Abstract

Skin color is a highly heritable human trait, and global variation in skin pigmentation has been shaped by natural selection, migration and admixture. Ethnically diverse African populations harbor extremely high levels of genetic and phenotypic diversity, and skin pigmentation varies widely across Africa. Recent genome-wide genetic studies of skin pigmentation in African populations have advanced our understanding of pigmentation biology and human evolutionary history. For example, novel roles in skin pigmentation for loci near MFSD12 and DDB1 have recently been identified in African populations. However, due to an underrepresentation of Africans in human genetic studies, there is still much to learn about the evolutionary genetics of skin pigmentation. Here, we summarize recent progress in skin pigmentation genetics in Africans and discuss the importance of including more ethnically diverse African populations in future genetic studies. In addition, we discuss methods for functional validation of adaptive variants related to skin pigmentation.

Introduction

The hypothesis that modern humans originated in Africa was first raised based on the fossil record (1,2). The ‘Out-of-Africa’ theory, which posits a single Homo sapiens origin in Africa ~300 kya, has since been confirmed at the molecular level based on mitochondrial and nuclear DNA sequencing studies (3–6). Around 50–80 kya, a small number of humans migrated out of Africa and subsequently populated the rest of the globe (3,4). During this migration, humans encountered diverse geographical, environmental and climatic conditions, which shaped many adaptive traits, including skin pigmentation (7). Variation in skin pigmentation is, thus, largely attributable to the forces of natural selection and is almost perfectly correlated with ultraviolet radiation (UVR) globally (7,8). The most widely held hypothesis is that darkly pigmented skin protects against UV-induced degradation of folate in high UVR environments, while lighter skin facilitates the production of vitamin D at higher latitudes (9,10). These biological processes could directly affect fitness; folate deficiencies lead to a diversity of birth defects and fertility complications (11), while vitamin D plays a critical role in bone health and reproductive physiology (12).

During the migration out of Africa to higher latitudes, modern humans encountered lower and more variable UVR environments, and depigmented skin evolved independently multiple times (13). As a result, genetic variants that are associated with lighter skin, including in genes SLC24A5, SLC45A2, MC1R, TYR, TYRP1 and OCA2, display strong statistical signals of positive selection in European and Asian populations (13–22). However, this is not to suggest that the ancestral state of human skin pigmentation was dark. Recent genomic studies show that the ancestral alleles of many predicted functional pigmentation variants in Africa are associated with lighter skin, suggesting our human ancestors may have had light or moderately pigmented skin (23). Combined with the fact that our closest evolutionary relatives, chimpanzees, have light skin (24), these results suggest that dark skin may be a derived trait in the Homo genus, and that both light and dark skin have continued to evolve over hominid history (23,25).

The majority of studies examining the genetics of skin pigmentation have been conducted in European and Asian populations, and African populations are vastly underrepresented (26,27). This is despite the fact that modern Africans harbor very high levels of genetic and phenotypic diversity (6), and skin pigmentation is no exception. For example, the KhoeSan hunter-gatherers in Botswana have relatively light skin, while Nilo-Saharan-speaking populations from East Africa have some of the darkest pigmented skin on Earth (23,28). Further, genetic variants associated with pigmentation in European and Asian populations explain only a small amount of the phenotypic variation in Africa, suggesting the genetic architecture underlying skin pigmentation differs between Africans and Eurasians (23,28).

In this review, we summarize current knowledge of the genes and variants influencing human skin pigmentation, with a particular focus on African populations. We highlight recent research on the evolutionary genetics of African skin pigmentation and suggest how best to utilize experimental methods to identify and functionally validate novel pigmentation loci that may be under selection. Finally, we argue that the inclusion of more ethnically, phenotypically and geographically diverse African populations will be essential for a full understanding of the evolutionary genetics of human skin pigmentation.

Human Skin Pigmentation Genes

Human skin color is determined by the composition, abundance and distribution of melanin pigments, which are biopolymers derived from tyrosine (29). Melanins come in multiple forms: black-brown eumelanin and yellow-red pheomelanin. Melanin is synthesized and stored in lysosome-like organelles called melanosomes, which are generated in melanocytes and transferred to surrounding keratinocytes (30). Skin melanocytes are differentiated from melanoblasts, which originate from neural crest cells during embryonic development (31,32).

Skin color is regulated by genes underlying the development of melanocytes and melanosome biogenesis. Mice and other model organisms have been used to validate the function of genes associated with human pigmentation. A recent review summarized a cross-species list of 650 genes with verified pigmentation phenotypes in humans, mice or zebrafish (33). The International Mouse Phenotyping Consortium has identified 224 genes associated with pigmentation phenotypes (34). Among them, SOX10, PAX3, MITF, EDN3, EDNRB, KIT and KITLG are key genes that regulate the proliferation, migration and differentiation of melanocyte precursor cells (32,35). TYR, TYRP1, DCT, OCA2, SLC24A5, SLC45A2, MC1R, ASIP, GPR143 and PMEL are well-characterized genes involved in the synthesis of melanin in melanocytes (29,30,36). Although here we mainly focus on the genetics and biology of skin pigmentation, it is important to note that many skin pigmentation genes (e.g. OCA2, SLC45A2, KITLG) also affect hair or eye color (37,38).

Searching for Signatures of Skin Pigmentation Evolution

An important goal of evolutionary genetics is to identify the variants and genes underlying phenotypic evolution and local adaptation (39,40). The advent of microarray and high throughput sequencing technologies now enables researchers to perform genome-wide association studies for complex human traits, which have been successful in identifying multiple genes associated with human skin pigmentation (27). These methods, coupled with genome-wide scans of natural selection, can provide novel insights into the genetic mechanisms underlying the evolution of skin pigmentation across populations (Fig. 1) (41,42). For example, allele frequency shifts at trait-associated variants between populations are indicators of local adaptation and positive selection (43–46).

Methods of detecting adaptive variants related to skin pigmentation. (1) Genome-wide selection scans identify genomic signatures of natural selection, and some are near pigmentation genes. (2) GWAS localizes variants associated with pigmentation. (3) Overlapping GWAS variants with signatures of natural selection reveals candidate adaptive pigmentation loci.
Figure 1

Methods of detecting adaptive variants related to skin pigmentation. (1) Genome-wide selection scans identify genomic signatures of natural selection, and some are near pigmentation genes. (2) GWAS localizes variants associated with pigmentation. (3) Overlapping GWAS variants with signatures of natural selection reveals candidate adaptive pigmentation loci.

Natural selection has shaped the frequency of genetic variants in human populations, and many statistical methods have been developed to identify its footprints in the genome (Fig. 1). Several reviews have described the methods for detecting selection signatures extensively (47–50), so we briefly mention the widely used techniques and their principles. These methods are mainly based on changes in the site frequency spectrum (Tajima’s D (51), Fay and Wu’s H (52)), changes in linkage disequilibrium around a positively selected site (LRH (53), LDD (54), IBD (55), iHS (56), Rsb (57), nSL (58)), genetic differentiation between populations (XP-CLR (45), FST (46), LSBL (59), PBS (60), XP-EHH (61), GRoSS (62)), their combinations or derivatives (CLR (63), CMS (64,65)) and changes in the density of singleton mutations (SDS) (66). Many of these methods have been used to detect significant selection signals near pigmentation genes in European, East Asian (13,15–20,67) and African populations (discussed below). Although many variants found to be associated with pigmentation from GWAS or identified from selection scans have been functionally validated, a large proportion of these variants have currently unknown functions (27).

Evolutionary genetics of skin pigmentation in Africans

The majority of evolutionary genetic studies on skin pigmentation have been conducted in European and Asian populations (68,69). However, some recent studies have examined pigmentation in African populations, providing novel insights into the evolution of skin color. In this section, we summarize these African studies on a gene-by-gene basis.

SLC24A5

SLC24A5 encodes a cation exchanger in melanosomes and was first confirmed to affect skin pigmentation in zebrafish (21). A derived, nonsynonymous mutation (rs1426654 (A), Ala111Thr) in SLC24A5 associated with light skin color has swept to near fixation in Europeans due to positive selection (13,22,70,71). Recent studies show that this SLC24A5 allele is also associated with light skin in Africans. Crawford et al. (23) performed a GWAS on 1570 ethnically diverse Africans that spanned the full spectrum of skin pigment variation in Africa and found the strongest association at SLC24A5. The authors show that rs1426654 is common in East African populations with high levels of Afroasiatic ancestry (Table 1), and that rs1426654 likely introgressed into these populations from a Eurasian source >5 kya. Further, SLC24A5 likely experienced positive selection in East Africa after this admixture event. The authors also find that rs1426654 is at moderate frequency (5–12%) in the KhoeSan from Botswana, who have substantially lighter skin than equatorial Africans (Table 1).

Table 1

Annotations and population allele frequencies of pigmentation-associated variants identified in African populations

Allele Frequency
SNPAlleleGeneFunctionAFR
1KG
EUR
1KG
SAS
1KG
EAS
1KG
AMR
1KG
Botswana
Khoesan*
Ethiopia
Afroasiatic*
Ethiopia
Nilo-Saharan*
rs1426654ASLC24A5Ala 111 Thr0.0740.9970.6850.0120.5890.1240.4410.127
rs12913832AHERC2Intronic0.9720.3640.9290.9980.7980.9950.9791
rs1800404TOCA2Synonymous0.1270.7860.3140.3860.5790.7660.320.069
rs4932620THERC2Intronic0.0850.0160.1650.0170.0140.0270.2760.48
rs6497271GHERC2Intronic0.3740.9840.8370.9790.9470.7730.5720.196
rs7948623TDDB1Intergenic0.2660.0060.1270.0030.030.0580.4170.716
rs2240751GMFSD12Tyr 182 His0.0020.010.0070.2690.16700.0010
rs56203814TMFSD12Synonymous0.2080.008000.0170.0240.2730.51
rs10424065TMFSD12Intronic0.2980.00900.0010.0220.0440.2560.55
Allele Frequency
SNPAlleleGeneFunctionAFR
1KG
EUR
1KG
SAS
1KG
EAS
1KG
AMR
1KG
Botswana
Khoesan*
Ethiopia
Afroasiatic*
Ethiopia
Nilo-Saharan*
rs1426654ASLC24A5Ala 111 Thr0.0740.9970.6850.0120.5890.1240.4410.127
rs12913832AHERC2Intronic0.9720.3640.9290.9980.7980.9950.9791
rs1800404TOCA2Synonymous0.1270.7860.3140.3860.5790.7660.320.069
rs4932620THERC2Intronic0.0850.0160.1650.0170.0140.0270.2760.48
rs6497271GHERC2Intronic0.3740.9840.8370.9790.9470.7730.5720.196
rs7948623TDDB1Intergenic0.2660.0060.1270.0030.030.0580.4170.716
rs2240751GMFSD12Tyr 182 His0.0020.010.0070.2690.16700.0010
rs56203814TMFSD12Synonymous0.2080.008000.0170.0240.2730.51
rs10424065TMFSD12Intronic0.2980.00900.0010.0220.0440.2560.55

AFR (African), EUR (European), SAS (South Asian), EAS (East Asian) and AMR (Admixed American) refer to super population codes from the 1000 genomes (1KG) project. Allele frequencies for populations marked with an asterisk were calculated from 295 Botswana KhoeSan, 451 Ethiopian Afroasiatic and 51 Ethiopian Nilo-Saharan individuals as described in Crawford et al. (23)

Table 1

Annotations and population allele frequencies of pigmentation-associated variants identified in African populations

Allele Frequency
SNPAlleleGeneFunctionAFR
1KG
EUR
1KG
SAS
1KG
EAS
1KG
AMR
1KG
Botswana
Khoesan*
Ethiopia
Afroasiatic*
Ethiopia
Nilo-Saharan*
rs1426654ASLC24A5Ala 111 Thr0.0740.9970.6850.0120.5890.1240.4410.127
rs12913832AHERC2Intronic0.9720.3640.9290.9980.7980.9950.9791
rs1800404TOCA2Synonymous0.1270.7860.3140.3860.5790.7660.320.069
rs4932620THERC2Intronic0.0850.0160.1650.0170.0140.0270.2760.48
rs6497271GHERC2Intronic0.3740.9840.8370.9790.9470.7730.5720.196
rs7948623TDDB1Intergenic0.2660.0060.1270.0030.030.0580.4170.716
rs2240751GMFSD12Tyr 182 His0.0020.010.0070.2690.16700.0010
rs56203814TMFSD12Synonymous0.2080.008000.0170.0240.2730.51
rs10424065TMFSD12Intronic0.2980.00900.0010.0220.0440.2560.55
Allele Frequency
SNPAlleleGeneFunctionAFR
1KG
EUR
1KG
SAS
1KG
EAS
1KG
AMR
1KG
Botswana
Khoesan*
Ethiopia
Afroasiatic*
Ethiopia
Nilo-Saharan*
rs1426654ASLC24A5Ala 111 Thr0.0740.9970.6850.0120.5890.1240.4410.127
rs12913832AHERC2Intronic0.9720.3640.9290.9980.7980.9950.9791
rs1800404TOCA2Synonymous0.1270.7860.3140.3860.5790.7660.320.069
rs4932620THERC2Intronic0.0850.0160.1650.0170.0140.0270.2760.48
rs6497271GHERC2Intronic0.3740.9840.8370.9790.9470.7730.5720.196
rs7948623TDDB1Intergenic0.2660.0060.1270.0030.030.0580.4170.716
rs2240751GMFSD12Tyr 182 His0.0020.010.0070.2690.16700.0010
rs56203814TMFSD12Synonymous0.2080.008000.0170.0240.2730.51
rs10424065TMFSD12Intronic0.2980.00900.0010.0220.0440.2560.55

AFR (African), EUR (European), SAS (South Asian), EAS (East Asian) and AMR (Admixed American) refer to super population codes from the 1000 genomes (1KG) project. Allele frequencies for populations marked with an asterisk were calculated from 295 Botswana KhoeSan, 451 Ethiopian Afroasiatic and 51 Ethiopian Nilo-Saharan individuals as described in Crawford et al. (23)

Martin et al. (28) also examined the genetic architecture of skin pigmentation in two KhoeSan-speaking populations: the ‡Khomani San and Nama. By examining detailed skin color measurements in 465 KhoeSan people, the authors found that previously known pigmentation-associated loci explain only a small percentage of the overall skin color variation, suggesting many pigmentation loci remain to be discovered. Despite this, they identified a significant association at SLC24A5 and reported the frequency of rs1426654 at 40% in the combined Nama and ‡Khomani dataset, which has ~11% European admixture. Lin et al. (72) further examined the evolution of SLC24A5 in the Nama and ‡Khomani populations by performing haplotype analysis and demographic modeling, finding that SLC24A5 experienced strong positive selection after being introduced from a non-African source within the last ~2k years. Further, rs1426654 likely arrived in the KhoeSan indirectly through a southern pastoralist migration from East Africa (72).

OCA2/HERC2

The OCA2 gene impacts skin pigmentation by regulating melanosome pH (73,74), and mutations in this gene cause oculocutaneous albinism type II (OCA2) (75). HERC2 is adjacent to OCA2 on chromosome 15, and although HERC2 does not directly impact pigmentation, regulatory elements in HERC2 can influence expression of OCA2 (76). For example, multiple studies report that the intronic HERC2 mutation rs12913832 is associated with skin (77–79), hair (37,80) and eye (81–83) pigmentation in Eurasians (84) and is located in a melanocyte-specific enhancer that regulates OCA2 expression (76). Within Africans, several novel variants in this region associated with skin pigmentation have recently been described. The derived allele of a synonymous variant in OCA2 (rs1800404 (T)), associated with light pigmentation, is at high frequency (>70%) in Europeans and the KhoeSan from Botswana (Table 1) (13,23). This variant is associated with alternative splicing of OCA2 at exon 10, which may impact African pigmentation by influencing the amount of functional OCA2 protein produced. This OCA2 region also displays a significantly elevated Tajima’s D value in both Africans and Eurasians, suggesting long-term balancing selection at this locus. Finally, two novel noncoding variants (rs4932620 and rs6497271) in the introns of HERC2 were also recently identified by Crawford et al. (23). They found that the derived allele of rs4932620 (T) is significantly associated with dark skin pigmentation and at highest frequency in Ethiopian Nilo-Saharan populations, while the derived allele of rs6497271 (G) is associated with light skin pigmentation and is common in the KhoeSan from Botswana (Table 1) (23).

D‌DB1

The damage-specific DNA binding protein (DDB1) gene functions in DNA repair after UV-induced damage and is associated with the disease xeroderma pigmentosum complementation group E (85). This gene, which underlies the fruit pigmentation phenotype in tomatoes (86), is located in a broad genomic region that was found to significantly associate with skin color in a GWAS in an African-European admixed population (87). Variants in and around this gene were also significantly associated with African skin pigmentation in a GWAS of ethnically diverse African people from Tanzania, Botswana and Ethiopia (23). Specifically, three noncoding genome-wide significant variants (rs7948623, rs1377457, rs148172827) were shown to change the activity of enhancers near the DDB1 gene in melanoma cell lines. The derived allele of rs7948623 (T), associated with dark pigmentation, is at highest global frequency in East African populations with Nilo-Saharan ancestry, who inhabit high UVR environments and have darkly pigmented skin. This allele is also at moderate to high frequency in South Asians and Austro-Melanesians with dark skin, while the light pigmentation allele is nearly fixed in Europeans (Table 1). Finally, the FST and Tajima’s D statistics show that DDB1 is under differential selection in Africans versus non-Africans. Specifically, the light pigment associated alleles have undergone a strong selective sweep in European and Asian populations (23).

MFSD12

The above-mentioned Crawford et al. (23) study also identified a significant association with skin color at MFSD12 in Africans, a gene never previously characterized. Specifically, the derived alleles at the synonymous variant rs56203814 (T) and the intronic variant rs10424065 (T) significantly associate with dark pigmentation and are at highest frequency in East Africans with Nilo-Saharan ancestry (Table 1). Silencing MFSD12 expression in skin melanocytes from the mouse resulted in an increase in eumelanin production and darker pigmentation. This observation correlates with the finding that melanocyte MFSD12 expression levels are lower in people of African ancestry. Further, CRISPR-Cas9 knockouts of MFSD12 in Agouti mice results in decreased pheomelanin production (23). Recently, MFSD12 was shown to affect pigmentation by mediating the import of cysteine, a key precursor in pheomelanin production, into melanosomes (88). Increased MFSD12 expression also promotes melanoma cell proliferation and may be an important therapeutic target for melanoma (89). The fact that completely novel, functionally validated skin pigmentation loci were identified by Crawford et al. (23) using sample sizes that are small by today’s standards reinforces the need to include more ethnically diverse African populations in modern genetic studies. Variants at MFSD12 also have highly divergent allele frequencies between East Africans and Europeans as measured by FST, consistent with a signature of local adaptation (23). Finally, a novel nonsynonymous variant in MFSD12 (rs2240751 (G) Tyr182His) was recently reported to be associated with skin pigmentation in GWAS of Latin American (90), East Asian (91) and European populations (Table 1) (92).

MC1R

Some early studies of skin pigmentation genes focus on the melanocortin 1 receptor (MC1R), which controls the type of melanin produced (eumelanin or pheomelanin) by melanocytes (93). Harding et al. (94) examined nucleotide variation at MC1R in European, Asian and African populations, finding that nonsynonymous variants are largely absent in African populations. In contrast, European and Asian populations harbor more diversity at MC1R, with many segregating nonsynonymous variants present (94). These data suggest strong purifying selection acting at MC1R within Africa, where any deviation from the eumelanin-producing form of the gene, which leads to darker pigmentation, is strongly selected against. This functional constraint has been relaxed in other global populations, where UVR is lower and lighter skin is adaptive. However, another study identified a small number of nonsynonymous MC1R mutations present in lightly pigmented people from Southern Africa, suggesting this functional constraint may also have been relaxed in low UVR regions of Africa (95). Several recent GWAS confirmed that the MC1R locus is significantly associated with skin (96–98) or hair color (37,38,98,99) in European populations. No GWAS have reported significant associations with pigmentation at the MC1R the locus in African populations thus far, likely due to a lack of variation at the locus in Africans.

TYRP1 and Other Genes

TYRP1 was the first gene found to be associated with oculocutaneous albinism in Africans, and the rs387906560 frameshift mutation (Lys368SerfsTer17) in TYRP1 causes oculocutaneous albinism 3 (OCA3 or rufous albinism) (100–102). This mutation is at a higher frequency in Africans (0.5%, GnomAD) than non-Africans (0.017%, GnomAD). TYRP1 shows a signature of strong positive selection in a Senegalese population, as indicated by a significantly negative Tajima’s D value (103). In addition, tests of extended haplotype homozygosity (EHH), which measure levels of linkage disequilibrium around a positively selected locus, find three pigmentation loci (LYST, TP53BP1, RAD50) showing signatures of positive selection in the Yoruban population from Nigeria (104). Another analysis of an African American population used the composite likelihood ratio (CLR) test to identify signals of recent positive selection near KITLG and OCA2 (63,105).

From association to causation: fine-mapping and functional validation

Signals of selection do not provide information about which trait is under selection or what the driving force of selection is (17,106). Similarly, although GWAS hits are statistically associated with a trait, their biological effects are unclear. Most genome-wide selection scans report top loci based on an empirical distribution and assume that targets of selection are outliers in the distribution of summary statistics, which is not always true. Thus, these empirical statistics (including FST, Tajima’s D and iHS) have high false-positive and false-negative rates (106,107) and cannot accurately pinpoint the causal variants. Prioritizing candidate causal variants in GWAS can be accomplished by performing computational fine-mapping based on summary statistics (108,109). However, to identify causal variants from GWAS and selection scans, it is necessary to evaluate and validate their biological functions experimentally (107,110,111).

Here, we use skin pigmentation as an example to illustrate how to prioritize and validate causal variants from GWAS or selection scans (Fig. 2). It is well known that melanocytes are the key cell types involved in skin pigmentation, and epidermal keratinocytes and fibroblasts play a secondary role (112,113). To study the genetic basis of human skin pigmentation, there are two levels of functional validation: (1) testing the role of candidate genes on skin pigmentation and (2) validating the effects of allelic variants on skin pigmentation. Testing the function of pigmentation-associated genes or coding variants is relatively straightforward. One could use CRISPR/Cas9-based tools to edit the coding region of a gene directly and then examine the effect on the pigmentation phenotype in cell lines or animal models. In addition, comparative genetics offers a powerful approach for predicting the function of genes or variants related to human skin pigmentation. For example, coding variants in MFSD12 (discussed above) are associated with coat color in dogs (114) and horses (115). Similarly, genetic studies of pigmentation in stickleback fish helped elucidate the function of a conserved regulatory variant in the human KITLG gene (116). However, validating the function of noncoding variants is more challenging, and comparative genetics are less informative for variants located in unconserved noncoding regions.

An integrative functional genomics approach for evolutionary studies. During the dispersal of modern humans, admixture between different populations, archaic introgression and natural selection have shaped the allele frequencies in human populations. The functional alleles (arising de novo or from standing variation) influence an individual’s phenotype (morphological or physiological) and fitness by modulating gene expression (noncoding variants) or changing the protein function (coding variants). The individuals with adaptive variants have a higher probability of survival or reproduction in certain environments (adaptation). The selection signals could be detected by genomic scans of selection and the causal variants could be identified and validated with experimental methods. GWAS could link the potential causal variants with phenotypes and eQTLs could be used to link variants with their target genes. However, both GWAS and eQTL analyses only indicate that association (dashed lines) and experimental validations are necessary for identifying causal variants.
Figure 2

An integrative functional genomics approach for evolutionary studies. During the dispersal of modern humans, admixture between different populations, archaic introgression and natural selection have shaped the allele frequencies in human populations. The functional alleles (arising de novo or from standing variation) influence an individual’s phenotype (morphological or physiological) and fitness by modulating gene expression (noncoding variants) or changing the protein function (coding variants). The individuals with adaptive variants have a higher probability of survival or reproduction in certain environments (adaptation). The selection signals could be detected by genomic scans of selection and the causal variants could be identified and validated with experimental methods. GWAS could link the potential causal variants with phenotypes and eQTLs could be used to link variants with their target genes. However, both GWAS and eQTL analyses only indicate that association (dashed lines) and experimental validations are necessary for identifying causal variants.

The majority of GWAS hits (~93%) and selection scan outliers (~90%) are located in noncoding regions (65,117–119). Noncoding variants in regulatory elements (REs) could impact the expression of target genes by altering the activities of enhancers or promoters, including some that act at long distances by chromatin looping (120,121). Overlapping candidate pigmentation variants with REs is an effective way to prioritize potential causal variants (122–124). However, the effects of non-coding causal variants can be highly cell-type specific. For example, rs12913832 in the HERC2 gene has been shown to be located in a melanocyte-specific enhancer that modulates nearby OCA2 expression by chromatin looping (76). Thus, to prioritize the functional variants related to pigmentation, one should overlap them with REs (e.g. annotations from the ENCODE (125) and/or RoadMap databases (126)) from pigmentation-related cells such as melanocytes, fibroblasts and keratinocytes. Another approach to localize potential functional variants is to overlap candidate pigmentation variants with expression quantitative trait loci (eQTL), which are genetic markers associated with the expression of one or more genes. However, one should be aware that about 50% of detected eQTL are cell-type specific and some African-specific variants may not be present in eQTL datasets such as GTEx (127–129).

To experimentally test the effects of regulatory variants in vitro, one could use luciferase reporter assays. With the development of massively parallel reporter assays (MPRA) (130), it is possible to screen thousands of candidate variants simultaneously in trait-related cells. Both methods are based on plasmid vectors, which harbor a reporter gene driven by the enhancer of interest, which contains the candidate variant. Luciferase assays use the light output of luciferase as an indicator of enhancer activity, while MPRA use read counts from high throughput sequencing as a measure of enhancer activity (130). Because enhancers have cell-type-dependent activity, it is important to perform these assays in trait-related cells (e.g. melanocytes for pigmentation studies). If the two alleles of a variant show different enhancer activity, the next step is to determine the target gene of the variant. Most studies link regulatory variants with their nearest gene based on genomic distance; however, this assumption is not always true. A recent multiplexed CRISPRi screen tested the target genes of 470 enhancers and found that only 67% of enhancers are linked to their most proximal gene (131). Two methods could refine the potential target genes: chromatin interaction capture-based approaches, which capture the interactions between regulatory elements (132–134), and CRISPR/Cas9 perturbations (131,135). The final step is using CRISPR/Cas9 tools to modify the variant in cell lines (e.g. melanocytes for skin pigmentation) or animal models, and then compare the phenotypic effects of the two alleles (136).

Taken together, there is an urgent need to validate candidate pigmentation variants discovered using GWAS and scans of natural selection, using functional experiments in vitro and in vivo. This will help to characterize the true causal variants underlying pigmentation variation, deepen our understanding of the selective processes acting on this trait, and provide guidance for the treatment of human skin diseases.

Conclusions and Future Directions

Skin pigmentation is a classic example of an adaptive trait and was historically thought to be determined by a small number of large-effect variants (13–22,67). However, recent GWAS in Africans and other populations have discovered multiple novel loci related to the evolution of human skin color, and at the same time revealed the complexity of this phenotype (23,28). The broad range of skin color in ethnically diverse Africans, which cannot be explained by previously identified large-effect alleles from Eurasian populations, indicates that African skin color is regulated by many other unknown alleles (23,28). As current African GWAS typically contain small sample sizes, the inclusion of more diverse populations from more variable African environments is needed in order to discover novel skin pigmentation loci.

One promising avenue of future research in skin pigmentation genomics is to utilize the increasing availability of ancient DNA, which has had an important impact on our understanding of skin pigmentation evolution. Recent studies examining ancient DNA samples from the last ~40k years in Europe and Asia find strong signatures of recent positive selection acting at loci underlying light skin color, including at SLC45A2, SLC24A5, TYR and HERC2 (14,137,138). While environmental conditions in Africa are not conducive to ancient DNA preservation, recent technical advances resulting in the generation of more ancient African genomes may improve our understanding of the evolutionary history of skin pigmentation in Africa (139).

Another avenue of future research is to examine the impact of structural genetic variation (SV) on skin pigmentation. Structural variants are polymorphisms larger than 50 bp, including insertions, deletions, inversions, duplications and translocations, which are difficult to detect using standard short-read sequencing methods (140,141). The recent development of high-fidelity long read sequencing technology has opened the door for SV analysis in human genomics (142). We anticipate that these long-read technologies will help to reveal the role of structural variants in skin pigmentation evolution and biology.

In conclusion, to decipher the full genetic map of skin pigmentation in Africa, it is necessary to include more African populations in genomic studies, develop novel methods to detect selection on standing genetic variation and validate candidate variants with functional experiments. Furthermore, analyzing ancient genomes, exploring the role of structural variation and combining all of these together in an integrated way will deepen our understanding of human phenotypic evolution.

Conflict of Interest statement. None declared.

Funding

National Institutes of Health (R35GM134957, R01AR076241-01A1 to S.A.T.); the Center of Excellence in Environmental Toxicology (CEET) T32 (T32ES019851 to M.A.M.).

References

1.

Klein
,
R.G.
(
2009
)
Darwin and the recent African origin of modern humans
.
Proc. Natl. Acad. Sci. USA
,
106
,
16007
16009
.

2.

Tattersall
,
I.
(
2009
)
Out of Africa: modern human origins special feature: human origins: out of Africa
.
Proc. Natl. Acad. Sci. USA
,
106
,
16018
16021
.

3.

Vigilant
,
L.
,
Stoneking
,
M.
,
Harpending
,
H.
,
Hawkes
,
K.
and
Wilson
,
A.
(
1991
)
African populations and the evolution of human mitochondrial DNA
.
Science
,
253
,
1503
1507
.

4.

Tishkoff
,
S.A.
,
Dietzsch
,
E.
,
Speed
,
W.
,
Pakstis
,
A.J.
,
Kidd
,
J.R.
,
Cheung
,
K.
,
Bonne-Tamir
,
B.
,
Santachiara-Benerecetti
,
A.S.
,
Moral
,
P.
,
Krings
,
M.
 et al. (
1996
)
Global patterns of linkage disequilibrium at the CD4 locus and modern human origins
.
Science
,
271
,
1380
1387
.

5.

Ingman
,
M.
,
Kaessmann
,
H.
,
Pääbo
,
S.
and
Gyllensten
,
U.
(
2000
)
Mitochondrial genome variation and the origin of modern humans
.
Nature
,
408
,
708
713
.

6.

Tishkoff
,
S.A.
,
Reed
,
F.A.
,
Friedlaender
,
F.R.
,
Ehret
,
C.
,
Ranciaro
,
A.
,
Froment
,
A.
,
Hirbo
,
J.B.
,
Awomoyi
,
A.A.
,
Bodo
,
J.M.
,
Doumbo
,
O.
 et al. (
2009
)
The genetic structure and history of Africans and African Americans
.
Science
,
324
,
1035
1044
.

7.

Chaplin
,
G.
(
2004
)
Geographic distribution of environmental factors influencing human skin coloration
.
Am. J. Phys. Anthropol.
,
125
,
292
302
.

8.

Jablonski
,
N.G.
and
Chaplin
,
G.
(
2010
)
Colloquium paper: human skin pigmentation as an adaptation to UV radiation
.
Proc. Natl. Acad. Sci. USA
,
107
(
Suppl 2
),
8962
8968
.

9.

Williams
,
J. D.
,
Jacobson
,
E. L.
,
Kim
,
H.
,
Kim
 
M.
,
Jacobson
 
M.K.
(
2012
) Folate in skin cancer prevention. In
Stanger
,
O
. (ed.),
Water Soluble Vitamins: Clinical Research and Future Application
,
Springer
 
Dordrecht, The Netherlands
, pp.
181
197
.

10.

Jones
,
P.
,
Lucock
,
M.
,
Veysey
,
M.
and
Beckett
,
E.
(
2018
)
The vitamin D(−)folate hypothesis as an evolutionary model for skin pigmentation: an update and integration of current ideas
.
Nutrients
,
10
, 554.

11.

Tamura
,
T.
and
Picciano
,
M.F.
(
2006
)
Folate and human reproduction
.
Am. J. Clin. Nutr.
,
83
,
993
1016
.

12.

Luk
,
J.
,
Torrealday
,
S.
,
Neal Perry
,
G.
and
Pal
,
L.
(
2012
)
Relevance of vitamin D in reproduction
.
Hum. Reprod.
,
27
,
3015
3027
.

13.

Norton
,
H.L.
,
Kittles
,
R.A.
,
Parra
,
E.
,
McKeigue
,
P.
,
Mao
,
X.
,
Cheng
,
K.
,
Canfield
,
V.A.
,
Bradley
,
D.G.
,
McEvoy
,
B.
and
Shriver
,
M.D.
(
2007
)
Genetic evidence for the convergent evolution of light skin in Europeans and East Asians
.
Mol. Biol. Evol.
,
24
,
710
722
.

14.

Wilde
,
S.
,
Timpson
,
A.
,
Kirsanow
,
K.
,
Kaiser
,
E.
,
Kayser
,
M.
,
Unterländer
,
M.
,
Hollfelder
,
N.
,
Potekhina
,
I.D.
,
Schier
,
W.
,
Thomas
,
M.G.
and
Burger
,
J.
(
2014
)
Direct evidence for positive selection of skin, hair, and eye pigmentation in Europeans during the last 5,000 y
.
Proc. Natl. Acad. Sci. USA
,
111
,
4832
4837
.

15.

McEvoy
,
B.
,
Beleza
,
S.
and
Shriver
,
M.D.
(
2006
)
The genetic architecture of normal variation in human pigmentation: an evolutionary perspective and model
.
Hum. Mol. Genet.
,
15 Spec No 2
,
R176
R181
.

16.

Myles
,
S.
,
Somel
,
M.
,
Tang
,
K.
,
Kelso
,
J.
and
Stoneking
,
M.
(
2007
)
Identifying genes underlying skin pigmentation differences among human populations
.
Hum. Genet.
,
120
,
613
621
.

17.

Pickrell
,
J.K.
,
Coop
,
G.
,
Novembre
,
J.
,
Kudaravalli
,
S.
,
Li
,
J.Z.
,
Absher
,
D.
,
Srinivasan
,
B.S.
,
Barsh
,
G.S.
,
Myers
,
R.M.
,
Feldman
,
M.W.
and
Pritchard
,
J.K.
(
2009
)
Signals of recent positive selection in a worldwide sample of human populations
.
Genome Res.
,
19
,
826
837
.

18.

Hider
,
J.L.
,
Gittelman
,
R.M.
,
Shah
,
T.
,
Edwards
,
M.
,
Rosenbloom
,
A.
,
Akey
,
J.M.
and
Parra
,
E.J.
(
2013
)
Exploring signatures of positive selection in pigmentation candidate genes in populations of East Asian ancestry
.
BMC Evol. Biol.
,
13
,
150
.

19.

Lao
,
O.
,
de
 
Gruijter
,
J.M.
,
van
 
Duijn
,
K.
,
Navarro
,
A.
and
Kayser
,
M.
(
2007
)
Signatures of positive selection in genes associated with human skin pigmentation as revealed from analyses of single nucleotide polymorphisms
.
Ann. Hum. Genet.
,
71
,
354
369
.

20.

Sturm
,
R.A.
and
Duffy
,
D.L.
(
2012
)
Human pigmentation genes under environmental selection
.
Genome Biol.
,
13
,
248
.

21.

Lamason
,
R.L.
,
Mohideen
,
M.-A.P.K.
,
Mest
,
J.R.
,
Wong
,
A.C.
,
Norton
,
H.L.
,
Aros
,
M.C.
,
Jurynec
,
M.J.
,
Mao
,
X.
,
Humphreville
,
V.R.
,
Humbert
,
J.E.
 et al. (
2005
)
SLC24A5, a putative cation exchanger, affects pigmentation in zebrafish and humans
.
Science
,
310
,
1782
1786
.

22.

Jonnalagadda
,
M.
,
Bharti
,
N.
,
Patil
,
Y.
,
Ozarkar
,
S.
,
K
,
S.M.
,
Joshi
,
R.
and
Norton
,
H.
(
2017
)
Identifying signatures of positive selection in pigmentation genes in two south Asian populations
.
Am. J. Hum. Biol.
,
29
.

23.

Crawford
,
N.G.
,
Kelly
,
D.E.
,
Hansen
,
M.E.B.
,
Beltrame
,
M.H.
,
Fan
,
S.
,
Bowman
,
S.L.
,
Jewett
,
E.
,
Ranciaro
,
A.
,
Thompson
,
S.
,
Lo
,
Y.
 et al. (
2017
)
Loci associated with skin pigmentation identified in African populations
.
Science
,
358
, eaan8433.

24.

Post
,
P.W.
,
Daniels
,
F.
, Jr.
and
Binford
,
R.T.
, Jr.
(
1975
)
Cold injury and the evolution of “white” skin
.
Hum. Biol.
,
47
,
65
80
.

25.

Jablonski
,
N.G.
and
Chaplin
,
G.
(
2017
)
The colours of humanity: the evolution of pigmentation in the human lineage
.
Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci.
,
372
, 20160349.

26.

Sirugo
,
G.
,
Williams
,
S.M.
and
Tishkoff
,
S.A.
(
2019
)
The missing diversity in human genetic studies
.
Cell
,
177
,
26
31
.

27.

MacArthur
,
J.
,
Bowler
,
E.
,
Cerezo
,
M.
,
Gil
,
L.
,
Hall
,
P.
,
Hastings
,
E.
,
Junkins
,
H.
,
McMahon
,
A.
,
Milano
,
A.
,
Morales
,
J.
 et al. (
2017
)
The new NHGRI-EBI Catalog of published genome-wide association studies (GWAS Catalog)
.
Nucleic Acids Res.
,
45
,
D896
D901
.

28.

Martin
,
A.R.
,
Lin
,
M.
,
Granka
,
J.M.
,
Myrick
,
J.W.
,
Liu
,
X.
,
Sockell
,
A.
,
Atkinson
,
E.G.
,
Werely
,
C.J.
,
Möller
,
M.
,
Sandhu
,
M.S.
 et al. (
2017
)
An unexpectedly complex architecture for skin pigmentation in Africans
.
Cell
,
171
,
1340
,
e14
1353
.

29.

Pavan
,
W.J.
and
Sturm
,
R.A.
(
2019
)
The genetics of human skin and hair pigmentation
.
Annu. Rev. Genomics Hum. Genet.
,
20
,
41
72
.

30.

Raposo
,
G.
and
Marks
,
M.S.
(
2007
)
Melanosomes—dark organelles enlighten endosomal membrane transport
.
Nat. Rev. Mol. Cell Biol.
,
8
,
786
797
.

31.

Thomas
,
A.J.
and
Erickson
,
C.A.
(
2008
)
The making of a melanocyte: the specification of melanoblasts from the neural crest
.
Pigment Cell Melanoma Res.
,
21
,
598
610
.

32.

Vandamme
,
N.
and
Berx
,
G.
(
2019
)
From neural crest cells to melanocytes: cellular plasticity during development and beyond
.
Cell. Mol. Life Sci.
,
76
,
1919
1934
.

33.

Baxter
,
L.L.
,
Watkins-Chow
,
D.E.
,
Pavan
,
W.J.
and
Loftus
,
S.K.
(
2019
)
A curated gene list for expanding the horizons of pigmentation biology
.
Pigment Cell Melanoma Res.
,
32
,
348
358
.

34.

The IMPC consortium
,
Muñoz-Fuentes
,
V.
,
Cacheiro
,
P.
,
Meehan
,
T.F.
,
Aguilar-Pimentel
,
J.A.
,
Brown
,
S.D.M.
,
Flenniken
,
A.M.
,
Flicek
,
P.
,
Galli
,
A.
,
Mashhadi
,
H.H.
 et al. (
2018
)
The International Mouse Phenotyping Consortium (IMPC): a functional catalogue of the mammalian genome that informs conservation
.
Conserv. Genet.
,
19
,
995
1005
.

35.

D’Mello
,
S.A.N.
,
Finlay
,
G.J.
,
Baguley
,
B.C.
and
Askarian-Amiri
,
M.
(
2016
)
Signaling pathways in melanogenesis
.
Int. J. Mol. Sci.
,
17
,
1144
.

36.

Lin
,
J.Y.
and
Fisher
,
D.E.
(
2007
)
Melanocyte biology and skin pigmentation
.
Nature
,
445
,
843
850
.

37.

Sulem
,
P.
,
Gudbjartsson
,
D.F.
,
Stacey
,
S.N.
,
Helgason
,
A.
,
Rafnar
,
T.
,
Magnusson
,
K.P.
,
Manolescu
,
A.
,
Karason
,
A.
,
Palsson
,
A.
,
Thorleifsson
,
G.
 et al. (
2007
)
Genetic determinants of hair, eye and skin pigmentation in Europeans
.
Nat. Genet.
,
39
,
1443
1452
.

38.

Morgan
,
M.D.
,
Pairo-Castineira
,
E.
,
Rawlik
,
K.
,
Canela-Xandri
,
O.
,
Rees
,
J.
,
Sims
,
D.
,
Tenesa
,
A.
and
Jackson
,
I.J.
(
2018
)
Genome-wide study of hair colour in UK Biobank explains most of the SNP heritability
.
Nat. Commun.
,
9
,
5271
.

39.

Fu
,
W.
and
Akey
,
J.M.
(
2013
)
Selection and adaptation in the human genome
.
Annu. Rev. Genomics Hum. Genet.
,
14
,
467
489
.

40.

Scheinfeldt
,
L.B.
and
Tishkoff
,
S.A.
(
2013
)
Recent human adaptation: genomic approaches, interpretation and insights
.
Nat. Rev. Genet.
,
14
,
692
702
.

41.

Koboldt
,
D.C.
,
Steinberg
,
K.M.
,
Larson
,
D.E.
,
Wilson
,
R.K.
and
Mardis
,
E.R.
(
2013
)
The next-generation sequencing revolution and its impact on genomics
.
Cell
,
155
,
27
38
.

42.

Nielsen
,
R.
,
Akey
,
J.M.
,
Jakobsson
,
M.
,
Pritchard
,
J.K.
,
Tishkoff
,
S.
and
Willerslev
,
E.
(
2017
)
Tracing the peopling of the world through genomics
.
Nature
,
541
,
302
310
.

43.

Novembre
,
J.
and
Barton
,
N.H.
(
2018
)
Tread lightly interpreting polygenic tests of selection
.
Genetics
,
208
,
1351
1355
.

44.

Pritchard
,
J.K.
,
Pickrell
,
J.K.
and
Coop
,
G.
(
2010
)
The genetics of human adaptation: hard sweeps, soft sweeps, and polygenic adaptation
.
Curr. Biol.
,
20
,
R208
R215
.

45.

Chen
,
H.
,
Patterson
,
N.
and
Reich
,
D.
(
2010
)
Population differentiation as a test for selective sweeps
.
Genome Res.
,
20
,
393
402
.

46.

Holsinger
,
K.E.
and
Weir
,
B.S.
(
2009
)
Genetics in geographically structured populations: defining, estimating and interpreting F(ST)
.
Nat. Rev. Genet.
,
10
,
639
650
.

47.

Nielsen
,
R.
(
2005
)
Molecular signatures of natural selection
.
Annu. Rev. Genet.
,
39
,
197
218
.

48.

Vitti
,
J.J.
,
Grossman
,
S.R.
and
Sabeti
,
P.C.
(
2013
)
Detecting natural selection in genomic data
.
Annu. Rev. Genet.
,
47
,
97
120
.

49.

Booker
,
T.R.
,
Jackson
,
B.C.
and
Keightley
,
P.D.
(
2017
)
Detecting positive selection in the genome
.
BMC Biol.
,
15
,
98
.

50.

Fan
,
S.
,
Hansen
,
M.E.B.
,
Lo
,
Y.
and
Tishkoff
,
S.A.
(
2016
)
Going global by adapting local: a review of recent human adaptation
.
Science
,
354
,
54
59
.

51.

Tajima
,
F.
(
1989
)
Statistical method for testing the neutral mutation hypothesis by DNA polymorphism
.
Genetics
,
123
,
585
595
.

52.

Fay
,
J.C.
and
Wu
,
C.I.
(
2000
)
Hitchhiking under positive Darwinian selection
.
Genetics
,
155
,
1405
1413
.

53.

Sabeti
,
P.C.
,
Reich
,
D.E.
,
Higgins
,
J.M.
,
Levine
,
H.Z.P.
,
Richter
,
D.J.
,
Schaffner
,
S.F.
,
Gabriel
,
S.B.
,
Platko
,
J.V.
,
Patterson
,
N.J.
,
McDonald
,
G.J.
 et al. (
2002
)
Detecting recent positive selection in the human genome from haplotype structure
.
Nature
,
419
,
832
837
.

54.

Wang
,
E.T.
,
Kodama
,
G.
,
Baldi
,
P.
and
Moyzis
,
R.K.
(
2006
)
Global landscape of recent inferred Darwinian selection for Homo sapiens
.
Proc. Natl. Acad. Sci. USA
,
103
,
135
140
.

55.

Cai
,
Z.
,
Camp
,
N.J.
,
Cannon-Albright
,
L.
and
Thomas
,
A.
(
2011
)
Identification of regions of positive selection using shared genomic segment analysis
.
Eur. J. Hum. Genet.
,
19
,
667
671
.

56.

Voight
,
B.F.
,
Kudaravalli
,
S.
,
Wen
,
X.
and
Pritchard
,
J.K.
(
2006
)
A map of recent positive selection in the human genome
.
PLoS Biol.
,
4
,
e72
.

57.

Tang
,
K.
,
Thornton
,
K.R.
and
Stoneking
,
M.
(
2007
)
A New approach for using genome scans to detect recent positive selection in the human genome
.
PLoS Biol.
,
5
,
e171
.

58.

Ferrer-Admetlla
,
A.
,
Liang
,
M.
,
Korneliussen
,
T.
and
Nielsen
,
R.
(
2014
)
On detecting incomplete soft or hard selective sweeps using haplotype structure
.
Mol. Biol. Evol.
,
31
,
1275
1291
.

59.

Shriver
,
M.D.
,
Kennedy
,
G.C.
,
Parra
,
E.J.
,
Lawson
,
H.A.
,
Sonpar
,
V.
,
Huang
,
J.
,
Akey
,
J.M.
and
Jones
,
K.W.
(
2004
)
The genomic distribution of population substructure in four populations using 8,525 autosomal SNPs
.
Hum. Genomics
,
1
,
274
286
.

60.

Yi
,
X.
,
Liang
,
Y.
,
Huerta-Sanchez
,
E.
,
Jin
,
X.
,
Cuo
,
Z.X.P.
,
Pool
,
J.E.
,
Xu
,
X.
,
Jiang
,
H.
,
Vinckenbosch
,
N.
,
Korneliussen
,
T.S.
 et al. (
2010
)
Sequencing of 50 human exomes reveals adaptation to high altitude
.
Science
,
329
,
75
78
.

61.

Sabeti
,
P.C.
,
Varilly
,
P.
,
Fry
,
B.
,
Lohmueller
,
J.
,
Hostetter
,
E.
,
Cotsapas
,
C.
,
Xie
,
X.
,
Byrne
,
E.H.
,
McCarroll
,
S.
,
Gaudet
,
R.
 et al. (
2007
)
Genome-wide detection and characterization of positive selection in human populations
.
Nature
,
449
,
913
918
.

62.

Refoyo-Martínez
,
A.
,
da
 
Fonseca
,
R.R.
,
Halldórsdóttir
,
K.
,
Árnason
,
E.
,
Mailund
,
T.
and
Racimo
,
F.
(
2019
)
Identifying loci under positive selection in complex population histories
.
Genome Res.
,
29
,
1506
1520
.

63.

Williamson
,
S.H.
,
Hubisz
,
M.J.
,
Clark
,
A.G.
,
Payseur
,
B.A.
,
Bustamante
,
C.D.
and
Nielsen
,
R.
(
2007
)
Localizing recent adaptive evolution in the human genome
.
PLoS Genet.
,
3
,
e90
.

64.

Grossman
,
S.R.
,
Shylakhter
,
I.
,
Karlsson
,
E.K.
,
Byrne
,
E.H.
,
Morales
,
S.
,
Frieden
,
G.
,
Hostetter
,
E.
,
Angelino
,
E.
,
Garber
,
M.
,
Zuk
,
O.
 et al. (
2010
)
A composite of multiple signals distinguishes causal variants in regions of positive selection
.
Science
,
327
,
883
886
.

65.

Grossman
,
S.R.
,
Andersen
,
K.G.
,
Shlyakhter
,
I.
,
Tabrizi
,
S.
,
Winnicki
,
S.
,
Yen
,
A.
,
Park
,
D.J.
,
Griesemer
,
D.
,
Karlsson
,
E.K.
,
Wong
,
S.H.
 et al. (
2013
)
Identifying recent adaptations in large-scale genomic data
.
Cell
,
152
,
703
713
.

66.

Field
,
Y.
,
Boyle
,
E.A.
,
Telis
,
N.
,
Gao
,
Z.
,
Gaulton
,
K.J.
,
Golan
,
D.
,
Yengo
,
L.
,
Rocheleau
,
G.
,
Froguel
,
P.
,
McCarthy
,
M.I.
and
Pritchard
,
J.K.
(
2016
)
Detection of human adaptation during the past 2000 years
.
Science
,
354
,
760
764
.

67.

Deng
,
L.
and
Xu
,
S.
(
2018
)
Adaptation of human skin color in various populations
.
Hereditas
,
155
,
1
.

68.

Quillen
,
E.E.
,
Norton
,
H.L.
,
Parra
,
E.J.
,
Lona-Durazo
,
F.
,
Ang
,
K.C.
,
Illiescu
,
F.M.
,
Pearson
,
L.N.
,
Shriver
,
M.D.
,
Lasisi
,
T.
,
Gokcumen
,
O.
 et al. (
2019
)
Shades of complexity: New perspectives on the evolution and genetic architecture of human skin
.
Am. J. Phys. Anthropol.
,
168 Suppl 67
,
4
26
.

69.

Buniello
,
A.
,
MacArthur
,
J.A.L.
,
Cerezo
,
M.
,
Harris
,
L.W.
,
Hayhurst
,
J.
,
Malangone
,
C.
,
McMahon
,
A.
,
Morales
,
J.
,
Mountjoy
,
E.
,
Sollis
,
E.
 et al. (
2019
)
The NHGRI-EBI GWAS Catalog of published genome-wide association studies, targeted arrays and summary statistics 2019
.
Nucleic Acids Res.
,
47
,
D1005
D1012
.

70.

Beleza
,
S.
,
Santos
,
A.M.
,
McEvoy
,
B.
,
Alves
,
I.
,
Martinho
,
C.
,
Cameron
,
E.
,
Shriver
,
M.D.
,
Parra
,
E.J.
and
Rocha
,
J.
(
2013
)
The timing of pigmentation lightening in Europeans
.
Mol. Biol. Evol.
,
30
,
24
35
.

71.

Basu Mallick
,
C.
,
Iliescu
,
F.M.
,
Möls
,
M.
,
Hill
,
S.
,
Tamang
,
R.
,
Chaubey
,
G.
,
Goto
,
R.
,
Ho
,
S.Y.W.
,
Gallego Romero
,
I.
,
Crivellaro
,
F.
 et al. (
2013
)
The light skin allele of SLC24A5 in South Asians and Europeans shares identity by descent
.
PLoS Genet.
,
9
,
e1003912
.

72.

Lin
,
M.
,
Siford
,
R.L.
,
Martin
,
A.R.
,
Nakagome
,
S.
,
Möller
,
M.
,
Hoal
,
E.G.
,
Bustamante
,
C.D.
,
Gignoux
,
C.R.
and
Henn
,
B.M.
(
2018
)
Rapid evolution of a skin-lightening allele in southern African KhoeSan
.
Proc. Natl. Acad. Sci. USA
,
115
,
13324
13329
.

73.

Bellono
,
N.W.
,
Escobar
,
I.E.
,
Lefkovith
,
A.J.
,
Marks
,
M.S.
and
Oancea
,
E.
(
2014
)
An intracellular anion channel critical for pigmentation
.
Elife
,
3
,
e04543
.

74.

Puri
,
N.
,
Gardner
,
J.M.
and
Brilliant
,
M.H.
(
2000
)
Aberrant pH of melanosomes in pink-eyed dilution (p) mutant melanocytes
.
J. Invest. Dermatol.
,
115
,
607
613
.

75.

Rinchik, E.M., Bultman, S.J., Horsthemke, B., Lee, S.T., Strunk, K.M., Spritz, R.A., Avidano, K.M., Jong, M.T. and Nicholls, R.D. (1993) A gene for the mouse pink-eyed dilution locus and for human type II oculocutaneous albinism.

Nature
,
361
, 72–76.

76.

Visser
,
M.
,
Kayser
,
M.
and
Palstra
,
R.J.
(
2012
)
HERC2 rs12913832 modulates human pigmentation by attenuating chromatin-loop formation between a long-range enhancer and the OCA2 promoter
.
Genome Res.
,
22
,
446
455
.

77.

Yuasa
,
I.
,
Umetsu
,
K.
,
Harihara
,
S.
,
Kido
,
A.
,
Miyoshi
,
A.
,
Saitou
,
N.
,
Dashnyam
,
B.
,
Jin
,
F.
,
Lucotte
,
G.
,
Chattopadhyay
,
P.K.
,
Henke
,
L.
and
Henke
,
J.
(
2007
)
Distribution of two Asian-related coding SNPs in the MC1R and OCA2 genes
.
Biochem. Genet.
,
45
,
535
542
.

78.

Branicki
,
W.
,
Brudnik
,
U.
and
Wojas-Pelc
,
A.
(
2009
)
Interactions between HERC2, OCA2 and MC1R may influence human pigmentation phenotype
.
Ann. Hum. Genet.
,
73
,
160
170
.

79.

Edwards
,
M.
,
Bigham
,
A.
,
Tan
,
J.
,
Li
,
S.
,
Gozdzik
,
A.
,
Ross
,
K.
,
Jin
,
L.
and
Parra
,
E.J.
(
2010
)
Association of the OCA2 polymorphism His615Arg with melanin content in east Asian populations: further evidence of convergent evolution of skin pigmentation
.
PLoS Genet.
,
6
,
e1000867
.

80.

Han
,
J.
,
Kraft
,
P.
,
Nan
,
H.
,
Guo
,
Q.
,
Chen
,
C.
,
Qureshi
,
A.
,
Hankinson
,
S.E.
,
Hu
,
F.B.
,
Duffy
,
D.L.
,
Zhao
,
Z.Z.
 et al. (
2008
)
A genome-wide association study identifies novel alleles associated with hair color and skin pigmentation
.
PLoS Genet.
,
4
,
e1000074
.

81.

Duffy
,
D.L.
,
Montgomery
,
G.W.
,
Chen
,
W.
,
Zhao
,
Z.Z.
,
le
,
L.
,
James
,
M.R.
,
Hayward
,
N.K.
,
Martin
,
N.G.
and
Sturm
,
R.A.
(
2007
)
A three–single-nucleotide polymorphism haplotype in intron 1 of OCA2 explains most human eye-color variation
.
Am. J. Hum. Genet.
,
80
,
241
252
.

82.

Sturm
,
R.A.
,
Duffy
,
D.L.
,
Zhao
,
Z.Z.
,
Leite
,
F.P.N.
,
Stark
,
M.S.
,
Hayward
,
N.K.
,
Martin
,
N.G.
and
Montgomery
,
G.W.
(
2008
)
A single SNP in an evolutionary conserved region within intron 86 of the HERC2 gene determines human blue-brown eye color
.
Am. J. Hum. Genet.
,
82
,
424
431
.

83.

Kayser
,
M.
,
Liu
,
F.
,
Janssens
,
A.C.J.W.
,
Rivadeneira
,
F.
,
Lao
,
O.
,
van Duijn
,
K.
,
Vermeulen
,
M.
,
Arp
,
P.
,
Jhamai
,
M.M.
,
van IJcken
,
W.F.J.
 et al. (
2008
)
Three genome-wide association studies and a linkage analysis identify HERC2 as a human iris color gene
.
Am. J. Hum. Genet.
,
82
,
411
423
.

84.

Donnelly
,
M.P.
,
Paschou
,
P.
,
Grigorenko
,
E.
,
Gurwitz
,
D.
,
Barta
,
C.
,
Lu
,
R.B.
,
Zhukova
,
O.V.
,
Kim
,
J.J.
,
Siniscalco
,
M.
,
New
,
M.
 et al. (
2012
)
A global view of the OCA2-HERC2 region and pigmentation
.
Hum. Genet.
,
131
,
683
696
.

85.

Kapetanaki
,
M.G.
,
Guerrero-Santoro
,
J.
,
Bisi
,
D.C.
,
Hsieh
,
C.L.
,
Rapic-Otrin
,
V.
and
Levine
,
A.S.
(
2006
)
The DDB1-CUL4ADDB2 ubiquitin ligase is deficient in xeroderma pigmentosum group E and targets histone H2A at UV-damaged DNA sites
.
Proc. Natl. Acad. Sci. USA
,
103
,
2588
2593
.

86.

Lieberman
,
M.
,
Segev
,
O.
,
Gilboa
,
N.
,
Lalazar
,
A.
and
Levin
,
I.
(
2004
)
The tomato homolog of the gene encoding UV-damaged DNA binding protein 1 (DDB1) underlined as the gene that causes the high pigment-1 mutant phenotype
.
Theor. Appl. Genet.
,
108
,
1574
1581
.

87.

Lloyd-Jones
,
L.R.
,
Robinson
,
M.R.
,
Moser
,
G.
,
Zeng
,
J.
,
Beleza
,
S.
,
Barsh
,
G.S.
,
Tang
,
H.
and
Visscher
,
P.M.
(
2017
)
Inference on the genetic basis of eye and skin color in an admixed population via Bayesian linear mixed models
.
Genetics
,
206
,
1113
1126
.

88.

Adelmann, C.H., Traunbauer, A.K., Chen, B., Condon, K.J., Chan, S.H., Kunchok, T., Lewis, C.A. and Sabatini, D.M. (2020) MFSD12 mediates the import of cysteine into melanosomes and lysosomes.

Nature
,
588
,
699
704
.

89.

Wei
,
C.-Y.
,
Zhu
,
M.-X.
,
Lu
,
N.-H.
,
Peng
,
R.
,
Yang
,
X.
,
Zhang
,
P.F.
,
Wang
,
L.
and
Gu
,
J.Y.
(
2019
)
Bioinformatics-based analysis reveals elevated MFSD12 as a key promoter of cell proliferation and a potential therapeutic target in melanoma
.
Oncogene
,
38
,
1876
1891
.

90.

Adhikari
,
K.
,
Mendoza-Revilla
,
J.
,
Sohail
,
A.
,
Fuentes-Guajardo
,
M.
,
Lampert
,
J.
,
Chacón-Duque
,
J.C.
,
Hurtado
,
M.
,
Villegas
,
V.
,
Granja
,
V.
,
Acuña-Alonzo
,
V.
 et al. (
2019
)
A GWAS in Latin Americans highlights the convergent evolution of lighter skin pigmentation in Eurasia
.
Nat. Commun.
,
10
,
358
.

91.

Shin, J.G., Leem, S., Kim, B., Kim, Y., Lee, S.G., Song, H.J., Seo, J.Y., Park, S.G., Won, H.H. and Kang, N.G. (2020) GWAS Analysis of 17,019 Korean Women Identifies the Variants Associated with Facial Pigmented Spots.

J. Invest. Dermatol.
, in press.

92.

Pan-UKB team
. (
2020
)
Pan-ancestry genetic analysis of the UK Biobank
. https://pan.ukbb.broadinstitute.org.
(accessed 2020)
.

93.

Rees
,
J.L.
and
Flanagan
,
N.
(
1999
)
Pigmentation, melanocortins and red hair
.
QJM
,
92
,
125
131
.

94.

Harding
,
R.M.
,
Healy
,
E.
,
Ray
,
A.J.
,
Ellis
,
N.S.
,
Flanagan
,
N.
,
Todd
,
C.
,
Dixon
,
C.
,
Sajantila
,
A.
,
Jackson
,
I.J.
,
Birch-Machin
,
M.A.
and
Rees
,
J.L.
(
2000
)
Evidence for variable selective pressures at MC1R
.
Am. J. Hum. Genet.
,
66
,
1351
1361
.

95.

JOHN
,
P.R.
,
MAKOVA
,
K.
,
LI
,
W.H.
,
JENKINS
,
T.
and
RAMSAY
,
M.
(
2003
)
DNA polymorphism and selection at the melanocortin-1 receptor gene in normally pigmented southern African individuals
.
Ann. N. Y. Acad. Sci.
,
994
,
299
306
.

96.

Jacobs
,
L.C.
,
Hamer
,
M.A.
,
Gunn
,
D.A.
,
Deelen
,
J.
,
Lall
,
J.S.
,
van Heemst
,
D.
,
Uh
,
H.W.
,
Hofman
,
A.
,
Uitterlinden
,
A.G.
,
Griffiths
,
C.E.M.
 et al. (
2015
)
A genome-wide association study identifies the skin color genes IRF4, MC1R, ASIP, and BNC2 influencing facial pigmented spots
.
J. Invest. Dermatol.
,
135
,
1735
1742
.

97.

Liu
,
F.
,
Visser
,
M.
,
Duffy
,
D.L.
,
Hysi
,
P.G.
,
Jacobs
,
L.C.
,
Lao
,
O.
,
Zhong
,
K.
,
Walsh
,
S.
,
Chaitanya
,
L.
,
Wollstein
,
A.
 et al. (
2015
)
Genetics of skin color variation in Europeans: genome-wide association studies with functional follow-up
.
Hum. Genet.
,
134
,
823
835
.

98.

Galván-Femenía
,
I.
,
Obón-Santacana
,
M.
,
Piñeyro
,
D.
,
Guindo-Martinez
,
M.
,
Duran
,
X.
,
Carreras
,
A.
,
Pluvinet
,
R.
,
Velasco
,
J.
,
Ramos
,
L.
,
Aussó
,
S.
 et al. (
2018
)
Multitrait genome association analysis identifies new susceptibility genes for human anthropometric variation in the GCAT cohort
.
J. Med. Genet.
,
55
,
765
778
.

99.

The International Visible Trait Genetics Consortium
,
Hysi
,
P.G.
,
Valdes
,
A.M.
,
Liu
,
F.
,
Furlotte
,
N.A.
,
Evans
,
D.M.
,
Bataille
,
V.
,
Visconti
,
A.
,
Hemani
,
G.
,
McMahon
,
G.
 et al. (
2018
)
Genome-wide association meta-analysis of individuals of European ancestry identifies new loci explaining a substantial fraction of hair color variation and heritability
.
Nat. Genet.
,
50
,
652
656
.

100.

King
,
R.A.
,
Creel
,
D.
,
Cervenka
,
J.
,
Okoro
,
A.N.
and
Witkop
,
C.J.
(
1980
)
Albinism in Nigeria with delineation of new recessive oculocutaneous type
.
Clin. Genet.
,
17
,
259
270
.

101.

King
,
R.A.
and
Rich
,
S.S.
(
1986
)
Segregation analysis of brown oculocutaneous albinism
.
Clin. Genet.
,
29
,
496
501
.

102.

Manga
,
P.
,
Kromberg
,
J.G.
,
Box
,
N.F.
,
Sturm
,
R.A.
,
Jenkins
,
T.
and
Ramsay
,
M.
(
1997
)
Rufous oculocutaneous albinism in southern African blacks is caused by mutations in the TYRP1 gene
.
Am. J. Hum. Genet.
,
61
,
1095
1101
.

103.

Alonso
,
S.
,
Izagirre
,
N.
,
Smith-Zubiaga
,
I.
,
Gardeazabal
,
J.
,
Díaz-Ramón
,
J.
,
Díaz-Pérez
,
J.
,
Zelenika
,
D.
,
Boyano
,
M.
,
Smit
,
N.
and
de la Rúa
,
C.
(
2008
)
Complex signatures of selection for the melanogenic loci TYR, TYRP1 and DCT in humans
.
BMC Evol. Biol.
,
8
,
74
.

104.

Izagirre
,
N.
,
García
,
I.
,
Junquera
,
C.
,
de la Rúa
,
C.́.
and
Alonso
,
S.
(
2006
)
A scan for signatures of positive selection in candidate loci for skin pigmentation in humans
.
Mol. Biol. Evol.
,
23
,
1697
1706
.

105.

Hinds
,
D.A.
,
Stuve
,
L.L.
,
Nilsen
,
G.B.
,
Halperin
,
E.
,
Eskin
,
E.
,
Ballinger
,
D.G.
,
Frazer
,
K.A.
and
Cox
,
D.R.
(
2005
)
Whole-genome patterns of common DNA variation in three human populations
.
Science
,
307
,
1072
1079
.

106.

Akey
,
J.M.
(
2009
)
Constructing genomic maps of positive selection in humans: where do we go from here?
 
Genome Res.
,
19
,
711
722
.

107.

Teshima
,
K.M.
,
Coop
,
G.
and
Przeworski
,
M.
(
2006
)
How reliable are empirical genomic scans for selective sweeps?
 
Genome Res.
,
16
,
702
712
.

108.

Hutchinson
,
A.
,
Asimit
,
J.
and
Wallace
,
C.
(
2020
)
Fine-mapping genetic associations
.
Hum. Mol. Genet.
,
29
,
R81
R88
.

109.

Broekema
,
R.V.
,
Bakker
,
O.B.
and
Jonkers
,
I.H.
(
2020
)
A practical view of fine-mapping and gene prioritization in the post-genome-wide association era
.
Open Biol.
,
10
,
190221
.

110.

Lichou
,
F.
and
Trynka
,
G.
(
2020
)
Functional studies of GWAS variants are gaining momentum
.
Nat. Commun.
,
11
,
6283
.

111.

Nei
,
M.
,
Suzuki
,
Y.
and
Nozawa
,
M.
(
2010
)
The neutral theory of molecular evolution in the genomic era
.
Annu. Rev. Genomics Hum. Genet.
,
11
,
265
289
.

112.

Minwalla
,
L.
,
Zhao
,
Y.
,
Boissy
,
R.E.
,
le Poole
,
I.C.
and
Wickett
,
R.R.
(
2001
)
Keratinocytes play a role in regulating distribution patterns of recipient melanosomes in vitro
.
J. Invest. Dermatol.
,
117
,
341
347
.

113.

Yamaguchi
,
Y.
,
Brenner
,
M.
and
Hearing
,
V.J.
(
2007
)
The regulation of skin pigmentation
.
J. Biol. Chem.
,
282
,
27557
27561
.

114.

Hédan
,
B.
,
Cadieu
,
E.
,
Botherel
,
N.
 et al. (
2019
)
Identification of a missense variant in MFSD12 involved in dilution of Phaeomelanin leading to white or cream coat color in dogs
.
Genes
,
10
, 386.

115.

Tanaka
,
J.
,
Leeb
,
T.
,
Rushton
,
J.
,
Famula
,
T.R.
,
Mack
,
M.
,
Jagannathan
,
V.
,
Flury
,
C.
,
Bachmann
,
I.
,
Eberth
,
J.
,
McDonnell
,
S.M.
,
Penedo
,
M.C.T.
and
Bellone
,
R.R.
(
2019
)
Frameshift variant in MFSD12 explains the mushroom coat color dilution in Shetland ponies
.
Genes
,
10
, 826.

116.

Miller
,
C.T.
,
Beleza
,
S.
,
Pollen
,
A.A.
,
Schluter
,
D.
,
Kittles
,
R.A.
,
Shriver
,
M.D.
and
Kingsley
,
D.M.
(
2007
)
Cis-regulatory changes in kit ligand expression and parallel evolution of pigmentation in sticklebacks and humans
.
Cell
,
131
,
1179
1189
.

117.

Maurano
,
M.T.
,
Humbert
,
R.
,
Rynes
,
E.
,
Thurman
,
R.E.
,
Haugen
,
E.
,
Wang
,
H.
,
Reynolds
,
A.P.
,
Sandstrom
,
R.
,
Qu
,
H.
,
Brody
,
J.
 et al. (
2012
)
Systematic localization of common disease-associated variation in regulatory DNA
.
Science
,
337
,
1190
1195
.

118.

Yu
,
F.
,
Lu
,
J.
,
Liu
,
X.
,
Gazave
,
E.
,
Chang
,
D.
,
Raj
,
S.
,
Hunter-Zinck
,
H.
,
Blekhman
,
R.
,
Arbiza
,
L.
,
van Hout
,
C.
 et al. (
2015
)
Population genomic analysis of 962 whole genome sequences of humans reveals natural selection in non-coding regions
.
PLoS One
,
10
,
e0121644
.

119.

Enard
,
D.
,
Messer
,
P.W.
and
Petrov
,
D.A.
(
2014
)
Genome-wide signals of positive selection in human evolution
.
Genome Res.
,
24
,
885
895
.

120.

Li
,
Y.
,
Hu
,
M.
and
Shen
,
Y.
(
2018
)
Gene regulation in the 3D genome
.
Hum. Mol. Genet.
,
27
,
R228
R233
.

121.

Schoenfelder
,
S.
and
Fraser
,
P.
(
2019
)
Long-range enhancer–promoter contacts in gene expression control
.
Nat. Rev. Genet.
,
20
,
437
455
.

122.

Farh
,
K.K.-H.
,
Marson
,
A.
,
Zhu
,
J.
,
Kleinewietfeld
,
M.
,
Housley
,
W.J.
,
Beik
,
S.
,
Shoresh
,
N.
,
Whitton
,
H.
,
Ryan
,
R.J.H.
,
Shishkin
,
A.A.
 et al. (
2015
)
Genetic and epigenetic fine mapping of causal autoimmune disease variants
.
Nature
,
518
,
337
343
.

123.

Westra
,
H.-J.
,
Martínez-Bonet
,
M.
,
Onengut-Gumuscu
,
S.
,
Lee
,
A.
,
Luo
,
Y.
,
Teslovich
,
N.
,
Worthington
,
J.
,
Martin
,
J.
,
Huizinga
,
T.
,
Klareskog
,
L.
 et al. (
2018
)
Fine-mapping and functional studies highlight potential causal variants for rheumatoid arthritis and type 1 diabetes
.
Nat. Genet.
,
50
,
1366
1374
.

124.

Vernot
,
B.
,
Stergachis
,
A.B.
,
Maurano
,
M.T.
,
Vierstra
,
J.
,
Neph
,
S.
,
Thurman
,
R.E.
,
Stamatoyannopoulos
,
J.A.
and
Akey
,
J.M.
(
2012
)
Personal and population genomics of human regulatory variation
.
Genome Res.
,
22
,
1689
1697
.

125.

Consortium
,
E.P.
(
2012
)
An integrated encyclopedia of DNA elements in the human genome
.
Nature
,
489
,
57
74
.

126.

Roadmap Epigenomics Consortium
,
Kundaje
,
A.
,
Meuleman
,
W.
 et al. (
2015
)
Integrative analysis of 111 reference human epigenomes
.
Nature
,
518
,
317
330
.

127.

GTEx Consortium, Laboratory, Data Analysis & Coordinating Center (LDACC)—Analysis Working Group, Statistical Methods groups—Analysis Working Group
 et al. (
2017
)
Genetic effects on gene expression across human tissues
.
Nature
,
550
,
204
213
.

128.

Dimas
,
A.S.
,
Deutsch
,
S.
,
Stranger
,
B.E.
,
Montgomery
,
S.B.
,
Borel
,
C.
,
Attar-Cohen
,
H.
,
Ingle
,
C.
,
Beazley
,
C.
,
Arcelus
,
M.G.
,
Sekowska
,
M.
 et al. (
2009
)
Common regulatory variation impacts gene expression in a cell type-dependent manner
.
Science
,
325
,
1246
1250
.

129.

Zhang
,
T.
,
Choi
,
J.
,
Kovacs
,
M.A.
,
Shi
,
J.
,
Xu
,
M.
, Program, N.C.S., Melanoma Meta-Analysis, C.
Goldstein
,
A.M.
,
Trower
,
A.J.
,
Bishop
,
D.T.
 et al. (
2018
)
Cell-type–specific eQTL of primary melanocytes facilitates identification of melanoma susceptibility genes
.
Genome Res
,
28
, 1621–1635.

130.

Inoue
,
F.
and
Ahituv
,
N.
(
2015
)
Decoding enhancers using massively parallel reporter assays
.
Genomics
,
106
,
159
164
.

131.

Gasperini
,
M.
,
Hill
,
A.J.
,
McFaline-Figueroa
,
J.L.
,
Martin
,
B.
,
Kim
,
S.
,
Zhang
,
M.D.
,
Jackson
,
D.
,
Leith
,
A.
,
Schreiber
,
J.
,
Noble
,
W.S.
 et al. (
2019
)
A genome-wide framework for mapping gene regulation via cellular genetic screens
.
Cell
,
176
,
377
390 e19
.

132.

Mumbach
,
M.R.
,
Rubin
,
A.J.
,
Flynn
,
R.A.
,
Dai
,
C.
,
Khavari
,
P.A.
,
Greenleaf
,
W.J.
and
Chang
,
H.Y.
(
2016
)
HiChIP: efficient and sensitive analysis of protein-directed genome architecture
.
Nat. Methods
,
13
,
919
922
.

133.

Mifsud
,
B.
,
Tavares-Cadete
,
F.
,
Young
,
A.N.
,
Sugar
,
R.
,
Schoenfelder
,
S.
,
Ferreira
,
L.
,
Wingett
,
S.W.
,
Andrews
,
S.
,
Grey
,
W.
,
Ewels
,
P.A.
 et al. (
2015
)
Mapping long-range promoter contacts in human cells with high-resolution capture Hi-C
.
Nat. Genet.
,
47
,
598
606
.

134.

Li
,
G.
,
Ruan
,
X.
,
Auerbach
,
R.K.
,
Sandhu
,
K.S.
,
Zheng
,
M.
,
Wang
,
P.
,
Poh
,
H.M.
,
Goh
,
Y.
,
Lim
,
J.
,
Zhang
,
J.
 et al. (
2012
)
Extensive promoter-centered chromatin interactions provide a topological basis for transcription regulation
.
Cell
,
148
,
84
98
.

135.

Gasperini
,
M.
,
Tome
,
J.M.
and
Shendure
,
J.
(
2020
)
Towards a comprehensive catalogue of validated and target-linked human enhancers
.
Nat. Rev. Genet.
,
21
,
292
310
.

136.

Lin
,
J.
and
Kiran
,
M.
(
2018
)
From genotype to phenotype
.
Circulation: Genomic and Precision Medicine
,
11
,
e001946
.

137.

Ju, D. and Mathieson, I. (2021) The evolution of skin pigmentation-associated variation in West Eurasia.

Proc Natl Acad Sci USA
 
118
, e2009227118.

138.

Mathieson
,
I.
,
Lazaridis
,
I.
,
Rohland
,
N.
,
Mallick
,
S.
,
Patterson
,
N.
,
Roodenberg
,
S.A.
,
Harney
,
E.
,
Stewardson
,
K.
,
Fernandes
,
D.
,
Novak
,
M.
 et al. (
2015
)
Genome-wide patterns of selection in 230 ancient Eurasians
.
Nature
,
528
,
499
503
.

139.

Wang
,
K.
,
Goldstein
,
S.
,
Bleasdale
,
M.
,
Clist
,
B.
,
Bostoen
,
K.
,
Bakwa-Lufu
,
P.
,
Buck
,
L.T.
,
Crowther
,
A.
,
Dème
,
A.
,
McIntosh
,
R.J.
 et al. (
2020
)
Ancient genomes reveal complex patterns of population movement, interaction, and replacement in sub-Saharan Africa
.
Sci. Adv.
,
6
,
eaaz0183
.

140.

Cameron
,
D.L.
,
Stefano
,
L.D.
and
Papenfuss
,
A.T.
(
2019
)
Comprehensive evaluation and characterisation of short read general-purpose structural variant calling software
.
Nat. Commun.
,
10
,
3240
.

141.

Kosugi
,
S.
,
Momozawa
,
Y.
,
Liu
,
X.
,
Terao
,
C.
,
Kubo
,
M.
and
Kamatani
,
Y.
(
2019
)
Comprehensive evaluation of structural variation detection algorithms for whole genome sequencing
.
Genome Biol.
,
20
,
117
.

142.

Amarasinghe
,
S.L.
,
Su
,
S.
,
Dong
,
X.
,
Zappia
,
L.
,
Ritchie
,
M.E.
and
Gouil
,
Q.
(
2020
)
Opportunities and challenges in long-read sequencing data analysis
.
Genome Biol.
,
21
,
30
.

Author notes

Yuanqing Feng and Michael A. McQuillan contributed equally.

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model)