duganzich DM. Predicting the fatness of lamb carcasses from carcass wall thickness measured by ruler or by a total depth indicator (TDI) probe. Livest Prod Sci. 1984;11(2):185–94.

    Article  Google Scholar 

  79. AOAC: Association of Official Analytical Chemists. Official Method of Chemical Analysis (18th ed.), Assoc. Off. Anal. Chem., Gaithersburg, MD, USA. 2005.

  80. Madruga MS, Medeiros EJLd, Sousa WHd, Cunha MdGG, Pereira Filho JM, Queiroga RdCRdE. Chemical composition and fat profile of meat from crossbred goats reared under feedlot systems. Rev Bras Zootec. 2009;38:547-552.

  81. Sambrook J, Maniatis Fritsch E. Molecular cloning: a laboratory manual. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press; 1989. p. 1997.

    Google Scholar 

  82. Nei M, Roychoudhury AK. Sampling variances of heterozygosity and genetic distance. Genetics. 1974;76(2):379–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Nei M, Li WH. Mathematical model for studying genetic variation in terms of restriction endonucleases. Proc Natl Acad Sci USA. 1979;76(10):5269–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Rebollo R, Miceli-Royer K, Zhang Y, Farivar S, Gagnier L, Mager DL. Epigenetic interplay between mouse endogenous retroviruses and host genes. Genome Biol. 2012;13(10):R89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Van de Lagemaat LN, Medstrand P, Mager DL. Multiple effects govern endogenous retrovirus survival patterns in human gene introns. Genome Biol. 2006;7(9):R86.

    Article  PubMed  PubMed Central  Google Scholar 

  86. Fueyo R, Judd J, Feschotte C, Wysocka J. Roles of transposable elements in the regulation of mammalian transcription. Nat Rev Mol Cell Biol. 2022;23(7):481–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Mita P, Boeke JD. How retrotransposons shape genome regulation. Curr Opin Genet Dev. 2016;37:90–100.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Gebrie A. Transposable elements as essential elements in the control of gene expression. Mob DNA. 2023;14(1):9.

    Article  PubMed  PubMed Central  Google Scholar 

  89. Grandi N, Pisano MP, Demurtas M, Blomberg J, Magiorkinis G, Mayer J, Tramontano E. Identification and characterization of ERV-W-like sequences in Platyrrhini species provides new insights into the evolutionary history of ERV-W in primates. Mob DNA. 2020;11(1):6.

    Article  PubMed  PubMed Central  Google Scholar 

  90. Guallar D, Climent M, Pérez-Palacios R, García-Tuñón I, Muniesa P, Schoorlemmer J. The pluripotency marker Rex-1/Zfp42 negatively regulates expression of endogenous retroviruses (ERV) and ERV-associated genes. Retrovirology. 2011;8(1):61.

    Article  Google Scholar 

  91. Halo JV, Pendleton AL, Jarosz AS, Gifford RJ, Day ML, Kidd JM. Origin and recent expansion of an endogenous gammaretroviral lineage in domestic and wild canids. Retrovirology. 2019;16(1):6.

    Article  PubMed  PubMed Central  Google Scholar 

  92. Ngo MH, Arnal M, Sumi R, Kawasaki J, Miyake A, Grant CK, Otoi T, Fernández de Luco D, Nishigaki K. Tracking the fate of endogenous retrovirus segregation in wild and domestic cats. J Virol. 2019;93(24):e01324-19. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/JVI.01324-19.

  93. Chong AY, Kojima KK, Jurka J, Ray DA, Smit AFA, Isberg SR, Gongora J. Evolution and gene capture in ancient endogenous retroviruses - insights from the crocodilian genomes. Retrovirology. 2014;11(1):71.

    Article  PubMed  PubMed Central  Google Scholar 

  94. Lee K-H, Horiuchi M, Itoh T, Greenhalgh DG, Cho K. Cerebellum-specific and age-dependent expression of an endogenous retrovirus with intact coding potential. Retrovirology. 2011;8(1):82.

    Article  PubMed  PubMed Central  Google Scholar 

  95. Gifford RJ, Blomberg J, Coffin JM, Fan H, Heidmann T, Mayer J, Stoye J, Tristem M, Johnson WE. Nomenclature for endogenous retrovirus (ERV) loci. Retrovirology. 2018;15(1):59.

    Article  PubMed  PubMed Central  Google Scholar 

  96. Kjeldbjerg AL, Villesen P, Aagaard L, Pedersen FS. Gene conversion and purifying selection of a placenta-specific ERV-V envelope gene during simian evolution. BMC Evol Biol. 2008;8(1):266.

    Article  PubMed  PubMed Central  Google Scholar 

  97. Russo M, Morelli S, Capranico G. Expression of down-regulated ERV LTR elements associates with immune activation in human small-cell lung cancers. Mob DNA. 2023;14(1):2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Bokor S, Legry V, Meirhaeghe A, Ruiz JR, Mauro B, Widhalm K, Manios Y, Amouyel P, Moreno LA, Molnàr D, et al. Single-nucleotide polymorphism of CD36 locus and obesity in European adolescents. Obesity. 2010;18(7):1398–403.

    Article  CAS  PubMed  Google Scholar 

  99. Clop A, Sharaf A, Castelló A, Ramos-Onsins S, Cirera S, Mercadé A, Derdak S, Beltran S, Huisman A, Fredholm M, et al. Identification of protein-damaging mutations in 10 swine taste receptors and 191 appetite-reward genes. BMC Genomics. 2016;17(1):685.

    Article  PubMed  PubMed Central  Google Scholar 

  100. Luo H, Zhang Y, Qin Y, Wu L. Association of ChREBP expression and polymorphism with serum biochemical indices in Cherry Valley duck. Pak J Zool. 2019;51:855–63.

    Article  CAS  Google Scholar 

  101. Rać ME, Safranow K, Poncyljusz W. Molecular Basis of Human CD36 Gene Mutations. Mol Med. 2007;13(5):288–96.

    Article  PubMed  PubMed Central  Google Scholar 

  102. Park YM. CD36, a scavenger receptor implicated in atherosclerosis. Exp Mol Med. 2014;46(6):e99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Ladanyi A, Mukherjee A, Kenny HA, Johnson A, Mitra AK, Sundaresan S, Nieman KM, Pascual G, Benitah SA, Montag A, et al. Adipocyte-induced CD36 expression drives ovarian cancer progression and metastasis. Oncogene. 2018;37(17):2285–301.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Vroegrijk IO, van Klinken JB, van Diepen JA, van den Berg SA, Febbraio M, Steinbusch LK, Glatz JF, Havekes LM, Voshol PJ, Rensen PC, et al. CD36 is important for adipocyte recruitment and affects lipolysis. Obesity. 2013;21(10):2037–45.

    Article  CAS  PubMed  Google Scholar 

  105. Christiaens V, Van Hul M, Lijnen HR, Scroyen I. CD36 promotes adipocyte differentiation and adipogenesis. Biochim Biophys Acta. 2012;1820(7):949–56.

    Article  CAS  PubMed  Google Scholar 

  106. Keller KL, Liang LC, Sakimura J, May D, Van Belle C, Breen C, Driggin E, Tepper BJ, Lanzano PC, Deng L. Common variants in the CD36 gene are associated with oral fat perception, fat preferences, and obesity in African Americans. Obesity. 2012;20(5):1066–73.

    Article  CAS  PubMed  Google Scholar 

  107. Holloway GP, Luiken JJ, Glatz JF, Spriet LL, Bonen A. Contribution of FAT/CD36 to the regulation of skeletal muscle fatty acid oxidation: an overview. Acta Physiol. 2008;194(4):293–309.

    Article  CAS  Google Scholar 

  108. Fagerberg L, Hallström BM, Oksvold P, Kampf C, Djureinovic D, Odeberg J, Habuka M, Tahmasebpoor S, Danielsson A, Edlund K, et al. Analysis of the human tissue-specific expression by genome-wide integration of transcriptomics and antibody-based proteomics. Mol Cell Proteomics. 2014;13(2):397–406.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

‘‘Not applicable.’’

Funding

This work was supported by the National Key Research and Development Program of China (2022YFD1300201), and the High-end Talent Support Program of Yangzhou University to Chengyi Song.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: C.S., and A.A.S.; Data curation: A.A.S., A.M., X.W., N.Y., C.C., and Y.Z.; Formal analysis: A.A.S., C.S, B.G., and Y.Z; Funding acquisition: C.S.; Investigation: A.A.S. Methodology: A.A.S., A.M., N.Y. Project administration: C.S., and A.A.S.; Supervision: C.S.; Writing—original draft: A.A.S.; Writing-review and editing: A.A.S. and C.S. All authors have read and agreed to the published version of the manuscript.

Corresponding author

Correspondence to Chengyi Song.

Ethics declarations

Ethics approval and consent to participate

All procedures were performed in strict accordance with the guidelines and regulations proposed by the Faculty of Agriculture (Al-Shatby), Alexandria University, Egypt. All the animal experiments were approved by the ethics committee of Alexandria University, Faculty of Agriculture (Al-Shatby), Egypt (No. AU082209203103) and the ethics committee of Yangzhou University, College of Animal Science and Technology, China (No. 202403005). All procedures and experimental protocols were in accordance with the Guide for the Care and Use of Agricultural Animals in Research and Teaching, Federation of Animal Science Societies (FASS, 2010) https://www.aaalac.org/about/Ag_Guide_3rd_ed.pdf. Also, the study was carried out in compliance with the ARRIVE guidelines (https://arriveguidelines.org).

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

13100_2025_349_MOESM1_ESM.xlsx

Additional File 1. Supplementary Tables: Table S1. The tested 57 assembled non-reference sheep genomes in addition to one reference genome. Table S2. The library included 28 ERV families sourced from LTR retrotransposons used in the current investigation. Table S3. The anticipated variations in the insertion of full-length ERVs in the genomes of sheep. Table S4. Primers utilized in PCR-Based activation tests for 31 Endogenous Retroviruses (ERVs). Table S5. Investigated Genomes for CD36 gene in sheep and eight other species. Table S6. Reference genome and 42 non-reference genomes in Sheep.Table S7. Body Condition Score (BCS). Table S8. Analysis of Endogenous Retroviruses (ERVs) in 58 sheep Genomes: Distribution and Characterization of Full-Length Sequences. Table S9. Investigating CD36 gene in sheep and eight other species. Table S9. Exploring the differentiation in Ov-ERV-R13-CD36 and CD36 gene for Reference genome and 42 non-reference genomes in Sheep. Table S11. Ov-ERV-R13-CD36 genotyping in the studied sheep breeds and population genetic parameters

Additional File 2. Mining Protocols: Graphs

13100_2025_349_MOESM3_ESM.pdf

Additional File 3. Data Sets: Data Set 1. 28 Endogenous Retrovirus (ERV) groups were identified through a comparison of genomic coordinates. Data Set 2. 31 ERV-RIP insertion polymorphisms were predicted through a comparison of genomic coordinates using 28 ERV groups. Data Set 3. 14 out of the 31 full-length Endogenous Retrovirus (ERV) insertion polymorphisms were chosen following a comparison of genomic coordinates. Data Set 4. Sequences for Chromosome 4 in sheep (Ovis aries), CD36 gene and Ov-ERV-R13-CD36

13100_2025_349_MOESM4_ESM.pdf

Additional File 4 Supplementary Figures: Fig. S1. Sheep breeds scrutinized in the present study; Barki (B), Rahmani (R), Rahmani x Barki cross (RB), Awassi (A), Ossimi (O) and Romanov x Rahmani (V). Fig. S2. A) Points of consideration; 1. the sterna (sternum region), 2. the ribs (the rib cage) and 3. the lumbar vertebrae region which involves the spinous and transverse processes of spine ''behind the ribs'' (a: spinous process, b: flank and c: transverse process) are palpated to determine BCS. B) Diagram ''cross-section'' for assigning BCS in small ruminants. Fig. S3. A) The Rib-eye area (longissimus dorsi) in sheep. B) The muscle longissimus thoracis et lumborum (LL) in sheep. Fig. S4. A) Design principles for full-length LTR polymorphic primers. 1&2) LTR-primer produces a shorter band 1,056 bp. 3&4) Side-primer produces a band of 1,159 bp. B) After PCR, a band of 1,056 bp with Primer (1 and 2) and no band of 1,159 bp with Primer (1 and 3) indicates ERV+/+. While a band of 1159 bp without the 1,056 bp band suggests ERV-/-, while both bands mean ERV+/-. Fig. S5. Analysis of CD36 gene and its genomic regions and nearby genes in the reference genomes of sheep (A), goats (B), cattle (C), buffaloes (D), pigs (E), rabbit (F), chicken(G), zebrafish (H) and domestic ferret (I). Fig. S6. PCR Verification for the selected Ov-ERV-R13-CD36 in 24 individuals (n=12♂ and 12♀) of Rahmani x Barki crossbred Breed, M: DNA Ladder 5 kbp

Additional File 5. Domains for CD36 and Ov-ERV-R13-CD36.

Additional File 6. The alignment of CD36 gene for one reference and 42 non-reference genomes.

13100_2025_349_MOESM7_ESM.fasta

Additional File 7. The insertion Ov-ERV-R13-CD36 in CD36 gene for Reference genome and selected 21 non-reference genomes in Sheep.

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/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Saleh, A.A., Moawad, A.S., Yang, N. et al. Association of a 7.9 kb Endogenous Retrovirus Insertion in Intron 1 of CD36 with Obesity and Fat Measurements in Sheep. Mobile DNA 16, 12 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13100-025-00349-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13100-025-00349-w

Keywords