Next Article in Journal
Solid Pseudopapillary Neoplasm of the Pancreas and Abdominal Desmoid Tumor in a Patient Carrying Two Different BRCA2 Germline Mutations: New Horizons from Tumor Molecular Profiling
Next Article in Special Issue
Genome-Wide Analysis for Early Growth-Related Traits of the Locally Adapted Egyptian Barki Sheep
Previous Article in Journal
Special Issue “Cardiovascular Genetics”
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Analyses Reveal Genetic Convergence of Prolificacy between Goats and Sheep

1
Key Laboratory of Animal Genetics, Breeding and Reproduction of Ministry of Agriculture and Rural Affairs, Institute of Animal Science, Chinese Academy of Agricultural Sciences, Beijing 100193, China
2
Yunnan Animal Science and Veterinary Institute, Kunming 650224, China
3
College of Life Science and Food Engineering, Hebei University of Engineering, Handan 056038, China
4
Annoroad Gene Technology Co., Ltd., Beijing 100176, China
*
Authors to whom correspondence should be addressed.
Genes 2021, 12(4), 480; https://doi.org/10.3390/genes12040480
Submission received: 24 February 2021 / Revised: 16 March 2021 / Accepted: 23 March 2021 / Published: 26 March 2021
(This article belongs to the Special Issue Genetics and Breeding of Small Ruminants)

Abstract

:
The litter size of domestic goats and sheep is an economically important trait that shows variation within breeds. Strenuous efforts have been made to understand the genetic mechanisms underlying prolificacy in goats and sheep. However, there has been a paucity of research on the genetic convergence of prolificacy between goats and sheep, which likely arose because of similar natural and artificial selection forces. Here, we performed comparative genomic and transcriptomic analyses to identify the genetic convergence of prolificacy between goats and sheep. By combining genomic and transcriptomic data for the first time, we identified this genetic convergence in (1) positively selected genes (CHST11 and SDCCAG8), (2) differentially expressed genes (SERPINA14, RSAD2, and PPIG at follicular phase, and IGF1, GPRIN3, LIPG, SLC7A11, and CHST15 at luteal phase), and (3) biological pathways (genomic level: osteoclast differentiation, ErbB signaling pathway, and relaxin signaling pathway; transcriptomic level: the regulation of viral genome replication at follicular phase, and protein kinase B signaling and antigen processing and presentation at luteal phase). These results indicated the potential physiological convergence and enhanced our understanding of the overlapping genetic makeup underlying litter size in goats and sheep.

1. Introduction

As ruminant livestock, sheep and goats are globally distributed and play important roles in providing meat, milk, and fur to humans. One primary objective of sheep and goat breeding programs is to increase the yield by phenotypic and genetic selection, and female fecundity is a basic and crucial key. Litter size (LS), the lambing number per parturition per ewe or the kidding number per parturition per doe, is an important component of goat and sheep productivity. Importantly and interestingly, variation in individual LS is common within breeds, such as 1–4 in Yunshang black goats, 1–5 Romanov sheep, and 1–6 in Finnsheep [1,2]. Breeds with exceptional prolificacy can provide genetic material to develop new lines/populations with ideal performance.
It is important to elucidate the genetic mechanisms of LS to efficiently utilize these prolific resources. Therefore, numerous studies have been separately performed to enhance our understanding of goat and sheep LS [3,4,5,6,7]. Compared with sheep, knowledge of hircine prolificacy is in its infancy. Thus, previous studies explored the association between hircine LS and homologous genetic markers first identified in sheep (such as FecB and FecX) [8,9], in which the genetic convergence in prolificacy between goats and sheep was expected in principle, at least at the single gene level. However, the approach using candidate genes/loci did not work well. Notwithstanding the convergent genomic signatures of domestication and high-altitude adaptation that have been explored in sheep and goats [10,11], thus far, there has been a paucity of research on the genetic convergence of prolificacy from a whole-genome perspective. We hypothesize that the genetic convergence of prolificacy occurred in goats and sheep, likely because they were exposed to similar natural and/or artificial selection forces.
Therefore, the objective of this study was to identify the genetic convergence of prolificacy between goats and sheep using whole-genome data by examining (1) positively selected genes (PSGs), (2) differentially expressed genes (DEGs), and (3) biological pathways. These results will enhance our knowledge of prolificacy and can be used to improve the molecular design breeding of both goats and sheep.

2. Materials and Methods

2.1. Animals, Phenotypes and Sample Collection

This study included two Yunshang black goat populations (Yunshang black goats used for DNA sequencing (YSD), n = 40; Yunshang black goats used ovarian mRNA sequencing (YSR), n = 20) and three sheep populations (Finnsheep, FIN, n = 19; Romanov, ROM, n = 19; and Small Tailed Han sheep, STH, n = 12); the YSD, FIN and ROM data were obtained from previous studies [1,2]. All animals were as unrelated as possible as determined by pedigree or farmers’ knowledge. On the basis of average LS, the five populations were classified into a low-yield group (LG, control) and a high-yield group (HG, case) to represent extreme individuals within each breed (Table S1). For each population, a highly significant difference was observed in LS between the HG and LG (p-value < 0.01, Mann–Whitney U test, Table 1).
To detect PSGs for LS, 20 YSD (LS ≤ 1.5), 9 FIN (LS ≤ 2), and 9 ROM (LS ≤ 1.8) were labeled as controls, and 20 YSD (LS > 2.3), 10 FIN (LS > 3.28), and 10 ROM (LS ≥ 3) were labeled as cases. Likewise, to identify DEGs for LS, 10 YSR (LS ≤ 2) and 6 STH (LS = 1) were selected as controls, and 10 YSR (LS ≥ 3) and 6 STH (LS > 2.66) were selected as cases. For ovary collection, estrus synchronization with controlled internal drug release (CIDR, progesterone 300 mg) for 16 and 12 days was separately performed for YSR and STH. Finally, 10 YSR (5 cases and 5 controls) and 6 STH (3 cases and 3 controls) were slaughtered within 45–48 h of CIDR removal for ovary harvest at follicular phase (FP), and the remaining 10 YSR and 6 STH were slaughtered on days 11 and 9 for ovary harvest at luteal phase (LP).

2.2. Whole-Genome Single Nucleotide Polymorphism (SNP)

Genomic DNA was extracted from the jugular venous blood of YSD and ear marginal tissues of FIN and ROM following a standard phenol–chloroform method. To obtain genomic SNPs of YSD, whole-genome sequencing of a single library was performed using the Illumina NovaSeq 6000 system (Illumina Inc., San Diego, CA, USA), and 150-bp paired-end reads were generated. To obtain clean reads, the following unqualified raw reads were removed: (1) adaptor-polluted reads, (2) reads with over 50% bases with a Phred score <19, and (3) reads with over 5% N bases. Clean reads were mapped to the ASR1 reference genome using a Burrows–Wheeler Aligner v0.7.9a [12], with an average depth of ~20-fold coverage, and SNPs were subsequently called using a Genome Analysis Toolkit v3.3 [13]. FIN and ROM were genotyped by the Ovine Infinium HD BeadChip, which comprises 606,006 SNPs. To obtain eligible SNPs and samples, the following quality control criteria were implemented in Plink v1.90 [14]: (1) call rate >90%, (2) SNPs on autosomes, (3) minor allele frequency >1%, (4) missing genotype frequency <10%, and (5) Hardy–Weinberg disequilibrium p-value > 10−6.

2.3. Population Structure

To explore the genetic relationship of individuals from YSD, FIN and ROM, Plink command “--indep-pairwise 50 5 0.2” was run to generate independent effective SNP sets for principal component analysis [14].

2.4. Selective Sweeps

To detect PSGs for LS, the fixation index (Fst), which is a measure of population differentiation, was calculated in VCFtools v0.1.14 [15] using Weir and Cockeram’s method [16] to compare LG and HG for YSD (Fst(YSD)), FIN (Fst(FIN)) and ROM (Fst(ROM)). For YSD, a 100-kb window with a 50-kb step size was used. For sheep, Fst(FIN) and Fst(ROM) were estimated SNP by SNP, and their root mean square (Fst_RMS) was then calculated with the following formula:
Fst _ RMS = 1 2   ( Fst ( FIN ) 2 + Fst ( ROM ) 2 )
To calculate Fst_RMS, only SNPs shared by FIN and ROM and those with nonnegative Fst values were used. The results were visualized using R package qqman [17].
To identify loci with extremely high Fst values that drive reproductive differences of LG and HG, we ranked the regions/loci in descending order on the basis of Fst(YSD) and Fst_RMS, respectively. The regions with extremely high Fst(YSD) (0.05, top 0.8%) were treated as candidates for goats, and the SNPs with extremely high Fst_RMS (0.3, top 0.7%) were selected as candidates for sheep. Finally, PSGs wholly or partially located at these regions or covering candidate SNPs were detected by ANNOVAR v2020-06-07 [18].
To explore the association of three HS3ST2 variations and reproductive performance, an additional 544 2–5-year-old female Yunshang black goats were genotyped using a MassARRAY® SNP system. The primer sequences are in Table S2. Chi-squared test (all expected frequencies ≥ 5) or Fisher’s exact test were employed to detect associations between the genotype and LS of Yunshang black goats.

2.5. Ovarian RNA-Seq

The total RNA of hircine ovaries for LG and HG at FP and LP (five biological replicates) was extracted and sequenced using the Illumina NovaSeq 6000 platform (Illumina Inc., San Diego, CA, USA). After filtering out the adapter reads and unqualified reads, clean reads were mapped to the ASR1 reference genome using HISAT2 [19], followed by the assembly and calculation of read count and fragments per kilobase per million values in StringTie [20], and identification of DEGs with DESeq2 [21]. In the two comparisons of LG vs. HG, genes were accepted as being differentially expressed when (1) q-value (p-value adjusted by the Benjamini–Hochberg method) <0.5, and |Log2FoldChang| ≥ 0.24. Likewise, the total RNA of 12 ovine ovaries for LG and HG at FP and LP (three biological replicates) were subjected to the same processing, except they were mapped to the Oar_v3.1 reference genome.

2.6. Identification of Genetic Convergence

To reveal the potential genetic convergence in goat and sheep prolificacy, we focused on (1) PSGs, (2) DEGs, and (3) biological pathways at the level of gene set. Given the availability of PSGs in HG using the number of segregating sites by length (nSL), a haplotype-based statistic with test power to both soft and hard sweeps [2,22], convergent PSGs were identified as the intersection of ovine candidates identified by Fst_RMS and hircine candidates shared by Fst(YSD) and nSL (≥6) of HG. Convergent DEGs at LP and FP were defined as these genes that were shared by goats and sheep, but excluded their expression trend. The number of genes shared by Fst(YSD) and nSL of HG is limited and nonideal for gene set analysis; therefore the Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway analyses were performed for the PSGs on the basis of Fst(YSD) and Fst_RMS using Metascape [23] to obtain the related gene set which was composed of functionally associated PSGs. Terms with p-value < 0.01 were considered to be significantly different. Similarly, DEGs were also used to identify the connected gene set for prolificacy between goats and sheep. Therefore, convergent biological pathways were identified as the mutual GO (including only biological processes) and KEGG Ontology (ko) terms.

3. Results and Discussion

3.1. Population Structure

The principal component analysis (Figure S1) demonstrated a clear split between FIN and ROM along the first principal component (PC), indicating that they were independent populations. However, they were northern European short-tailed breeds with common ancestry components [24]. Therefore, they are thought to be useful for exploring the genetic mechanisms underlying LS as single and separate herds. For both goats and sheep, the cases and controls within each breed were not split by the first two PCs, which was consistent with the circumstances of dogs in a previous report on behavior [25]. These findings indicate that population structure was not linked to LS of each breed.

3.2. Convergent PSGs Linked to LS

To reveal the signatures of selection for goat prolificacy, a set of 350 PSGs was identified by Fst(YSD), of which HS3ST2 on chromosome 25 was found to have the strongest selection signature (Figure 1 and Table S3). The variation g.20423876C>T at HS3ST2 was significantly associated with LS (Table S4). Coding heparan sulfate-glucosamine 3-sulfotransferase 2, HS3ST2 was reported to be linked to circadian rhythm regulation in the pineal gland of rats [26]. HS3ST5, an HS3ST2 paralog, was associated with reproductive seasonality traits in sheep [27]. Given the association of seasonality-associated DIO3 and LS in pigs and goats, and the empirical concurrence of higher LS and year-round estrus in goats [2,28], these results indicate that HS3ST2 may be involved in female reproductive health, and seasonal breeding in goats.
Eight overlapping PSGs were revealed on the basis of Fst(YSD) and nSL of HG in goats: CHST11, SDCCAG8, ATF6, FAM13A, HSP90B1, IBSP, IFNLR1, and STAB2 (Figure 2). Of these, two genes (CHST11 and SDCCAG8) were also under selection in sheep for which 162 PSGs were identified by Fst_RMS (Figure 1 and Figure 2). This number was significantly larger than expected (p-value = 0.05), suggesting that CHST11 and SDCCAG8 are significantly convergent in goat and sheep prolificacy. CHST11 (also known as C4ST-1), which encodes carbohydrate sulfotransferase 11, is a target gene of transforming growth factor β (TGFβ)-like growth factors (such as TGFβ, BMP2, and Activin) [29]. Chst11 was found to regulate the proliferation and apoptosis of embryo chondrocytes in mice [30]. The mRNA level of CHST11 increases upon the interaction of BMP4 with a phospho-Smad3 binding site in the CHST11 promoter [31]. Furthermore, CHST11 was linked to ovarian cancer [32], endometrial carcinomas [33], and porcine farrowing interval [34]. However, the CHST11 transcripts were not differentially expressed in goat ovary (Table S5). Another PSG, SDCCAG8, is a risk gene for Bardet–Biedl syndrome, which is primarily characterized by postaxial polydactyly, hypogonadotropic hypogonadism, and/or genitourinary malformations, and secondly characterized by polycystic ovary syndrome [35]. SDCCAG8 is a candidate gene for hind limb and rib cage malformations in mice [36,37] and teat number in pigs [38]. Interestingly, BMPR1B, a major gene for sheep LS, plays important roles in both chondrogenesis and embryogenesis [39,40], indicating the pleiotropism of candidate genes for bone formation and reproduction. This was, to some extent, supported by the “osteoclast differentiation” pathway, which was enriched by some PSGs for LS (Table 2). Moreover, there is a transcription factor binding site at SDCCAG8 for SMAD5, a member of the TGFβ signaling pathway, and strong differentiation signals have been observed in SMAD5 of Shaanbei White Cashmere goats with different LS [7]; this may contribute to the differential expression of SDCCAG8 at both LP and FP (Table S5). Taken together, these findings indicate that CHST11 and SDCCAG8 regulate LS via TGFβ and BMP signaling in goats and sheep.

3.3. Convergent DEGs Associated with LS

For goats, compared with LG, 2207 (1126 up- and 1081 down-regulated) and 1665 DEGs (826 up- and 839 down-regulated) were identified in HG at LP and FP, respectively (Figure 3 and Tables S6 and S7). For sheep, 104 and 39 DEGs were identified in HG compared with LG at LP and FP, respectively (Figure 3 and Tables S8 and S9). In total, eight overlapping DEGs were shared by goats and sheep, including three (SERPINA14, RSAD2, and PPIG) at FP and five (IGF1, GPRIN3, LIPG, SLC7A11, and CHST15) at LP (Figure 2 and Figure 4). Of these, IGF1 and SLC7A11 were identified as DEGs in the transcriptomic analysis of ovaries from pigs with high and low LS [41]. Additionally, SLC7A11 is differentially expressed in ovaries among sheep breeds with different fecundity [42]. IGF1 is a candidate gene for LS in goats and sheep [43,44,45], suggesting the convergence at the levels of DNA and change in mRNA expression.
Previous studies reported the differential expression of LIPG mRNA or protein in the theca interna during the transition from small to large antral bovine follicles [46], in the placenta of Berkshire pigs with different LS [47], and in the human follicular fluid of poor ovarian responders compared with a control group [48]. Additionally, a genome-wide association study on Large White pigs identified LIPG as a candidate gene for piglet uniformity measured as the coefficient of variation of birth weights within one litter [49]. These findings highlight the potential roles in follicular development and pregnancy. Secreted from uterine endometrium and induced by progesterone, uterine serpin (SERPINA14) plays crucial roles in the pregnancy of farm animals, such as nutrient supply to the conceptus, immunoregulation, and maternal recognition of pregnancy [50,51,52]. RSAD2 is a marker for dendritic cell maturation [53] and early pregnancy diagnosis in dairy cows [54]. Moreover, compared with healthy follicles, RSAD2 was up-regulated in early atresia and progressive atresia groups in pigs [55]. The functions of GPRIN3, CHST15, and PPIG in reproduction are unclear. Collectively, it is likely that these DEGs influence female reproduction by multiple processes, such as follicular development, immunoregulation, and pregnancy maintenance.

3.4. Convergent Biological Pathways

At the single gene level, it is possible to infer convergent signals from overlapping DEGs or PSGs between goats and sheep. At the gene set level, GO and KEGG pathway analyses for PSGs and DEGs were implemented to identify the overlapping biological pathways that include a series of functionally connected genes. For goat PSGs, 192 GO and 30 ko terms were revealed, as expected, six acrosome reaction-related GO terms were detected: binding of sperm to zona pellucida (GO:0007339), regulation of acrosome reaction (GO:0060046), single fertilization (GO:0007338), regulation of fertilization (GO:0080154), sperm–egg recognition (GO:0035036), and fertilization (GO:0009566) (Table S10). Additionally, some endocrine resistance ko terms were identified: endocrine resistance (ko01522), estrogen signaling pathway (ko04915 and hsa04915), relaxin signaling pathway (hsa04926), and fluid shear stress and atherosclerosis (hsa05418) (Table S11). For sheep PSGs, we identified 132 GO and 58 ko terms, of which osteoclast differentiation (ko04380 and hsa04380), ErbB signaling pathway (hsa04012 and ko04012), and relaxin signaling pathway (hsa04926) were shared by goats (Figure 2 and Table 3).
For goat DEGs, as expected, the phosphatidylinositol 3-kinase (PI3K)-Akt signaling pathway (ko04151), oocyte meiosis (ko04114), progesterone-mediated oocyte maturation (ko04914), and estrogen signaling pathway (ko04915) were in the top 10 ko terms (Figure S2), which reemphasizes the significance of some biological processes in LS. Consequently, goats shared three terms with sheep: protein kinase B signaling (GO:0043491) and antigen processing and presentation (ko04612) at LP, and the regulation of viral genome replication (GO:0045069) at FP (Table 3 and Figure 2). An implication of these convergent biological pathways is that there may be differences in relaxin concentration during pregnancy, immunity, and disease resistance that contribute to LS in goats and sheep. Moreover, it is helpful to pinpoint the potential causal genes for LS, such as ADCY3, which is necessary for male fertility [56]; TGFB2, which participates in ovine fecundity [57]; and IGFBP5, which is associated with LS in pigs [58]. Therefore, the convergence of biological pathways from the perspective of gene sets and single genes, indicates potential physiological convergence and enhances our understanding of the genetic basis underlying goat and sheep LS.
From a genomic and transcriptomic standpoint, we attempted to provide a genome-wide landscape to reveal the genetic convergence of prolificacy between goats and sheep. The results may indicate the physiological convergence of goats and sheep; however, more research is needed to characterize reproductive hormones and biological factors. However, these results may differ among other populations/breeds because of the sample size and/or genetic backgrounds in this study. Therefore, it would be useful to integrate an abundance of multi-omics data, such as genomics, transcriptomics, proteomics, and metabolomics, for additional populations or reproductive organs in a large population.

4. Conclusions

By combining genomic and transcriptomic data, this study identified the genetic convergence of prolificacy between goats and sheep for the first time on the basis of (1) PSGs (CHST11 and SDCCAG8), (2) DEGs (SERPINA14, RSAD2, and PPIG at FP, and IGF1, GPRIN3, LIPG, SLC7A11, and CHST15 at LP), and (3) biological pathways (PSG: osteoclast differentiation, ErbB signaling pathway and relaxin signaling pathway; DEG: regulation of viral genome replication at FP, and protein kinase B signaling and antigen processing and presentation at LP). These results indicated potential physiological convergence and enhanced our understanding of the overlapping genetic makeup underlying LS between goats and sheep.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/genes12040480/s1, Figure S1: Principal component analysis of goats and sheep, Figure S2: Top 10 KEGG terms of differentially expressed genes from luteal phase (LP) and follicular phase (FP) in Yunshang black goats, Table S1: Phenotypes of animals involved in this study, Table S2: Primer sequences of HS3ST2, Table S3: Positively selected genes identified by Fst, Table S4: Association between loci on HS3ST2 and litter size of Yunshang black goats, Table S5: Ovarian mRNA expression of CHST11 and SDCCAG8 in Yunshang black goats, Table S6: Differentially expressed mRNAs of Yunshang black goat ovaries at luteal phase, Table S7: Differentially expressed mRNAs of Yunshang black goat ovaries at follicular phase, Table S8: Differentially expressed mRNAs of Small Tailed Han sheep ovaries at luteal phase, Table S9: Differentially expressed mRNAs of Small Tailed Han sheep ovaries at follicular phase, Table S10: Functional category of positively selected genes in Yunshang black goats, Table S11: Terms related to acrosome reaction and endocrine resistance.

Author Contributions

Conceptualization, M.C. and Q.H.; methodology, L.T.; software, L.T. and Y.S.; validation, L.T., X.H. and Y.L.; formal analysis, L.T.; investigation, L.T., Y.J. and Y.O.; resources, M.C. and Q.H.; data curation, L.T. and Y.S.; writing—original draft preparation, L.T.; writing—review and editing, X.H., Q.H. and M.C.; visualization, L.T.; supervision, M.C. and Q.H.; project administration, M.C.; funding acquisition, M.C. and Q.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (31772580), the Earmarked Fund for China Agriculture Research System (CARS-38), the Agricultural Science and Technology Innovation Program of China (CAAS-ZDRW202106 and ASTIP-IAS13), the Science and Technology Innovation Talents Program of Yunnan Province (2018HC017), the Basic Research Foundation Key Project of Yunnan Province (202001AS070002). The APC was funded by 31772580.

Institutional Review Board Statement

STH, YSD and YSR involved in this study were approved by the Science Research Department (in charge of animal welfare issues) of the Institute of Animal Sciences, Chinese Academy of Agricultural Sciences (IAS-CAAS) (Beijing, China) and ethical approval was given by the Animal Ethics Committee of the IAS-CAAS (No. IASCAAS-AE-03 for STH, and No. IAS2019-63 for YSD and YSR).

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw sequencing data of 40 Yunshang black goats (YSD) have been deposited in the NCBI SRA database under accession code PRJNA611688. Genotype and phenotype datasets for Finnsheep and Romanov sheep were available at the following link: https://www.animalgenome.org/repository/pub/CAAS2018.0302/ (accessed on 24 September 2020). The RNA-seq data reported in this article are available upon request for research purpose.

Acknowledgments

We are grateful to all volunteers who assisted in sampling. We thank Liwen Bianji, Edanz Group China (www.liwenbianji.cn/ac, accessed on 10 March 2021), for editing the English text of a draft of this manuscript.

Conflicts of Interest

The authors declare no conflict of interest. Yezhen Shen is employee of Annoroad Gene Technology Co. Ltd., Beijing, China. This article has no conflict of interest with the company.

Abbreviations

LSlitter size
LGlow-yield group
HGhigh-yield group
FPfollicular phase
LPluteal phase
PSGspositively selected genes
DEGsdifferentially expressed genes
GOgene ontology
KEGGKyoto Encyclopedia of Genes and Genomes
koKEGG ontology
YSDYunshang black goats for DNA sequencing
YSRYunshang black goats for mRNA sequencing
FINFinnsheep
ROMRomanov sheep
STHSmall Tailed Han sheep
Fst_RMSroot mean square of fixation index (Fst)
nSLnumber of segregating sites by length

References

  1. Xu, S.S.; Gao, L.; Xie, X.L.; Ren, Y.L.; Shen, Z.Q.; Wang, F.; Shen, M.; Eyϸórsdóttir, E.; Hallsson, J.H.; Kiseleva, T.; et al. Genome-wide association analyses highlight the potential for different genetic mechanisms for litter size among sheep breeds. Front. Genet. 2018, 9, 118. [Google Scholar] [CrossRef]
  2. Tao, L.; He, X.Y.; Jiang, Y.T.; Lan, R.; Li, M.; Li, Z.M.; Yang, W.F.; Hong, Q.H.; Chu, M.X. Combined approaches to reveal genes associated with litter size in Yunshang black goats. Anim. Genet. 2020, 51, 924–934. [Google Scholar] [CrossRef] [PubMed]
  3. Yao, Y.; Pan, Z.; Di, R.; Liu, Q.; Hu, W.; Guo, X.; He, X.; Gan, S.; Wang, X.; Chu, M. Whole genome sequencing reveals the effects of recent artificial selection on litter size of Bamei mutton sheep. Animals 2021, 11, 157. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, J.J.; Zhang, T.; Chen, Q.M.; Zhang, R.Q.; Li, L.; Cheng, S.F.; Shen, W.; Lei, C.Z. Genomic signatures of selection associated with litter size trait in Jining Gray goat. Front. Genet. 2020, 11, 286. [Google Scholar] [CrossRef] [PubMed]
  5. Demars, J.; Fabre, S.; Sarry, J.; Rossetti, R.; Gilbert, H.; Persani, L.; Tosser-Klopp, G.; Mulsant, P.; Nowak, Z.; Drobik, W. Genome-wide association studies identify two novel BMP15 mutations responsible for an atypical hyperprolificacy phenotype in sheep. PLoS Genet. 2013, 9, e1003482. [Google Scholar] [CrossRef]
  6. Lai, F.N.; Zhai, H.L.; Cheng, M.; Ma, J.Y.; Cheng, S.F.; Ge, W.; Zhang, G.L.; Wang, J.J.; Zhang, R.Q.; Wang, X. Whole-genome scanning for the litter size trait associated genes and SNPs under selection in dairy goat (Capra hircus). Sci. Rep. 2016, 6, 38096. [Google Scholar] [CrossRef]
  7. Wang, K.; Liu, X.; Qi, T.; Hui, Y.; Yan, H.; Qu, L.; Lan, X.; Pan, C. Whole-genome sequencing to identify candidate genes for litter size and to uncover the variant function in goats (Capra hircus). Genomics 2021, 113, 142–150. [Google Scholar] [CrossRef]
  8. Hua, G.H.; Chen, S.L.; Ai, J.T.; Yang, L.G. None of polymorphism of ovine fecundity major genes FecB and FecX was tested in goat. Anim. Reprod. Sci. 2008, 108, 279–286. [Google Scholar] [CrossRef]
  9. Mulsant, P.; Lecerf, F.; Fabre, S.; Schibler, L.; Monget, P.; Lanneluc, I.; Pisselet, C.; Riquet, J.; Monniaux, D.; Callebaut, I.; et al. Mutation in bone morphogenetic protein receptor-IB is associated with increased ovulation rate in Booroola Mérino ewes. Proc. Natl. Acad. Sci. USA 2001, 98, 5104–5109. [Google Scholar] [CrossRef] [Green Version]
  10. Alberto, F.J.; Boyer, F.; Orozco-terWengel, P.; Streeter, I.; Servin, B.; de Villemereuil, P.; Benjelloun, B.; Librado, P.; Biscarini, F.; Colli, L.; et al. Convergent genomic signatures of domestication in sheep and goats. Nat. Commun. 2018, 9, 813. [Google Scholar] [CrossRef]
  11. Wu, D.-D.; Yang, C.-P.; Wang, M.-S.; Dong, K.-Z.; Yan, D.-W.; Hao, Z.-Q.; Fan, S.-Q.; Chu, S.-Z.; Shen, Q.-S.; Jiang, L.-P.; et al. Convergent genomic signatures of high-altitude adaptation among domestic mammals. Nat. Sci. Rev. 2019, 7, 952–963. [Google Scholar] [CrossRef] [Green Version]
  12. Li, H.; Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 2009, 25, 1754–1760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Mckenna, A.; Hanna, M.G.; Banks, E.; Sivachenko, A.; Cibulskis, K.; Kernytsky, A.M.; Garimella, K.; Altshuler, D.S.; Gabriel, S.; Daly, M. The Genome Analysis Toolkit: A MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 2010, 20, 1297–1303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Purcell, S.; Neale, B.; Todd-Brown, K.; Thomas, L.; Ferreira, M.A.R.; Bender, D.; Maller, J.; Sklar, P.; de Bakker, P.I.W.; Daly, M.J.; et al. PLINK: A tool set for whole-genome association and population-based linkage analyses. Am. J. Hum. Genet. 2007, 81, 559–575. [Google Scholar] [CrossRef] [Green Version]
  15. Danecek, P.; Auton, A.; Abecasis, G.R.; Albers, C.A.; Banks, E.; Depristo, M.A.; Handsaker, R.E.; Lunter, G.; Marth, G.T.; Sherry, S.T. The variant call format and VCFtools. Bioinformatics 2011, 27, 2156–2158. [Google Scholar] [CrossRef] [PubMed]
  16. Weir, B.S.; Cockerham, C.C. Estimating F-statistics for the analysis of population structure. Evolution 1984, 38, 1358–1370. [Google Scholar] [CrossRef]
  17. Turner, S.D. qqman: An R package for visualizing GWAS results using QQ and manhattan plots. bioRxiv 2014. [Google Scholar] [CrossRef]
  18. Wang, K.; Li, M.Y.; Hakonarson, H. ANNOVAR: Functional annotation of genetic variants from high-throughput sequencing data. Nucleic Acids Res. 2010, 38, e164. [Google Scholar] [CrossRef]
  19. Kim, D.; Langmead, B.; Salzberg, S.L. HISAT: A fast spliced aligner with low memory requirements. Nat. Methods 2015, 12, 357–360. [Google Scholar] [CrossRef] [Green Version]
  20. Kovaka, S.; Zimin, A.V.; Pertea, G.M.; Razaghi, R.; Salzberg, S.L.; Pertea, M. Transcriptome assembly from long-read RNA-seq alignments with StringTie2. Genome Biol. 2019, 20, 278. [Google Scholar] [CrossRef] [Green Version]
  21. Love, M.I.; Huber, W.; Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014, 15, 550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ferrer-Admetlla, A.; Liang, M.; Korneliussen, T.; Nielsen, R. On detecting incomplete soft or hard selective sweeps using haplotype structure. Mol. Biol. Evol. 2014, 31, 1275–1291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zhou, Y.; Zhou, B.; Pache, L.; Chang, M.; Khodabakhshi, A.H.; Tanaseichuk, O.; Benner, C.; Chanda, S.K. Metascape provides a biologist-oriented resource for the analysis of systems-level datasets. Nat. Commun. 2019, 10, 1523. [Google Scholar] [CrossRef] [PubMed]
  24. Rochus, C.M.; Jonas, E.; Johansson, A.M. Population structure of five native sheep breeds of Sweden estimated with high density SNP genotypes. BMC Genet. 2020, 21, 27. [Google Scholar] [CrossRef] [PubMed]
  25. Cao, X.; Liu, W.-P.; Cheng, L.-G.; Li, H.-J.; Wu, H.; Liu, Y.-H.; Chen, C.; Xiao, X.; Li, M.; Wang, G.-D.; et al. Whole genome analyses reveal significant convergence in obsessive-compulsive disorder between humans and dogs. Sci. Bull. 2021, 66, 187–196. [Google Scholar] [CrossRef]
  26. Kuberan, B.; Lech, M.; Borjigin, J.; Rosenberg, R.D. Light-induced 3-O-sulfotransferase expression alters pineal heparan sulfate fine structure: A surprising link to circadian rhythm. J. Biol. Chem. 2004, 279, 5053–5054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Martinez-Royo, A.; Alabart, J.L.; Sarto, P.; Serrano, M.; Lahoz, B.; Folch, J.; Calvo, J.H. Genome-wide association studies for reproductive seasonality traits in Rasa Aragonesa sheep breed. Theriogenology 2017, 99, 21–29. [Google Scholar] [CrossRef]
  28. Coster, A.; Madsen, O.; Heuven, H.C.; Dibbits, B.; Groenen, M.A.; van Arendonk, J.A.; Bovenhuis, H. The imprinted gene DIO3 is a candidate gene for litter size in pigs. PLoS ONE 2012, 7, e31825. [Google Scholar] [CrossRef] [Green Version]
  29. Klüppel, M.; Vallis, K.A.; Wrana, J.L. A high-throughput induction gene trap approach defines C4ST as a target of BMP signaling. Mech. Dev. 2002, 118, 77–89. [Google Scholar] [CrossRef]
  30. Klüppel, M.; Wight, T.N.; Chan, C.; Hinek, A.; Wrana, J.L. Maintenance of chondroitin sulfation balance by chondroitin-4-sulfotransferase 1 is required for chondrocyte development and growth factor signaling during cartilage morphogenesis. Development 2005, 132, 3989–4003. [Google Scholar] [CrossRef] [Green Version]
  31. Bhattacharyya, S.; Feferman, L.; Tobacman, J.K. Regulation of chondroitin-4-sulfotransferase (CHST11) expression by opposing effects of arylsulfatase B on BMP4 and Wnt9A. Biochim. Biophys. Acta 2015, 1849, 342–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Oliveira-Ferrer, L.; Heßling, A.; Trillsch, F.; Mahner, S.; Milde-Langosch, K. Prognostic impact of chondroitin-4-sulfotransferase CHST11 in ovarian cancer. Tumor Biol. 2015, 36, 9023–9030. [Google Scholar] [CrossRef] [PubMed]
  33. Farkas, S.A.; Sorbe, B.G.; Nilsson, T.K. Epigenetic changes as prognostic predictors in endometrial carcinomas. Epigenetics 2017, 12, 19–26. [Google Scholar] [CrossRef] [Green Version]
  34. Wu, P.; Wang, K.; Zhou, J.; Chen, D.; Yang, Q.; Yang, X.; Liu, Y.; Feng, B.; Jiang, A.; Shen, L.; et al. GWAS on imputed whole-genome resequencing from genotyping-by-sequencing data for farrowing interval of different parities in pigs. Front. Genet. 2019, 10, 1012. [Google Scholar] [CrossRef]
  35. Forsyth, R.; Gunay-Aygun, M. Bardet-Biedl syndrome overview. In GeneReviews(®); Adam, M.P., Ardinger, H.H., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Stephens, K., Amemiya, A., Eds.; University of Washington: Seattle, WA, USA, 1993. [Google Scholar]
  36. Weihbrecht, K.; Goar, W.A.; Carter, C.S.; Sheffield, V.C.; Seo, S. Genotypic and phenotypic characterization of the Sdccag8Tn(sb-Tyr)2161B.CA1C2Ove mouse model. PLoS ONE 2018, 13, e0192755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Airik, R.; Schueler, M.; Airik, M.; Cho, J.; Ulanowicz, K.A.; Porath, J.D.; Hurd, T.W.; Bekker-Jensen, S.; Schrøder, J.M.; Andersen, J.S.; et al. SDCCAG8 interacts with RAB effector proteins RABEP2 and ERC1 and is required for Hedgehog signaling. PLoS ONE 2016, 11, e0156081. [Google Scholar] [CrossRef] [Green Version]
  38. Zhuang, Z.; Ding, R.; Peng, L.; Wu, J.; Ye, Y.; Zhou, S.; Wang, X.; Quan, J.; Zheng, E.; Cai, G.; et al. Genome-wide association analyses identify known and novel loci for teat number in Duroc pigs using single-locus and multi-locus models. BMC Genom. 2020, 21, 344. [Google Scholar] [CrossRef] [PubMed]
  39. Wu, M.; Chen, G.; Li, Y.-P. TGF-β and BMP signaling in osteoblast, skeletal development, and bone formation, homeostasis and disease. Bone Res. 2016, 4, 16009. [Google Scholar] [CrossRef]
  40. Abdoli, R.; Zamani, P.; Mirhoseini, S.Z.; Ghavi Hossein-Zadeh, N.; Nadri, S. A review on prolificacy genes in sheep. Reprod. Domest. Anim. 2016, 51, 631–637. [Google Scholar] [CrossRef] [PubMed]
  41. Zhang, X.; Huang, L.; Wu, T.; Feng, Y.; Ding, Y.; Ye, P.; Yin, Z. Transcriptomic analysis of ovaries from pigs with high and low litter size. PLoS ONE 2015, 10, e0139514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Pokharel, K.; Peippo, J.; Honkatukia, M.; Seppala, A.; Rautiainen, J.; Ghanem, N.; Hamama, T.M.; Crowe, M.A.; Andersson, M.; Li, M.H.; et al. Integrated ovarian mRNA and miRNA transcriptome profiling characterizes the genetic basis of prolificacy traits in sheep (Ovis aries). BMC Genom. 2018, 19, 104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. He, J.; Zhang, B.; Chu, M.; Wang, P.; Feng, T.; Cao, G.; Di, R.; Fang, L.; Huang, D.; Tang, Q. Polymorphism of insulin-like growth factor 1 gene and its association with litter size in Small Tail Han sheep. Mol. Biol. Rep. 2012, 39, 9801–9807. [Google Scholar] [CrossRef] [PubMed]
  44. Thomas, N.; Venkatachalapathy, T.; Aravindakshan, T.; Raghavan, K.C. Molecular cloning, SNP detection and association analysis of 5′ flanking region of the goat IGF1 gene with prolificacy. Anim. Reprod. Sci. 2016, 167, 8–15. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, P.-Q.; Tan, Y.; Zhang, B.-Y.; Chu, M.-X.; Deng, L.-M.; Fan, Q.; Liu, C.-X. DNA polymorphisms of 5′-flanking region of insulin-like growth factor 1 gene and their association with reproduction traits in goats. Agric. Sci. China 2011, 10, 1609–1617. [Google Scholar] [CrossRef]
  46. Hatzirodos, N.; Hummitzsch, K.; Irving-Rodgers, H.F.; Rodgers, R.J. Transcriptome profiling of the theca interna in transition from small to large antral ovarian follicles. PLoS ONE 2014, 9, e97489. [Google Scholar] [CrossRef] [Green Version]
  47. Kwon, S.G.; Hwang, J.H.; Park, D.H.; Kim, T.W.; Kang, D.G.; Kang, K.H.; Kim, I.-S.; Park, H.C.; Na, C.-S.; Ha, J. Identification of differentially expressed genes associated with litter size in Berkshire pig placenta. PLoS ONE 2016, 11, e0153311. [Google Scholar] [CrossRef]
  48. Oh, J.W.; Kim, S.K.; Cho, K.C.; Kim, M.S.; Suh, C.S.; Lee, J.R.; Kim, K.P. Proteomic analysis of human follicular fluid in poor ovarian responders during in vitro fertilization. Proteomics 2017, 17, 1600333. [Google Scholar] [CrossRef]
  49. Wang, Y.; Ding, X.; Tan, Z.; Ning, C.; Xing, K.; Yang, T.; Pan, Y.; Sun, D.; Wang, C. Genome-wide association study of piglet uniformity and farrowing interval. Front. Genet. 2017, 8, 194. [Google Scholar] [CrossRef] [Green Version]
  50. Padua, M.B.; Hansen, P.J. Evolution and function of the uterine serpins (SERPINA14). Am. J. Reprod. Immunol. 2010, 64, 265–274. [Google Scholar] [CrossRef]
  51. Ulbrich, S.E.; Frohlich, T.; Schulke, K.; Englberger, E.; Waldschmitt, N.; Arnold, G.J.; Reichenbach, H.D.; Reichenbach, M.; Wolf, E.; Meyer, H.H.D.; et al. Evidence for estrogen-dependent uterine serpin (SERPINA14) expression during estrus in the bovine endometrial glandular epithelium and lumen. Biol. Reprod. 2009, 81, 795–805. [Google Scholar] [CrossRef] [Green Version]
  52. Rohit, K.; Ramteke, P.W.; Amar, N.; Singh, S.P. Role of candidate genes in regulation of embryonic survival and maternal recognition of pregnancy in farm animals. Vet. World 2013, 6, 280–284. [Google Scholar] [CrossRef] [Green Version]
  53. Jang, J.-S.; Lee, J.-H.; Jung, N.-C.; Choi, S.-Y.; Park, S.-Y.; Yoo, J.-Y.; Song, J.-Y.; Seo, H.G.; Lee, H.S.; Lim, D.-S. Rsad2 is necessary for mouse dendritic cell maturation via the IRF7-mediated signaling pathway. Cell Death Dis. 2018, 9, 823. [Google Scholar] [CrossRef] [PubMed]
  54. Cheng, L.; Xiang, M.; Hu, X.; Yu, J.; Xia, Y.; Tao, B.; Zhao, S. Duplex quantitative polymerase chain reaction of ISG15 and RSAD2 increases accuracy of early pregnancy diagnosis in dairy cows. Ann. Anim. Sci. 2019, 19, 383–401. [Google Scholar] [CrossRef] [Green Version]
  55. Zhang, J.; Liu, Y.; Yao, W.; Li, Q.; Liu, H.; Pan, Z. Initiation of follicular atresia: Gene networks during early atresia in pig ovaries. Reproduction 2018, 156, 23–33. [Google Scholar] [CrossRef] [Green Version]
  56. Livera, G.; Xie, F.; Garcia, M.A.; Jaiswal, B.; Chen, J.; Law, E.; Storm, D.R.; Conti, M. Inactivation of the mouse adenylyl cyclase 3 gene disrupts male fertility and spermatozoon function. Mol. Endocrinol. 2005, 19, 1277–1290. [Google Scholar] [CrossRef] [Green Version]
  57. Zhang, Y.; Li, F.; Feng, X.; Yang, H.; Zhu, A.; Pang, J.; Han, L.; Zhang, T.; Yao, X.; Wang, F. Genome-wide analysis of DNA methylation profiles on sheep ovaries associated with prolificacy using whole-genome Bisulfite sequencing. BMC Genom. 2017, 18, 759. [Google Scholar] [CrossRef]
  58. Sironen, A.I.; Uimari, P.; Serenius, T.; Mote, B.; Rothschild, M.; Vilkki, J. Effect of polymorphisms in candidate genes on reproduction traits in Finnish pig populations. J. Anim. Sci. 2010, 88, 821–827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Manhattan plots for fixation index (Fst)(YSD) and root mean square of fixation index (Fst_RMS). The red lines indicate the cut-offs of Fst(YSD) (0.05, top 0.8%) and Fst_RMS (0.3, top 0.7%).
Figure 1. Manhattan plots for fixation index (Fst)(YSD) and root mean square of fixation index (Fst_RMS). The red lines indicate the cut-offs of Fst(YSD) (0.05, top 0.8%) and Fst_RMS (0.3, top 0.7%).
Genes 12 00480 g001
Figure 2. Venn diagrams of genetic convergence of prolificacy between goats and sheep. The top panels denote the convergence at the single gene level (positively selected genes and differentially expressed genes of ovaries), whereas the convergence at the gene set level is represented as Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) Ontology (ko) terms in the bottom panels; only biological processes (BP) were considered in GO analysis.
Figure 2. Venn diagrams of genetic convergence of prolificacy between goats and sheep. The top panels denote the convergence at the single gene level (positively selected genes and differentially expressed genes of ovaries), whereas the convergence at the gene set level is represented as Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) Ontology (ko) terms in the bottom panels; only biological processes (BP) were considered in GO analysis.
Genes 12 00480 g002
Figure 3. Volcano plots of differentially expressed genes in the ovaries of Yunshang black goats (A,B) and Small Tail Han sheep (C,D). (A,C) indicate luteal phases, and (B,D) denote follicular phases. Compared with the low-yield groups, points in red, blue and green are the down-regulated, up-regulated, and non-differentially expressed genes in high-yield groups, respectively. The dashed lines are thresholds of 1.30 (−lg(0.05)).
Figure 3. Volcano plots of differentially expressed genes in the ovaries of Yunshang black goats (A,B) and Small Tail Han sheep (C,D). (A,C) indicate luteal phases, and (B,D) denote follicular phases. Compared with the low-yield groups, points in red, blue and green are the down-regulated, up-regulated, and non-differentially expressed genes in high-yield groups, respectively. The dashed lines are thresholds of 1.30 (−lg(0.05)).
Genes 12 00480 g003
Figure 4. Differentially expressed genes of ovaries shared by sheep and goats. FP and LP denote follicular and luteal phases, respectively.
Figure 4. Differentially expressed genes of ovaries shared by sheep and goats. FP and LP denote follicular and luteal phases, respectively.
Genes 12 00480 g004
Table 1. Phenotypes of does and ewes used in this study.
Table 1. Phenotypes of does and ewes used in this study.
Population 1Average Litter Size (Mean ± SD)p-Value 2
Low-Yield GroupHigh-Yield Group
YSD1.32 ± 0.19 (n = 20)3.00 ± 0.38 (n = 20)5.16 × 10−8
YSR1.75 ± 0.26 (n = 10)3.30 ± 0.35 (n = 10)1.21 × 10−4
FIN1.66 ± 0.33 (n = 9)3.59 ± 0.41 (n = 10)2.76 × 10−4
ROM1.53 ± 0.25 (n = 9)3.13 ± 0.17 (n = 10)2.01 × 10−4
STH1.00 ± 0.00 (n = 6)2.89 ± 0.17 (n = 6)2.22 × 10−3
1 YSD and YSR indicate Yunshang black goats used for DNA sequencing and ovarian mRNA sequencing, respectively. FIN: Finnsheep; ROM: Romanov sheep; STH: Small Tailed Han sheep. 2 Mann–Whitney U test. The number in the parentheses is sample size.
Table 2. Pathways shared by positively selected genes of sheep and goats.
Table 2. Pathways shared by positively selected genes of sheep and goats.
TermName−lg(p)Gene
SheepGoatSheepGoat
hsa04012ErbB signaling pathway3.922.18CBL PIK3CA MAPK10 KCNH8 NRG4CAMK2B GRB2 MAPK10 SRC AKT3
ko04380Osteoclast differentiation3.272.89MAPK14 NFATC2 PIK3CA MAPK10 TGFB2FCGR2A FCGR2B FCGR3A FCGR3B GRB2 MAPK10 TEC AKT3
hsa04380Osteoclast differentiation3.142.72MAPK14 NFATC2 PIK3CA MAPK10 TGFB2FCGR2A FCGR2B FCGR3A FCGR3B GRB2 MAPK10 TEC AKT3 RHBDF2
ko04012ErbB signaling pathway3.012.34CBL PIK3CA MAPK10 NRG4CAMK2B GRB2 MAPK10 SRC AKT3
hsa04926Relaxin signaling pathway2.292.13MAPK14 PIK3CA MAPK10 CREB5ADCY3 GNG5 GRB2 MMP9 MAPK10 SRC AKT3
Table 3. Terms shared by differentially expressed genes of sheep and goats.
Table 3. Terms shared by differentially expressed genes of sheep and goats.
TermName−lg(p)Gene
GoatSheepGoatSheep
Protein kinase B signalingGO:00434912.992.58TCF7L2 IGFBP5 RPS6KB2 MYOC LOC102175889IGF1 KIT LOX TRK2 PKHD1
Antigen processing and presentationko046122.083.86PDIA3 CREB1 CALR NFYCHLA-A HLA-BLA-C HLA-F
Regulation of viral genome replicationGO:00450692.256.15RSAD2 VAPAIFIT1 MX1 ISG15 RSAD2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tao, L.; He, X.; Jiang, Y.; Liu, Y.; Ouyang, Y.; Shen, Y.; Hong, Q.; Chu, M. Genome-Wide Analyses Reveal Genetic Convergence of Prolificacy between Goats and Sheep. Genes 2021, 12, 480. https://doi.org/10.3390/genes12040480

AMA Style

Tao L, He X, Jiang Y, Liu Y, Ouyang Y, Shen Y, Hong Q, Chu M. Genome-Wide Analyses Reveal Genetic Convergence of Prolificacy between Goats and Sheep. Genes. 2021; 12(4):480. https://doi.org/10.3390/genes12040480

Chicago/Turabian Style

Tao, Lin, Xiaoyun He, Yanting Jiang, Yufang Liu, Yina Ouyang, Yezhen Shen, Qionghua Hong, and Mingxing Chu. 2021. "Genome-Wide Analyses Reveal Genetic Convergence of Prolificacy between Goats and Sheep" Genes 12, no. 4: 480. https://doi.org/10.3390/genes12040480

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop