Abstract

CD36 and GNAT3 mediate taste responses, with CD36 acting as a lipid detector and GNAT3 acting as the α subunit of gustducin, a G protein governing sweet, savory, and bitter transduction. Strikingly, the genes encoding CD36 and GNAT3 are genomically superimposed, with CD36 completely encompassing GNAT3. To characterize genetic variation across the CD36-GNAT3 region, its implications for phenotypic diversity, and its recent evolution, we analyzed from ~2,500 worldwide subjects sequenced by the 1000 Genomes Project (1000GP). CD36-GNAT3 harbored extensive diversity including 8,688 single-nucleotide polymorphisms (SNPs), 414 indels, and other complex variants. Sliding window analyses revealed that nucleotide diversity and population differentiation across CD36-GNAT3 were consistent with genome-wide trends in the 1000GP (π = 0.10%, P = 0.64; FST = 9.0%, P = 0.57). In addition, functional predictions using SIFT and PolyPhen-2 identified 60 variants likely to alter protein function, and they were in weak linkage disequilibrium (r2 < 0.17), suggesting their effects are largely independent. However, the frequencies of predicted functional variants were low (P¯ = 0.0013), indicating their contributions to phenotypic variance on population scales are limited. Tests using Tajima’s D statistic revealed that pressures from natural selection have been relaxed across most of CD36-GNAT3 during its recent history (0.39 < P < 0.67). However, CD36 exons showed signs of local adaptation consistent with prior reports (P < 0.035). Thus, CD36 and GNAT3 harbor numerous variants predicted to affect taste sensitivity, but most are rare and phenotypic variance on a population level is likely mediated by a small number of sites.

Introduction

Taste perception is a fundamental mechanism of diet selection and control. By allowing animals to evaluate the nutritional properties and safety of foods before they are consumed, taste provides a powerful means of enhancing health and evolutionary fitness (Lindemann 2001; Reed and Knaapila 2010; Roper and Chaudhari 2017). For instance, bitter sensations, which are triggered by plant toxins, signal the presence of noxious components, allowing avoidance. Sweet sensations, which are triggered by sugars, signal carbohydrate richness. Salty, sour, and umami/savory sensations signal the presence of electrolytes, acidity indicative of ripeness, and protein content. Together these modalities provide a nutrient profile that can be used to guide intake, a major foraging advantage. The significance of this role is evident in the diversity of taste receptors found throughout vertebrates (Fischer et al. 2005; Wooding et al. 2006; Zhao et al. 2010, 2015; Wooding 2011; Jiang et al. 2012; Baldwin et al. 2014; Feng et al. 2014; Li and Zhang 2014; Antinucci and Risso 2017; Behrens et al. 2020).

A key feature of taste perception in humans is that it varies due to polymorphism in genes encoding receptors and other signaling components (Kim et al. 2004; Bachmanov et al. 2014). For example, TAS2R38, which encodes a bitter receptor, harbors alleles associated with taste responses to goitrin, a thyroid toxin synthesized by plants in the Brassicaceae family (Wooding et al. 2010). Similar associations are found between variants in TAS1R3 (an umami receptor subunit) and monosodium glutamate and between CA6 variants (a salivary carbonic anhydrase) and sodium salt (Chen et al. 2009; Feeney and Hayes 2014). Polymorphism in taste pathways also associates with preferences and consumption of foods such as alcoholic beverages and cruciferous vegetables, and health measures such as body mass index, susceptibility to colorectal cancers, and kidney disease (Greene 1974; Basson et al. 2005; Wooding et al. 2012; Behrens et al. 2013; Allen et al. 2014; Hayes et al. 2015; Choi et al. 2016; Barontini et al. 2017). These affect evolutionary fitness, and taste genes in humans harbor signatures of natural selection including evidence of local adaptation, balancing pressures, and purifying effects (Wooding et al. 2004; Drayna 2005; Kim et al. 2005, 2006; Campbell et al. 2012, 2014; Risso et al. 2016, 2018). Thus, modern patterns of variation in taste sensitivity, nutrition, and health reflect ancient evolutionary influences on taste.

Mounting evidence suggests that human taste abilities extend to the detection of fats, particularly long-chain fatty acids (LCFAs), and that fat taste sensitivity varies from person to person as the result of genetic polymorphism. In psychophysical assays, subjects are capable of discriminating fat content in controlled preparations even when non-gustatory cues are masked, supporting a role for taste (Chalé-Rush et al. 2007; Cartoni et al. 2010; Mattes 2011). In addition, like other taste signals, neural signals generated by oral fat exposure originate in taste receptor cells and travel via the chorda tympani and glossopharyngeal nerves in mice (Gaillard et al. 2008). Oral fat exposure also activates the brain’s insular cortex, which is activated during sweet perception (De Araujo and Rolls 2004). Several lines of evidence indicate that CD36, a fatty acid translocase, is the receptor accounting for these effects. It localizes to taste receptor cells, natively responds to fatty acids in vitro, and knockout of CD36 in rats and mice alters their preferences for fat-containing solutions and foods (Laugerette et al. 2005). CD36 also harbors alleles associated with both orosensory detection of fats and preferences for them (Keller et al. 2012; Pepino et al. 2012). In addition, CD36 plays known roles beyond taste, contributing to immune system function, lipid metabolism, and cell adhesion (Pepino et al. 2014). These findings raise questions about the extent of genetic polymorphism at CD36 and its effects on fat perception and other phenotypes.

The potential contributions of CD36 to fat taste also raise evolutionary questions. The high nutritional value of lipids, which are calorically rich but environmentally scarce, suggests that CD36’s role as a taste sensor placed it under selective pressures in the course of human evolution. In particular, humans’ population expansion and migration out of Africa 50–60 thousand years ago introduced them to new physical and nutritional environments that likely altered the advantages of fat perception and metabolism. For instance, they may have been shaped by factors such as the accessibility of fats when hunting and foraging, or climate, which poses thermoregulatory challenges. They could also have arisen from CD36’s non-gustatory roles in processes such as cell adhesion, which makes it vulnerable to exploitation by pathogens (Silverstein and Febbraio 2009). Such pressures leave signatures in genetic diversity including effects on allele frequencies and population differentiation (Bamshad and Wooding 2003). Thus, modern patterns of diversity in CD36 may provide clues to the evolutionary factors driving responses to fats.

Strikingly, CD36 genomically encompasses a second gene participating in taste perception, GNAT3 (Fig. 1). GNAT3 encodes a G-protein subunit mediating detection of sweet, savory, and bitter substances and, like CD36, harbors variants associated with taste sensitivity (Fushan et al. 2010; Farook et al. 2012; Pepino et al. 2012). GNAT3 plays non-gustatory roles as well, such as the detection of foreign compounds in the gut and airways (Egan and Margolskee 2008; Deshpande et al. 2010). The nested arrangement of the two genes suggests that patterns of diversity in them may be correlated due to their genetic linkage and shared evolutionary histories. If so, GNAT3-mediated taste responses (bitter, sweet, and umami) and CD36-mediated responses (fat) may be correlated. However, the CD36-GNAT3 region is sufficiently large (~305 kb) that linkage disequilibrium (LD) may not be high across its entirety. LD can also be shaped by natural selection, which can affect its range and magnitude. Establishing the structure of genetic variation across CD36-GNAT3 has the potential to reveal the extent of such effects.

Genomic organization of CD36 and GNAT3. The CD36-GNAT3 region is ~305 kb in length. GNAT3 is nested within CD36, with exons 1–3 located in CD36 intron 3 and exons 4–8 located in CD36 exon 1.
Fig. 1.

Genomic organization of CD36 and GNAT3. The CD36-GNAT3 region is ~305 kb in length. GNAT3 is nested within CD36, with exons 1–3 located in CD36 intron 3 and exons 4–8 located in CD36 exon 1.

We addressed these issues in a population genetic analysis of CD36-GNAT3 in >2,500 subjects from the 1000 Genomes Project (1000GP; Li 2011; The 1000 Genomes Project Consortium 2015). To establish the extent of diversity at CD36-GNAT3 and its implications for genotype–phenotype associations, we comprehensively identified variable sites in the region, their allele frequencies in worldwide populations, and their linkage structure. We then used computational prediction to detect sites likely to have functional effects, and evolutionary analyses to determine the role of natural selection in shaping this variability. Our results shed light on the architecture of diversity in CD36 and GNAT3, its potential contributions to taste and metabolism, and its evolutionary origins.

Methods

We examined genetic variation across CD36-GNAT3 in 2,504 subjects included in Phase 3 of the 1000GP (The 1000 Genomes Project Consortium 2015). The 1000GP subjects comprise a random, demographically representative sample of 26 worldwide populations in five superpopulations, providing a diverse hierarchical perspective on human genetic variation (Table 1).

Table 1.

Population sample.

Super populationPopulation
Africa (N = 661)African Caribbeans in Barbados (N = 96)
Americans of African Ancestry in SW USA (N = 61)
Esan in Nigeria (N = 99)
Gambian in Western Divisions in the Gambia (N = 113)
Luhya in Webuye, Kenya (N = 99)
Mende in Sierra Leone (N = 85)
Yoruba in Ibadan, Nigeria (N = 108)
Americas (N = 347)Colombians from Medellin, Colombia (N = 94)
Mexican Ancestry from Los Angeles, USA (N = 64)
Peruvian from Lima, Peru (N = 85)
Puerto Rican in Puerto Rico (N = 104)
East Asia (N = 504)Chinese Dai in Xishuangbanna, China (N = 93)
Han Chinese in Beijing, China (N = 103)
Japanese in Toyko, Japan (N = 104)
Kinh in Ho Chi Minh City, Vietnam (N = 99)
Southern Han Chinese (N = 105)
Europe (N = 503)British in England and Scotland (N = 91)
Finnish in Finland (N = 99)
Iberian population in Spain (N = 107)
Toscani in Italy (N = 107)
Utah residents (CEPH) with European ancestry (N = 99)
South Asia (N = 489)Bengali from Bangladesh (N = 86)
Gujarati Indian from Houston, TX (N = 103)
Indian Telegu from the UK (N = 102)
Punjabi from Lahore, Pakistan (N = 96)
Sri Lankan Tamil from the UK (N = 102)
Super populationPopulation
Africa (N = 661)African Caribbeans in Barbados (N = 96)
Americans of African Ancestry in SW USA (N = 61)
Esan in Nigeria (N = 99)
Gambian in Western Divisions in the Gambia (N = 113)
Luhya in Webuye, Kenya (N = 99)
Mende in Sierra Leone (N = 85)
Yoruba in Ibadan, Nigeria (N = 108)
Americas (N = 347)Colombians from Medellin, Colombia (N = 94)
Mexican Ancestry from Los Angeles, USA (N = 64)
Peruvian from Lima, Peru (N = 85)
Puerto Rican in Puerto Rico (N = 104)
East Asia (N = 504)Chinese Dai in Xishuangbanna, China (N = 93)
Han Chinese in Beijing, China (N = 103)
Japanese in Toyko, Japan (N = 104)
Kinh in Ho Chi Minh City, Vietnam (N = 99)
Southern Han Chinese (N = 105)
Europe (N = 503)British in England and Scotland (N = 91)
Finnish in Finland (N = 99)
Iberian population in Spain (N = 107)
Toscani in Italy (N = 107)
Utah residents (CEPH) with European ancestry (N = 99)
South Asia (N = 489)Bengali from Bangladesh (N = 86)
Gujarati Indian from Houston, TX (N = 103)
Indian Telegu from the UK (N = 102)
Punjabi from Lahore, Pakistan (N = 96)
Sri Lankan Tamil from the UK (N = 102)
Table 1.

Population sample.

Super populationPopulation
Africa (N = 661)African Caribbeans in Barbados (N = 96)
Americans of African Ancestry in SW USA (N = 61)
Esan in Nigeria (N = 99)
Gambian in Western Divisions in the Gambia (N = 113)
Luhya in Webuye, Kenya (N = 99)
Mende in Sierra Leone (N = 85)
Yoruba in Ibadan, Nigeria (N = 108)
Americas (N = 347)Colombians from Medellin, Colombia (N = 94)
Mexican Ancestry from Los Angeles, USA (N = 64)
Peruvian from Lima, Peru (N = 85)
Puerto Rican in Puerto Rico (N = 104)
East Asia (N = 504)Chinese Dai in Xishuangbanna, China (N = 93)
Han Chinese in Beijing, China (N = 103)
Japanese in Toyko, Japan (N = 104)
Kinh in Ho Chi Minh City, Vietnam (N = 99)
Southern Han Chinese (N = 105)
Europe (N = 503)British in England and Scotland (N = 91)
Finnish in Finland (N = 99)
Iberian population in Spain (N = 107)
Toscani in Italy (N = 107)
Utah residents (CEPH) with European ancestry (N = 99)
South Asia (N = 489)Bengali from Bangladesh (N = 86)
Gujarati Indian from Houston, TX (N = 103)
Indian Telegu from the UK (N = 102)
Punjabi from Lahore, Pakistan (N = 96)
Sri Lankan Tamil from the UK (N = 102)
Super populationPopulation
Africa (N = 661)African Caribbeans in Barbados (N = 96)
Americans of African Ancestry in SW USA (N = 61)
Esan in Nigeria (N = 99)
Gambian in Western Divisions in the Gambia (N = 113)
Luhya in Webuye, Kenya (N = 99)
Mende in Sierra Leone (N = 85)
Yoruba in Ibadan, Nigeria (N = 108)
Americas (N = 347)Colombians from Medellin, Colombia (N = 94)
Mexican Ancestry from Los Angeles, USA (N = 64)
Peruvian from Lima, Peru (N = 85)
Puerto Rican in Puerto Rico (N = 104)
East Asia (N = 504)Chinese Dai in Xishuangbanna, China (N = 93)
Han Chinese in Beijing, China (N = 103)
Japanese in Toyko, Japan (N = 104)
Kinh in Ho Chi Minh City, Vietnam (N = 99)
Southern Han Chinese (N = 105)
Europe (N = 503)British in England and Scotland (N = 91)
Finnish in Finland (N = 99)
Iberian population in Spain (N = 107)
Toscani in Italy (N = 107)
Utah residents (CEPH) with European ancestry (N = 99)
South Asia (N = 489)Bengali from Bangladesh (N = 86)
Gujarati Indian from Houston, TX (N = 103)
Indian Telegu from the UK (N = 102)
Punjabi from Lahore, Pakistan (N = 96)
Sri Lankan Tamil from the UK (N = 102)

The genomic structure of the CD36-GNAT3 region was determined from the Ensembl GRch37 human genome assembly, the reference for the 1000GP. These placed CD36 (Ensembl ENSG00000135218, ENST00000435819) at position 7:79998891–7:80308593 (~305 kb) and GNAT3 (ENSG00000214415, ENST00000398291) at position 7:80087987–7:80141336 (53 kb), with GNAT3 located in introns 1 and 2 of CD36. Data for the region were extracted from 1000GP databases in variant call format (VCF) using the Tabix software package (Li 2011).

Genetic variation was assessed with respect to three factors, allelic polymorphism, population substructure, and LD. Nucleotide diversity (π), the mean pairwise nucleotide difference among sequences normalized to sequence length, was calculated across CD36-GNAT3 as well as separately for CD36 exons, CD36 introns, GNAT3 exons, and GNAT3 introns (Tajima 1983). Population substructure was measured within and across 1000GP superpopulations using Weir and Hill’s weighted FST (Weir and Hill 2002). These calculations were performed using VCFtools and the R packages PopGenome, pegas, hierfstat, and adgenet (Goudet 2004; Paradis 2010; Danecek et al. 2011; Jombart and Ahmed 2011; Pfeifer et al. 2014). LD was assessed for variants with frequencies >0.05 using two measures, and r2 (Mueller 2004; Slatkin 2008). , a measure of correlation between sites relative to the maximum possible given their allele frequencies, was used to determine the extent to which recombination has shaped diversity across the region. A raw measure of correlation among genetic markers, r2, was used to determine the extent to which sites are expected to exhibit similar genotype–phenotype associations. These measures were calculated and visualized using the VariantAnnotation and Ldheatmap packages in the R statistical analysis environment and Bioconductor library (Ihaka and Gentleman 1996; Gentleman et al. 2004; Graham et al. 2006; Obenchain et al. 2014).

Two algorithms were used to predict the potential functional impact of exon variants, PolyPhen-2 and SIFT (Kumar et al. 2009; Adzhubei et al. 2010). PolyPhen-2 predicts the impact of amino acid changes on protein function on the basis of the location of the changed site within the protein structure, level of conservation relative to homologous genes, and the biochemical characteristics of the substituted amino acids. It denotes the impact of substitutions on a scale from 0.0 (benign) to 1.0 (damaging). SIFT predicts whether amino acid substitutions affect protein function on the basis of probabilities estimated from gene homologs. It denotes impact on a scale from 0.0 to 1.0, categorizing scores below 0.05 as deleterious and higher scores as tolerated. Both scores were obtained using the Variant Effect Predictor (VEP) software package (McLaren et al. 2016). Regulatory variants were identified by using VEP to query the Ensembl Regulatory Build.

Tests for natural selection were performed using Tajima’s D statistic (Tajima 1989). D compares the number of variable sites and mean nucleotide difference between alleles in a sample, which are differentially affected by selective processes. It is designed to test for selective effects nonrecombining genomic regions but is applicable with elevated conservativeness to regions with recombination. These analyses were performed using the PopGenome R package. As with pi estimates, these were calculated both overall and with respect to CD36 and GNAT3 exons and introns.

Because the vast majority of the human genome (>98%) is noncoding it is can be assumed to be evolving neutrally or nearly so with respect to natural selection. Therefore, to provide a neutral baseline for evaluating diversity measures in the CD36-GNAT3, we generated empirical distributions for three measures (π, FST, and D) using sliding window analyses. These were obtained by iteratively calculating each measure in ~270,000 adjacent 10 kb windows spanning the length of the 1000GP genomes, excluding known unstable and repetitive regions such as telomeres. The probability of the observed values given genome-wide trends was then determined by comparing the values observed in CD36-GNAT3 with their genome-wide distributions. We denoted P-values from these empirical tests PE to distinguish them from P-values obtained in parametric tests.

Results

The CD36-GNAT3 region harbored extensive variation. A total of 9,111 polymorphic sites were identified. The majority (95.3%) were single-nucleotide polymorphisms (SNPs; 8,653 biallelic and 32 multiallelic). A smaller number (4.5%) were insertion/deletion (indel) polymorphisms (390 biallelic and 21 multiallelic). The remainder (<0.2%) were rare complex variants, including three sites with both SNP and indel alleles, 11 copy number variants (CNVs), and one Alu insertion. CD36 exons, which totaled 2,365 bp in length, contained 112 SNPs (all biallelic) and 8 indels, which occurred in 7 exons (Fig. 2). Four of the eight were frameshift deletions, two were 1 bp insertions in untranslated regions, and two were in-frame deletions. GNAT3 exons, which totaled 1,159 bp in length, contained 28 SNPs (all biallelic), with one indel in the 5ʹ-untranslated region of exon 1. The number of variants also differed among exons. No polymorphism was found in GNAT3 exon 2, 17 SNPs were present in CD36 exon 17, and the mean across exons was 6. The number of SNPs per nucleotide ranged from 0.0 (GNAT3 exon 2) to 0.1 (CD36 exon 12), with an average of 0.037.

Variant types and their frequencies across CD36 and GNAT3 exons. Both GNAT3 and CD36 harbored extensive variation, including numerous variants likely to affect function.
Fig. 2.

Variant types and their frequencies across CD36 and GNAT3 exons. Both GNAT3 and CD36 harbored extensive variation, including numerous variants likely to affect function.

Among the 140 SNPs in CD36 and GNAT3 exons, PolyPhen and SIFT detected 57 likely to alter function in the CD36 and GNAT3 proteins. Fifty-one had PolyPhen scores of Possibly or Probably Damaging and 51 had SIFT scores of Deleterious. The scores were largely in agreement between the two measures. Forty variants were scored as Possibly/Probably Damaging by PolyPhen and Deleterious by SIFT, and a further 15 were sites scored as Benign by PolyPhen and Tolerated by SIFT. Predictions disagreed for 12 sites, with 6 scored as tolerated by SIFT but Possibly/Probably Damaging by PolyPhen, and 6 scored as Deleterious by SIFT but tolerated by PolyPhen. Sites without SIFT and PolyPhen scores but likely to have functional impact were also found. Eighteen variants in CD36 were indels causing frameshifts (4 sites) stop gains (10 sites), stop losses (2 sites), or in-frame deletions (2 sites). GNAT3 harbored one stop gain (in exon 3) and one stop loss (in exon 8). For further analyses, the 60 sites with variants scored as possibly damaging by PolyPhen and Deleterious by SIFT, frameshifts, and those occurring in start or stop codons were denoted putatively high impact (PHI) sites, and their derived alleles denoted PHI alleles (Table 2).

Table 2.

Putative high impact (PHI) sites.

GenersidExonVariant typeReference codonAlternate codonReference amino acidAlternate amino acid
GNAT3rs5335248661NsaCcaAcTN
GNAT3rs5730823243SgtTgtAgLStop
GNAT3rs2000104944NsGatAatDN
GNAT3rs5711203134NscTgcCgLP
GNAT3rs5700301585NsttGttTLF
GNAT3rs1868772328SlTaaCaaStopQ
GNAT3rs5349021398NsTtcGtcFV
CD36rs5598762706NsgGggAgGE
CD36rs753269247NsCctTctPS
CD36rs1390670667SgtgGtgAWStop
CD36rs1500376127NsaCgaTgTM
CD36rs5345778787NstgGtgCWC
CD36rs5454892047SgCagTagQStop
CD36rs5561812107NsaTcaAcIN
CD36rs5719750657FsAaaaaKna
CD36rs5744167057SgtaCtaGYStop
CD36rs709617158NscGtcCtRP
CD36rs2017653318NstCatTaSL
CD36rs5485078598NsTcaCcaSP
CD36rs5564386558SgGaaTaaEStop
CD36rs5630978478NsCtcTtcLF
CD36rs5722958238FsaACaNna
CD36rs2017593079NsTggCggWR
CD36rs5685039179NsgGcgTcGV
CD36rs5699597769SgtaTtaGYStop
CD36rs3577609510NsgGagAaGE
CD36rs37382957810SgAaaTaaKStop
CD36rs20006732210NsGgaAgaGR
CD36rs53515093610NsgGtgTtGV
CD36rs20124576610SgaggagGTAAgRStop
CD36rs14917814211NsaCaaTaTI
CD36rs14998598811SgtgCtgACStop
CD36rs55773273611NsaTtaAtIN
CD36rs14218640412NsTttGttFV
CD36rs14590880312NscCacTaPL
CD36rs19968163112NsaGgaCgRT
CD36rs20115545212NscCtcAtPH
CD36rs53554916812NstTgtCgLS
CD36rs321193813SgtaTtaGYStop
CD36rs20075778813NsaGaaTaRI
CD36rs55401917013NsgAcgGcDG
CD36rs55811506713FsctGctLna
CD36rs56749185613NstGttTtCF
CD36rs57155318413IfaAAGaaaaaKEK
CD36rs14790373514NsCatTatHY
CD36rs37070121014NsGcaCcaAP
CD36rs37188408214NscAtcGtHR
CD36rs37631104514NsCcaTcaPS
CD36rs56497157114NsCctTctPS
CD36rs14891022715NsCggTggRW
CD36rs20019448615NsaCtaGtTS
CD36rs20090646215NsActCctTP
CD36rs20135571115NscAgcTgQL
CD36rs55160778415FsGcacaAna
CD36rs55056580016IftaTATTGTGCCTATttatYIVPIY
CD36rs20155860817NsGgtAgtGS
CD36rs55016379917SlTaaGaaStopE
CD36rs55991652817NsgGtgCtGA
CD36rs56377233717SgCaaTaaQStop
CD36rs57017191717SltAatCaStopS
GenersidExonVariant typeReference codonAlternate codonReference amino acidAlternate amino acid
GNAT3rs5335248661NsaCcaAcTN
GNAT3rs5730823243SgtTgtAgLStop
GNAT3rs2000104944NsGatAatDN
GNAT3rs5711203134NscTgcCgLP
GNAT3rs5700301585NsttGttTLF
GNAT3rs1868772328SlTaaCaaStopQ
GNAT3rs5349021398NsTtcGtcFV
CD36rs5598762706NsgGggAgGE
CD36rs753269247NsCctTctPS
CD36rs1390670667SgtgGtgAWStop
CD36rs1500376127NsaCgaTgTM
CD36rs5345778787NstgGtgCWC
CD36rs5454892047SgCagTagQStop
CD36rs5561812107NsaTcaAcIN
CD36rs5719750657FsAaaaaKna
CD36rs5744167057SgtaCtaGYStop
CD36rs709617158NscGtcCtRP
CD36rs2017653318NstCatTaSL
CD36rs5485078598NsTcaCcaSP
CD36rs5564386558SgGaaTaaEStop
CD36rs5630978478NsCtcTtcLF
CD36rs5722958238FsaACaNna
CD36rs2017593079NsTggCggWR
CD36rs5685039179NsgGcgTcGV
CD36rs5699597769SgtaTtaGYStop
CD36rs3577609510NsgGagAaGE
CD36rs37382957810SgAaaTaaKStop
CD36rs20006732210NsGgaAgaGR
CD36rs53515093610NsgGtgTtGV
CD36rs20124576610SgaggagGTAAgRStop
CD36rs14917814211NsaCaaTaTI
CD36rs14998598811SgtgCtgACStop
CD36rs55773273611NsaTtaAtIN
CD36rs14218640412NsTttGttFV
CD36rs14590880312NscCacTaPL
CD36rs19968163112NsaGgaCgRT
CD36rs20115545212NscCtcAtPH
CD36rs53554916812NstTgtCgLS
CD36rs321193813SgtaTtaGYStop
CD36rs20075778813NsaGaaTaRI
CD36rs55401917013NsgAcgGcDG
CD36rs55811506713FsctGctLna
CD36rs56749185613NstGttTtCF
CD36rs57155318413IfaAAGaaaaaKEK
CD36rs14790373514NsCatTatHY
CD36rs37070121014NsGcaCcaAP
CD36rs37188408214NscAtcGtHR
CD36rs37631104514NsCcaTcaPS
CD36rs56497157114NsCctTctPS
CD36rs14891022715NsCggTggRW
CD36rs20019448615NsaCtaGtTS
CD36rs20090646215NsActCctTP
CD36rs20135571115NscAgcTgQL
CD36rs55160778415FsGcacaAna
CD36rs55056580016IftaTATTGTGCCTATttatYIVPIY
CD36rs20155860817NsGgtAgtGS
CD36rs55016379917SlTaaGaaStopE
CD36rs55991652817NsgGtgCtGA
CD36rs56377233717SgCaaTaaQStop
CD36rs57017191717SltAatCaStopS
Table 2.

Putative high impact (PHI) sites.

GenersidExonVariant typeReference codonAlternate codonReference amino acidAlternate amino acid
GNAT3rs5335248661NsaCcaAcTN
GNAT3rs5730823243SgtTgtAgLStop
GNAT3rs2000104944NsGatAatDN
GNAT3rs5711203134NscTgcCgLP
GNAT3rs5700301585NsttGttTLF
GNAT3rs1868772328SlTaaCaaStopQ
GNAT3rs5349021398NsTtcGtcFV
CD36rs5598762706NsgGggAgGE
CD36rs753269247NsCctTctPS
CD36rs1390670667SgtgGtgAWStop
CD36rs1500376127NsaCgaTgTM
CD36rs5345778787NstgGtgCWC
CD36rs5454892047SgCagTagQStop
CD36rs5561812107NsaTcaAcIN
CD36rs5719750657FsAaaaaKna
CD36rs5744167057SgtaCtaGYStop
CD36rs709617158NscGtcCtRP
CD36rs2017653318NstCatTaSL
CD36rs5485078598NsTcaCcaSP
CD36rs5564386558SgGaaTaaEStop
CD36rs5630978478NsCtcTtcLF
CD36rs5722958238FsaACaNna
CD36rs2017593079NsTggCggWR
CD36rs5685039179NsgGcgTcGV
CD36rs5699597769SgtaTtaGYStop
CD36rs3577609510NsgGagAaGE
CD36rs37382957810SgAaaTaaKStop
CD36rs20006732210NsGgaAgaGR
CD36rs53515093610NsgGtgTtGV
CD36rs20124576610SgaggagGTAAgRStop
CD36rs14917814211NsaCaaTaTI
CD36rs14998598811SgtgCtgACStop
CD36rs55773273611NsaTtaAtIN
CD36rs14218640412NsTttGttFV
CD36rs14590880312NscCacTaPL
CD36rs19968163112NsaGgaCgRT
CD36rs20115545212NscCtcAtPH
CD36rs53554916812NstTgtCgLS
CD36rs321193813SgtaTtaGYStop
CD36rs20075778813NsaGaaTaRI
CD36rs55401917013NsgAcgGcDG
CD36rs55811506713FsctGctLna
CD36rs56749185613NstGttTtCF
CD36rs57155318413IfaAAGaaaaaKEK
CD36rs14790373514NsCatTatHY
CD36rs37070121014NsGcaCcaAP
CD36rs37188408214NscAtcGtHR
CD36rs37631104514NsCcaTcaPS
CD36rs56497157114NsCctTctPS
CD36rs14891022715NsCggTggRW
CD36rs20019448615NsaCtaGtTS
CD36rs20090646215NsActCctTP
CD36rs20135571115NscAgcTgQL
CD36rs55160778415FsGcacaAna
CD36rs55056580016IftaTATTGTGCCTATttatYIVPIY
CD36rs20155860817NsGgtAgtGS
CD36rs55016379917SlTaaGaaStopE
CD36rs55991652817NsgGtgCtGA
CD36rs56377233717SgCaaTaaQStop
CD36rs57017191717SltAatCaStopS
GenersidExonVariant typeReference codonAlternate codonReference amino acidAlternate amino acid
GNAT3rs5335248661NsaCcaAcTN
GNAT3rs5730823243SgtTgtAgLStop
GNAT3rs2000104944NsGatAatDN
GNAT3rs5711203134NscTgcCgLP
GNAT3rs5700301585NsttGttTLF
GNAT3rs1868772328SlTaaCaaStopQ
GNAT3rs5349021398NsTtcGtcFV
CD36rs5598762706NsgGggAgGE
CD36rs753269247NsCctTctPS
CD36rs1390670667SgtgGtgAWStop
CD36rs1500376127NsaCgaTgTM
CD36rs5345778787NstgGtgCWC
CD36rs5454892047SgCagTagQStop
CD36rs5561812107NsaTcaAcIN
CD36rs5719750657FsAaaaaKna
CD36rs5744167057SgtaCtaGYStop
CD36rs709617158NscGtcCtRP
CD36rs2017653318NstCatTaSL
CD36rs5485078598NsTcaCcaSP
CD36rs5564386558SgGaaTaaEStop
CD36rs5630978478NsCtcTtcLF
CD36rs5722958238FsaACaNna
CD36rs2017593079NsTggCggWR
CD36rs5685039179NsgGcgTcGV
CD36rs5699597769SgtaTtaGYStop
CD36rs3577609510NsgGagAaGE
CD36rs37382957810SgAaaTaaKStop
CD36rs20006732210NsGgaAgaGR
CD36rs53515093610NsgGtgTtGV
CD36rs20124576610SgaggagGTAAgRStop
CD36rs14917814211NsaCaaTaTI
CD36rs14998598811SgtgCtgACStop
CD36rs55773273611NsaTtaAtIN
CD36rs14218640412NsTttGttFV
CD36rs14590880312NscCacTaPL
CD36rs19968163112NsaGgaCgRT
CD36rs20115545212NscCtcAtPH
CD36rs53554916812NstTgtCgLS
CD36rs321193813SgtaTtaGYStop
CD36rs20075778813NsaGaaTaRI
CD36rs55401917013NsgAcgGcDG
CD36rs55811506713FsctGctLna
CD36rs56749185613NstGttTtCF
CD36rs57155318413IfaAAGaaaaaKEK
CD36rs14790373514NsCatTatHY
CD36rs37070121014NsGcaCcaAP
CD36rs37188408214NscAtcGtHR
CD36rs37631104514NsCcaTcaPS
CD36rs56497157114NsCctTctPS
CD36rs14891022715NsCggTggRW
CD36rs20019448615NsaCtaGtTS
CD36rs20090646215NsActCctTP
CD36rs20135571115NscAgcTgQL
CD36rs55160778415FsGcacaAna
CD36rs55056580016IftaTATTGTGCCTATttatYIVPIY
CD36rs20155860817NsGgtAgtGS
CD36rs55016379917SlTaaGaaStopE
CD36rs55991652817NsgGtgCtGA
CD36rs56377233717SgCaaTaaQStop
CD36rs57017191717SltAatCaStopS

Potential regulatory variation was also found. VEP identified 48 regulatory regions and 10 features as transcription factor binding sites, which together harbored 685 variants annotated by Ensembl as expression modifiers. One hundred and thirty-one of these were most proximal to GNAT3 exons, and 554 were most proximal to CD36 exons. The majority (667; >95%) were SNPs, but other variant types were also present. Ten short insertions and 16 short deletions were found, along with six copy numbers and structural variants ranging from ~2 kb to ~130 kb in length, which spanned numerous regulatory and transcription factor binding sites.

As expected given that the majority of the CD36-GNAT3 region is intronic, most variants (7,827) occurred in noncoding regions (Table 3). A smaller number, 686, occurred in regulatory regions. In addition, consistent with their relative lengths, CD36 harbored more segregating sites than did GNAT3 (112 vs. 28). Derived allele frequencies at noncoding sites ranged from 0.0002 (singletons) to 0.9998 with a mean of 0.033. Consistent with the low mean, the majority of alleles were rare, with 92% having frequencies below 0.05, and 83% having frequencies below 0.01 (Table 4). Alleles with intermediate frequencies accounted for a small proportion of sites, with 4% having frequencies between 0.25 and 0.75. These patterns extended to CD36 and GNAT3 exons and regulatory regions, with >80% of alleles having frequencies below 0.01 in all three cases.

Table 3.

Genetic diversity in populations and superpopulations. Numbers in parentheses indicate nonsynonymous variants.

Site categoryWorldwideAfricaAmericasEast AsiaEuropeSouth Asia
S
Noncoding7,8273,8852,6242,1242,1202,443
CD36 exons112 (98)43 (38)22 (19)30 (27)21 (17)24 (19)
GNAT3 exons28 (15)10 (6)10 (6)5 (2)6 (2)9 (5)
PHI sites60241024814
Regulatory sites686333237180194227
π (%)
Noncoding0.1000.1210.0870.0920.0790.087
CD36 exons0.0450.0540.0380.0420.0410.042
GNAT3 exons0.0360.0380.0350.0260.0330.034
PHI sites
Regulatory sites
FST
Noncoding0.0900.0150.0190.0150.0020.005
CD36 exons0.0450.0790.0000.0110.0040.000
GNAT3 exons0.0930.0060.0300.0010.0070.015
PHI sites0.0660.1080.0000.0150.0000.007
Regulatory sites0.0800.0160.0310.0030.0010.007
Site categoryWorldwideAfricaAmericasEast AsiaEuropeSouth Asia
S
Noncoding7,8273,8852,6242,1242,1202,443
CD36 exons112 (98)43 (38)22 (19)30 (27)21 (17)24 (19)
GNAT3 exons28 (15)10 (6)10 (6)5 (2)6 (2)9 (5)
PHI sites60241024814
Regulatory sites686333237180194227
π (%)
Noncoding0.1000.1210.0870.0920.0790.087
CD36 exons0.0450.0540.0380.0420.0410.042
GNAT3 exons0.0360.0380.0350.0260.0330.034
PHI sites
Regulatory sites
FST
Noncoding0.0900.0150.0190.0150.0020.005
CD36 exons0.0450.0790.0000.0110.0040.000
GNAT3 exons0.0930.0060.0300.0010.0070.015
PHI sites0.0660.1080.0000.0150.0000.007
Regulatory sites0.0800.0160.0310.0030.0010.007
Table 3.

Genetic diversity in populations and superpopulations. Numbers in parentheses indicate nonsynonymous variants.

Site categoryWorldwideAfricaAmericasEast AsiaEuropeSouth Asia
S
Noncoding7,8273,8852,6242,1242,1202,443
CD36 exons112 (98)43 (38)22 (19)30 (27)21 (17)24 (19)
GNAT3 exons28 (15)10 (6)10 (6)5 (2)6 (2)9 (5)
PHI sites60241024814
Regulatory sites686333237180194227
π (%)
Noncoding0.1000.1210.0870.0920.0790.087
CD36 exons0.0450.0540.0380.0420.0410.042
GNAT3 exons0.0360.0380.0350.0260.0330.034
PHI sites
Regulatory sites
FST
Noncoding0.0900.0150.0190.0150.0020.005
CD36 exons0.0450.0790.0000.0110.0040.000
GNAT3 exons0.0930.0060.0300.0010.0070.015
PHI sites0.0660.1080.0000.0150.0000.007
Regulatory sites0.0800.0160.0310.0030.0010.007
Site categoryWorldwideAfricaAmericasEast AsiaEuropeSouth Asia
S
Noncoding7,8273,8852,6242,1242,1202,443
CD36 exons112 (98)43 (38)22 (19)30 (27)21 (17)24 (19)
GNAT3 exons28 (15)10 (6)10 (6)5 (2)6 (2)9 (5)
PHI sites60241024814
Regulatory sites686333237180194227
π (%)
Noncoding0.1000.1210.0870.0920.0790.087
CD36 exons0.0450.0540.0380.0420.0410.042
GNAT3 exons0.0360.0380.0350.0260.0330.034
PHI sites
Regulatory sites
FST
Noncoding0.0900.0150.0190.0150.0020.005
CD36 exons0.0450.0790.0000.0110.0040.000
GNAT3 exons0.0930.0060.0300.0010.0070.015
PHI sites0.0660.1080.0000.0150.0000.007
Regulatory sites0.0800.0160.0310.0030.0010.007
Table 4.

Allele frequency quantiles.

Frequency quantile
Site category<0.01<0.02<0.03<0.04<0.05<0.05
Noncoding0.830.040.020.020.010.09
CD36 exon0.960.010.000.010.000.02
GNAT3 exon0.960.000.000.000.000.04
PHI0.970.020.000.020.000.00
Regulatory0.830.050.020.010.000.09
Frequency quantile
Site category<0.01<0.02<0.03<0.04<0.05<0.05
Noncoding0.830.040.020.020.010.09
CD36 exon0.960.010.000.010.000.02
GNAT3 exon0.960.000.000.000.000.04
PHI0.970.020.000.020.000.00
Regulatory0.830.050.020.010.000.09
Table 4.

Allele frequency quantiles.

Frequency quantile
Site category<0.01<0.02<0.03<0.04<0.05<0.05
Noncoding0.830.040.020.020.010.09
CD36 exon0.960.010.000.010.000.02
GNAT3 exon0.960.000.000.000.000.04
PHI0.970.020.000.020.000.00
Regulatory0.830.050.020.010.000.09
Frequency quantile
Site category<0.01<0.02<0.03<0.04<0.05<0.05
Noncoding0.830.040.020.020.010.09
CD36 exon0.960.010.000.010.000.02
GNAT3 exon0.960.000.000.000.000.04
PHI0.970.020.000.020.000.00
Regulatory0.830.050.020.010.000.09

Consistent with the abundance of low-frequency alleles, π across noncoding sites was low, 0.10% (Table 3). It had a PE of 0.64 in the sliding window analysis, indicating that is consistent with expectations given the 1000GP sample. Values observed within superpopulations were similar to π values across the sample as a whole, ranging from 0.08 in Europe to 0.12 in Africa. Little difference in diversity was observed with respect to CD36 and GNAT3 exons within superpopulations, with π in CD36 ranging from 0.04% (in Europe) to 0.05% (in Africa) and π in GNAT3 ranging from 0.03% (in East Asians) to 0.04% (in Africans). Differences in diversity between introns and exons at both GNAT3 and CD36 were also small, with CD36 having π values of 0.08% and 0.05% in introns and exons, respectively, and GNAT3 having values of 0.07% and 0.04%.

The frequencies of PHI alleles ranged from 0.0002 to 0.031 with a mean of 0.0013. Thus, while the minimum frequency of PHI alleles was identical to that across noncoding sites, the mean was substantially lower (0.0013 vs. 0.0333). Nucleotide diversity could not be calculated for PHI sites or regulatory variants because the denominator in π calculations, the length of the analyzed sequence, is indeterminate because not all sites in the genome have potential to harbor PHI or regulatory variants. However, heterozygosity estimates were consistent with the distributions of allele frequencies at both noncoding and PHI sites, with 80% of noncoding and 97% of PHI sites having heterozygosities below 0.025. As with π, this pattern extended to superpopulations. Alleles scored as modifiers in regulatory regions ranged in frequency from 0.0002 to 0.970 with a mean of 0.037, and 83% having frequencies below 1%.

The overall FST of CD36-GNAT3 among superpopulations was 9.0% with a PE of 0.57 (Table 3). Pairwise FST values ranged from a low of 1.1% (between Europe and the Americas) to a high of 13.2% (between Africa and Europe). FST values among populations within superpopulations were smaller, ranging from 0.003 in Europe to 0.018 in the Americas. The FST among superpopulations was lower for both PHI sites (6.6%) and regulatory sites (8.0%) than for noncoding sites (9.0%; Table 3). This pattern held for pairwise FSTs between superpopulations, in which FST for noncoding variants was always higher than FST for PHI and regulatory sites. However, FST within superpopulations did not follow this pattern (Table 3). FST for PHI sites was similar to or less than FST for noncoding sites in four populations (Americas, East Asia, Europe, and South Asia). In contrast, FST in Africa was substantially higher for CD36 exons than for noncoding sites (7.9% vs. 1.5%) and higher still for PHI sites, 10.8%.

values across CD36-GNAT were high (0.7–1.0) across localized regions separated by regions of lower LD (<0.5), pointing to the presence of 7 haplotype blocks ranging from ~10 kb to ~100 kb in length (Fig. 3). A ~70 kb block was centered on GNAT3 and spanned its full length (53 kb). CD36 exons were distributed across 5 of the 7 blocks, which contained exons 1, 2–3, 4, 5–8, and 9–17, respectively. In contrast to , r2 was low across CD36-GNAT3 with the exception of highly localized areas. CD36 exons 10–17 were located in a block with r2 near 1.0. LD calculated among all PHI sites was consistent with patterns expected when allele frequencies are low. In general, pairwise LD measures between rare variants take on extreme values, with values near 1 and r2 values near 0. This pattern held across all but one pair of PHI sites. No pair of PHI sites had an r2 above 0.015 or below 0.995 with the exception of rs563097847 and rs558115067, which had an r2 of 0.17.

Linkage disequilibrium across CD36-GNAT3. (A) Pairwise Dʹ. (B) Pairwise r2.
Fig. 3.

Linkage disequilibrium across CD36-GNAT3. (A) Pairwise . (B) Pairwise r2.

Tajima’s D values were strongly negative for the primary site categories (all sites, exons, and introns; Table 5). However, their statistical significance depended on the test used. While the standard D test rejected neutrality at a high level of significance (P < 0.001), comparisons with the sliding window distribution yielded different results. In the sliding window analyses, only the D value of CD36 exons departed from expectations, and marginally so with PE = 0.035. Across the other four categories, PE ranged from 0.390 to 0.665, well within expectations (Table 5).

Table 5.

Neutrality Tests. Tests rejecting the null hypothesis with P < 0.001 are indicated by ∗∗∗. Tests rejecting the null hypothesis with P < 0.05 are indicated by ∗.

RegionTajima’s DPPE
All sites−1.87<0.001∗∗∗0.608
CD36 exons−2.41<0.001∗∗∗0.035∗
CD36 introns−1.86<0.001∗∗∗0.620
GNAT3 exons−1.81<0.001∗∗∗0.665
GNAT3 introns−2.05<0.001∗∗∗0.390
RegionTajima’s DPPE
All sites−1.87<0.001∗∗∗0.608
CD36 exons−2.41<0.001∗∗∗0.035∗
CD36 introns−1.86<0.001∗∗∗0.620
GNAT3 exons−1.81<0.001∗∗∗0.665
GNAT3 introns−2.05<0.001∗∗∗0.390
Table 5.

Neutrality Tests. Tests rejecting the null hypothesis with P < 0.001 are indicated by ∗∗∗. Tests rejecting the null hypothesis with P < 0.05 are indicated by ∗.

RegionTajima’s DPPE
All sites−1.87<0.001∗∗∗0.608
CD36 exons−2.41<0.001∗∗∗0.035∗
CD36 introns−1.86<0.001∗∗∗0.620
GNAT3 exons−1.81<0.001∗∗∗0.665
GNAT3 introns−2.05<0.001∗∗∗0.390
RegionTajima’s DPPE
All sites−1.87<0.001∗∗∗0.608
CD36 exons−2.41<0.001∗∗∗0.035∗
CD36 introns−1.86<0.001∗∗∗0.620
GNAT3 exons−1.81<0.001∗∗∗0.665
GNAT3 introns−2.05<0.001∗∗∗0.390

Discussion

Contemporary human populations are the product of a rapid expansion of small ancestral populations out of Africa 50–60 thousand years ago (Bergström et al. 2021). This process is evident in diversity patterns in human genes, which harbor signatures of ancient demography and natural selection (Bamshad and Wooding 2003; Marth et al. 2003). Most obvious on a genome-wide scale are a paucity of variation and downward skew in allele frequencies, patterns attributed to a combination of early population bottlenecks and pervasive pressure from purifying selection (Rogers 1995; Cvijović et al. 2018). However, while prevalent, these processes can act simultaneously with others, and patterns of diversity in individual genes can reveal the evolutionary underpinnings of specific traits (Hancock and Di Rienzo 2008). For instance, unexpectedly high LD and FST in the lactase (LCT) gene indicate that positive selection favored variants conferring lactase persistence in early herding peoples, where ability to digest milk was a fitness advantage (Bersaglieri et al. 2004). Conversely, high diversity and low LD in calpain-10 (CAPN10) signal the presence of long-term balancing selection at a locus implicated in energy use and storage, with the explanation being that more than one allele has been selectively maintained for an extended period and LD has decayed (Vander Molen et al. 2005). The overlapping structures of CD36 and GNAT3 together with their roles in taste and metabolism raise questions about the extent of diversity at these loci, its potential impact on phenotypes, and the roles of demography and natural selection in shaping it.

Worldwide diversity across CD36-GNAT3

We found that patterns of nucleotide diversity across CD36-GNAT3 were consistent with prevailing genome-wide trends and inferences about human origins. Diversity in noncoding regions, π = 0.10%, was well within expectations given the empirical distribution (PE = 0.64) and similar to numerous prior estimates, which vary depending on the populations sampled but are typically 0.075%–0.125% (Table 3; The 1000 Genomes Project Consortium 2010). Such values fall far below theoretical expectations given humans’ current population size. They are more consistent with paleoanthropological and genetic evidence that ancient populations sizes were small and attained their current size relatively recently (Henn et al. 2012; Osada 2015; Bergström et al. 2021). Nucleotide diversity within continents was similarly consistent with contemporary findings on human evolution. In particular, it was somewhat higher in Africa (π = 0.12%) than in other superpopulations (π = 0.09%). This pattern is widely observed and usually ascribed to the antiquity of African populations, which have diverged over a longer period than populations founded during humans’ recent expansion (Yu et al. 2002; Campbell and Tishkoff 2008; Henn et al. 2012; Bergström et al. 2021).

Patterns of population differentiation with respect to noncoding variation in our sample were also consistent with genome-wide trends and human origins. FST among superpopulations with respect to noncoding sites, 9%, was typical for values in the 1000GP (PE = 0.57) and similar to previous estimates among continents, which are typically near 10% (Table 3; Holsinger and Weir 2009). For instance, it is only slightly higher than the mean observed in HapMap phase 3 data (8.0%), which have a similar population composition (Elhaik 2012). Pairwise FST values between superpopulations were also consistent with previous estimates, varying widely depending on the continents being compared (1.1%–13.2%; Elhaik 2012). The mean within-superpopulation FST in our sample, 1.2%, was similarly consistent with previous genome-wide estimates, such as the 1% reported in HapMap phase 3 populations by Elhaik (2012). As with π, these values reflect expectations following humans’ history of population growth and dispersal, with groups near each other having more recent common ancestry and higher rates of intermigration than populations farther apart.

The presence of nonsynonymous SNPs and PHI sites in CD36 and GNAT3 exons suggests that functional polymorphism is present in both genes, which could affect fat taste and metabolism as well as non-gustatory phenotypes. However, patterns of diversity at PHI sites indicated that their contributions to phenotypic variance on a population scale are limited. Of the 113 coding changes we found, only 60 were scored as PHI variants. Further, the frequencies of PHI alleles were low, with only two of the 60 having frequencies above 1% and the most common having a frequency of 3% (Table 4). These patterns extended to diversity within superpopulations, with the number of PHI sites being lower than the number of nonsynonymous variants on all continents (Table 3). These patterns indicate that while polymorphism affecting phenotypes is almost certainly present, it is low in frequency.

FSTs within superpopulations revealed potentially important regional trends. In particular, FSTs were far higher in Africa than elsewhere for two site categories (Table 3). The FST of CD36 in Africa was 7-fold greater than in any other superpopulation (0.079 vs. 0.011 in East Asia) and 5-fold greater than in noncoding sites in Africa (0.079 vs. 0.015). The FST of PHI sites in Africa was also 7-fold greater than in any other superpopulation (0.108 vs. 0.015 in East Asia) and 5-fold greater than in noncoding sites. These patterns suggest that populations may be differentiated with respect to functional variation and, if so, it is likely most pronounced with respect to CD36 in Africa. However, it is important to note that while patterns in FST with respect to PHI sites point to the presence of phenotypic differences among populations they provide little information about the magnitude of the differences, which depend not just on allele frequencies, but on effect sizes as well.

Signatures of natural selection

The high-calorie content of fats makes them highly valuable nutritionally, so the mechanisms underlying their detection must be under strong selective pressure. This is a familiar issue in human evolutionary biology, most famously as part of Neel’s “thrifty gene” hypothesis, which posits that ability to store fat provides fitness advantages (Neel 1962; Reales et al. 2017). Under the thrifty gene hypothesis, the advantages of perceiving and metabolizing fats are evident. However, its implications specifically for CD36 and GNAT3 are not. On the one hand, the roles of CD36 and GNAT3 in sensory signaling, lipid metabolism, and other processes imply that they are intolerant of novel mutations and under pressure from purifying natural selection. However, humans’ diffusion out of Africa placed CD36 and GNAT3 into novel environments, which could have produced pressures to adapt. Moreover, a single genomic region can be under more than one selective pressure at a time. For instance, selective pressures on the LCT gene appear to have been strong in some parts of Europe but not others (Bersaglieri et al. 2004). Similarly, selective pressures on TAS2R38 have been stronger on some haplotype backgrounds than others (Risso et al. 2016).

The results of our Tajima’s D tests across CD36-GNAT3 as a whole were most consistent with an absence of selective pressure during its recent history. As expected given the low diversity and downward skew in allele frequencies, observed D values were strongly negative (Table 5). This can result from two common selective pressures: purifying selection and positive selection. However, it can also be caused by rapid population growth, during which genetic drift slows and new variants accumulate (Wooding 2003; Adams and Hudson 2004). In our study, the observed D was below expectations at a high level of statistical significance (P < 0.001). However, this did not reveal whether the cause of the shift was selection, demography, or both. D tests using the empirical distribution in the 1000GP clarified the probable cause. Because the vast majority of the human genome (>98%) is noncoding it is can be assumed to be evolving neutrally or nearly so with respect to selection, but it is still shaped by demography. Thus, the empirical distribution of D we obtained from the 1000GP represents expectations adjusted for demography while the standard distribution does not. We found that when compared with the empirical distribution, D values were well within expectations for four of the five site categories we analyzed (0.390 < PE < 0.665; Table 5). The exception was in CD36 exons, where the observed D was significantly lower than expected, albeit marginally so (D = −2.41, PE = 0.035). This is consistent with localized purifying or weak positive selection. In our view, the empirical tests are the more convincing and conclude that on a worldwide scale selective pressures have been largely absent across CD36-GNAT3 during its recent history with exception of exonic regions in CD36.

Evidence for selection in CD36 exons in the 1000GP is consistent with prior findings by Fry et al. (2009). In a study of CD36 in Africa, Fry et al. detected signatures of positive selection favoring a premature stop allele at rs3211938, which is located in CD36 exon 13. Fry et al.’s investigation centered on resistance to malaria infection, which is hypothesized to be affected by variation in CD36 because the receptor is an antigenic target for Plasmodium falciparum. The study found signatures of selection surrounding rs3211938 but rejected the malaria resistance hypothesis because association studies failed to detect correlations with malaria severity. However, the selective signatures were robust; suggesting that factors other than malaria resistance must be responsible. Evidence that lipid perception and metabolism are mediated by CD36 offers a possible explanation. In clinical studies, variation in CD36 associates with lipid perception and metabolic phenotypes, and rs3211938 is specifically implicated as the source of the associations (Love-Gregory et al. 2008, 2011). This supports the hypothesis that CD36 harbors genotypes affecting responses to fats, which could exert selective pressures on genes mediating fat taste, metabolism, or both. Thus, a speculative explanation for the high frequency of the rs3211938 stop allele in localized areas of Africa is that it was favored by the nutritional environment, and contemporary associations are a holdover from those pressures.

Our findings on population differentiation are also consistent with Fry et al.’s (2009) proposal that selection has promoted differentiation among African populations. In our data, FST with respect to noncoding sites in Africa (FST = 1.5%) was similar to or slightly greater than in other superpopulations (FST = 0.2%–1.8%), indicating they are differentiated to roughly the same extent with respect to neutral variation (Table 3). However, FST with respect to CD36 exons was far higher in Africa than in other superpopulations (7.9% vs. <1.5%), suggesting that some factor has driven differentiation specifically at that locus. Moreover, the trend was amplified in PHI sites, which is expected under local adaptation because selection most affects sites with functional effects. Further, differentiation in Africa was greatest specifically with respect to the premature stop in CD36 (rs3211938), a major mutation particularly likely to experience selective effects. These patterns are consistent with the effects of local adaptation and that evolutionary pressures on fat responses had important effects on diversity in the continent.

Implications for association studies

Our finding that CD36-GNAT3 harbors numerous sites with putative functional effects supports predictions that the region bears variants affecting lipid perception and metabolism. To date, more than 80 variants in CD36 and GNAT3 have been reported to associate with both gustatory and non-gustatory phenotypes, and the potential for connections of CD36 with obesity and related diseases draws ongoing interest (MacArthur et al. 2017). Our diversity estimates and functional predictions support these findings. They specifically buttress evidence for associations at eight sites identified in six previous studies (Table 6). Cross tabulating our list of PHI and regulatory sites against the list of previously reported associations revealed six with putative effects: rs1527479, rs1534314, rs3211883, rs3211938, rs12706912, and rs6960369. And of these, three (rs1527479, rs3211883, and rs3211938) showed evidence of associations in more than one study. One (rs3211938, the premature stop in CD36) showed evidence of being under pressure from positive natural selection, which can only act on functional sites. Thus, these are particularly strong candidates for future investigation.

Table 6.

PHI and regulatory sites with reported genotype–phenotype associations.

Allele frequency
rsidEffectWorldwideAfricaAmericasEast AsiaEuropeSouth AsiaReports
rs1527479Expression modifier0.350.220.460.310.540.30Bokor et al. 2010
Jayewardene et al. 2016
Lecompte et al. 2011
rs1534314Expression modifier0.260.260.190.440.070.30Ghosh et al. 2011
rs3211883Expression modifier0.640.340.760.620.900.69Bokor et al. 2010
Ghosh et al. 2011
Heni et al. 2011
rs3211938Stop gained (CD36 Exon 13)0.030.120.000.000.000.00Fry et al. 2009
Love-Gregory et al. 2011
Love-Gregory et al. 2008
rs6960369Expression modifier0.230.330.130.340.070.28Ghosh et al. 2011
rs12706912Expression modifier0.560.600.450.590.440.68Ghosh et al. 2011
Allele frequency
rsidEffectWorldwideAfricaAmericasEast AsiaEuropeSouth AsiaReports
rs1527479Expression modifier0.350.220.460.310.540.30Bokor et al. 2010
Jayewardene et al. 2016
Lecompte et al. 2011
rs1534314Expression modifier0.260.260.190.440.070.30Ghosh et al. 2011
rs3211883Expression modifier0.640.340.760.620.900.69Bokor et al. 2010
Ghosh et al. 2011
Heni et al. 2011
rs3211938Stop gained (CD36 Exon 13)0.030.120.000.000.000.00Fry et al. 2009
Love-Gregory et al. 2011
Love-Gregory et al. 2008
rs6960369Expression modifier0.230.330.130.340.070.28Ghosh et al. 2011
rs12706912Expression modifier0.560.600.450.590.440.68Ghosh et al. 2011
Table 6.

PHI and regulatory sites with reported genotype–phenotype associations.

Allele frequency
rsidEffectWorldwideAfricaAmericasEast AsiaEuropeSouth AsiaReports
rs1527479Expression modifier0.350.220.460.310.540.30Bokor et al. 2010
Jayewardene et al. 2016
Lecompte et al. 2011
rs1534314Expression modifier0.260.260.190.440.070.30Ghosh et al. 2011
rs3211883Expression modifier0.640.340.760.620.900.69Bokor et al. 2010
Ghosh et al. 2011
Heni et al. 2011
rs3211938Stop gained (CD36 Exon 13)0.030.120.000.000.000.00Fry et al. 2009
Love-Gregory et al. 2011
Love-Gregory et al. 2008
rs6960369Expression modifier0.230.330.130.340.070.28Ghosh et al. 2011
rs12706912Expression modifier0.560.600.450.590.440.68Ghosh et al. 2011
Allele frequency
rsidEffectWorldwideAfricaAmericasEast AsiaEuropeSouth AsiaReports
rs1527479Expression modifier0.350.220.460.310.540.30Bokor et al. 2010
Jayewardene et al. 2016
Lecompte et al. 2011
rs1534314Expression modifier0.260.260.190.440.070.30Ghosh et al. 2011
rs3211883Expression modifier0.640.340.760.620.900.69Bokor et al. 2010
Ghosh et al. 2011
Heni et al. 2011
rs3211938Stop gained (CD36 Exon 13)0.030.120.000.000.000.00Fry et al. 2009
Love-Gregory et al. 2011
Love-Gregory et al. 2008
rs6960369Expression modifier0.230.330.130.340.070.28Ghosh et al. 2011
rs12706912Expression modifier0.560.600.450.590.440.68Ghosh et al. 2011

Statistical power is a critical consideration in association studies aimed at dissecting phenotypic effects, and our findings predict that it will be weak in CD36-GNAT3. Although numerous sites predicted to affect protein function and gene expression were present in the region, their frequencies were low. For instance, 83% of regulatory alleles had frequencies below 1%, and 97% of PHI alleles had frequencies below 1%. This is important because even when such alleles have functional impact they cannot contribute much phenotypic variance overall. Conversely, it suggests that if CD36-GNAT3 does harbor alleles contributing substantial variance, they must be limited to a small number of sites where allele frequencies are elevated. This pattern is evident in the cross-tabulation of PHI and regulatory sites against reported associations, in which all alleles exhibiting associations were found at high frequencies (>10%) in at least one superpopulation (Table 6).

Patterns of LD in our sample also have implications for association studies. When high, LD can produce false associations where noncausal variants cosegregate with causal ones, making their effects difficult to discriminate. However, it can enhance efforts to dissect associations by reducing the density of genotyping needed to localize the sources of effects, which can subsequently be pinpointed using fine mapping strategies. In contrast, when LD is low, sites’ independent effects on phenotypes can be pinpointed directly through dense genotyping; however, this approach is vulnerable to missing effects if marker density is insufficient. In the case of CD36 and GNAT3, the structure of LD is a particularly important consideration because the two genes underlie similar chemosensory traits. Thus, if present, high LD could cause GNAT3- and CD36-mediated phenotypes to spuriously associate with variants in both genes, making their functional contributions difficult to distinguish. We saw low potential for such effects in CD36-GNAT3. While the high values we observed are consistent with a risk of confounds, r2 was low or 0 between almost all site pairs, including PHI and regulatory sites. This makes confounds unlikely.

Conflict of Interest

None declared.

References

Adams
AM
,
Hudson
RR.
2004
.
Maximum-likelihood estimation of demographic parameters using the frequency spectrum of unlinked single-nucleotide polymorphisms.
Genetics
.
168
(
3
):
1699
1712
.

Adzhubei
IA
,
Schmidt
S
,
Peshkin
L
,
Ramensky
VE
,
Gerasimova
A
,
Bork
P
,
Kondrashov
AS
,
Sunyaev
SR.
2010
.
A method and server for predicting damaging missense mutations.
Nat Methods
.
7
(
4
):
248
249
.

Allen
AL
,
McGeary
JE
,
Hayes
JE.
2014
.
Polymorphisms in TRPV1 and TAS2Rs associate with sensations from sampled ethanol.
Alcohol Clin Exp Res
.
38
(
10
):
2550
2560
.

Antinucci
M
,
Risso
D.
2017
.
A matter of taste: lineage-specific loss of function of taste receptor genes in vertebrates.
Front Mol Biosci
.
4
:
81
. doi:10.3389/fmolb.2017.00081.

Bachmanov
AA
,
Bosak
NP
,
Lin
C
,
Matsumoto
I
,
Ohmoto
M
,
Reed
DR
,
Nelson
TM.
2014
.
Genetics of taste receptors.
Curr Pharm Des
.
20
(
16
):
2669
2683
.

Baldwin
MW
,
Toda
Y
,
Nakagita
T
,
O’Connell
MJ
,
Klasing
KC
,
Misaka
T
,
Edwards
SV
,
Liberles
SD.
2014
.
Sensory biology. Evolution of sweet taste perception in hummingbirds by transformation of the ancestral umami receptor.
Science
.
345
(
6199
):
929
933
.

Bamshad
M
,
Wooding
SP.
2003
.
Signatures of natural selection in the human genome.
Nat Rev Genet
.
4
(
2
):
99
111
.

Barontini
J
,
Antinucci
M
,
Tofanelli
S
,
Cammalleri
M
,
Dal Monte
M
,
Gemignani
F
,
Vodicka
P
,
Marangoni
R
,
Vodickova
L
,
Kupcinskas
J
, et al.
2017
.
Association between polymorphisms of TAS2R16 and susceptibility to colorectal cancer.
BMC Gastroenterol
.
17
(
1
):
104
.

Basson
MD
,
Bartoshuk
LM
,
Dichello
SZ
,
Panzini
L
,
Weiffenbach
JM
,
Duffy
VB.
2005
.
Association between 6-n-propylthiouracil (PROP) bitterness and colonic neoplasms.
Dig Dis Sci
.
50
(
3
):
483
489
.

Behrens
M
,
Di Pizio
A
,
Redel
U
,
Meyerhof
W
,
Korsching
SI.
2020
.
At the root of T2R gene evolution: recognition profiles of coelacanth and zebrafish bitter receptors.
Genome Biol Evol
.
13
(
1
). doi:10.1093/gbe/evaa264.

Behrens
M
,
Gunn
HC
,
Ramos
PC
,
Meyerhof
W
,
Wooding
SP.
2013
.
Genetic, functional, and phenotypic diversity in TAS2R38-mediated bitter taste perception.
Chem Senses
.
38
(
6
):
475
484
.

Bergström
A
,
Stringer
C
,
Hajdinjak
M
,
Scerri
EML
,
Skoglund
P.
2021
.
Origins of modern human ancestry.
Nature
.
590
(
7845
):
229
237
.

Bersaglieri
T
,
Sabeti
PC
,
Patterson
N
,
Vanderploeg
T
,
Schaffner
SF
,
Drake
JA
,
Rhodes
M
,
Reich
DE
,
Hirschhorn
JN.
2004
.
Genetic signatures of strong recent positive selection at the lactase gene.
Am J Hum Genet
.
74
(
6
):
1111
1120
.

Bokor
S
,
Legry
V
,
Meirhaeghe
A
,
Ruiz
JR
,
Mauro
B
,
Widhalm
K
,
Manios
Y
,
Amouyel
P
,
Moreno
LA
,
Molnar
D
, et al.
2010
.
Single-nucleotide polymorphism of CD36 locus and obesity in European adolescents.
Obesity (Silver Spring)
.
18
:
1398
1403
.

Campbell
MC
,
Ranciaro
A
,
Froment
A
,
Hirbo
J
,
Omar
S
,
Bodo
JM
,
Nyambo
T
,
Lema
G
,
Zinshteyn
D
,
Drayna
D
, et al.
2012
.
Evolution of functionally diverse alleles associated with PTC bitter taste sensitivity in Africa.
Mol Biol Evol
.
29
(
4
):
1141
1153
.

Campbell
MC
,
Ranciaro
A
,
Zinshteyn
D
,
Rawlings-Goss
R
,
Hirbo
J
,
Thompson
S
,
Woldemeskel
D
,
Froment
A
,
Rucker
JB
,
Omar
SA
, et al.
2014
.
Origin and differential selection of allelic variation at TAS2R16 associated with salicin bitter taste sensitivity in Africa.
Mol Biol Evol
.
31
(
2
):
288
302
.

Campbell
MC
,
Tishkoff
SA.
2008
.
African genetic diversity: implications for human demographic history, modern human origins, and complex disease mapping.
Annu Rev Genomics Hum Genet
.
9
:
403
433
. doi:10.1146/annurev.genom.9.081307.164258.

Cartoni
C
,
Yasumatsu
K
,
Ohkuri
T
,
Shigemura
N
,
Yoshida
R
,
Godinot
N
,
le Coutre
J
,
Ninomiya
Y
,
Damak
S.
2010
.
Taste preference for fatty acids is mediated by GPR40 and GPR120.
J Neurosci
.
30
(
25
):
8376
8382
.

Chalé-Rush
A
,
Burgess
JR
,
Mattes
RD.
2007
.
Evidence for human orosensory (taste?) sensitivity to free fatty acids.
Chem Senses
.
32
(
5
):
423
431
.

Chen
QY
,
Alarcon
S
,
Tharp
A
,
Ahmed
OM
,
Estrella
NL
,
Greene
TA
,
Rucker
J
,
Breslin
PA.
2009
.
Perceptual variation in umami taste and polymorphisms in TAS1R taste receptor genes.
Am J Clin Nutr
.
90
(
3
):
770S
779S
.

Choi
JH
,
Lee
J
,
Choi
IJ
,
Kim
YW
,
Ryu
KW
,
Kim
J.
2016
.
Genetic variation in the TAS2R38 bitter taste receptor and gastric cancer risk in Koreans.
Sci Rep
.
6
:
26904
. doi:10.1038/srep26904.

Cvijović
I
,
Good
BH
,
Desai
MM.
2018
.
The effect of strong purifying selection on genetic diversity.
Genetics
.
209
(
4
):
1235
1278
.

Danecek
P
,
Auton
A
,
Abecasis
G
,
Albers
CA
,
Banks
E
,
DePristo
MA
,
Handsaker
RE
,
Lunter
G
,
Marth
GT
,
Sherry
ST
, et al. ;
1000 Genomes Project Analysis Group
.
2011
.
The variant call format and VCFtools.
Bioinformatics
.
27
(
15
):
2156
2158
.

De Araujo
IE
,
Rolls
ET.
2004
.
Representation in the human brain of food texture and oral fat.
J Neurosci
.
24
(
12
):
3086
3093
.

Deshpande
DA
,
Wang
WC
,
McIlmoyle
EL
,
Robinett
KS
,
Schillinger
RM
,
An
SS
,
Sham
JS
,
Liggett
SB.
2010
.
Bitter taste receptors on airway smooth muscle bronchodilate by localized calcium signaling and reverse obstruction.
Nat Med
.
16
(
11
):
1299
1304
.

Drayna
D.
2005
.
Human taste genetics.
Annu Rev Genomics Hum Genet
.
6
:
217
235
. doi:10.1146/annurev.genom.6.080604.162340.

Egan
JM
,
Margolskee
RF.
2008
.
Taste cells of the gut and gastrointestinal chemosensation.
Mol Interv
.
8
(
2
):
78
81
.

Elhaik
E.
2012
.
Empirical distributions of FST from large-scale human polymorphism data.
PLoS One
.
7
(
11
):
e49837
.

Farook
VS
,
Puppala
S
,
Schneider
J
,
Fowler
SP
,
Chittoor
G
,
Dyer
TD
,
Allayee
H
,
Cole
SA
,
Arya
R
,
Black
MH
, et al.
2012
.
Metabolic syndrome is linked to chromosome 7q21 and associated with genetic variants in CD36 and GNAT3 in Mexican Americans.
Obesity (Silver Spring)
.
20
(
10
):
2083
2092
.

Feeney
EL
,
Hayes
JE.
2014
.
Exploring associations between taste perception, oral anatomy and polymorphisms in the carbonic anhydrase (gustin) gene CA6.
Physiol Behav
.
128
:
148
154
. doi:10.1016/j.physbeh.2014.02.013.

Feng
P
,
Zheng
J
,
Rossiter
SJ
,
Wang
D
,
Zhao
H.
2014
.
Massive losses of taste receptor genes in toothed and baleen whales.
Genome Biol Evol
.
6
(
6
):
1254
1265
.

Fischer
A
,
Gilad
Y
,
Man
O
,
Pääbo
S.
2005
.
Evolution of bitter taste receptors in humans and apes.
Mol Biol Evol
.
22
(
3
):
432
436
.

Fry
AE
,
Ghansa
A
,
Small
KS
,
Palma
A
,
Auburn
S
,
Diakite
M
,
Green
A
,
Campino
S
,
Teo
YY
,
Clark
TG
, et al.
2009
.
Positive selection of a CD36 nonsense variant in sub-Saharan Africa, but no association with severe malaria phenotypes.
Hum Mol Genet
.
18
(
14
):
2683
2692
.

Fushan
AA
,
Simons
CT
,
Slack
JP
,
Drayna
D.
2010
.
Association between common variation in genes encoding sweet taste signaling components and human sucrose perception.
Chem Senses
.
35
(
7
):
579
592
.

Gaillard
D
,
Laugerette
F
,
Darcel
N
,
El-Yassimi
A
,
Passilly-Degrace
P
,
Hichami
A
,
Khan
NA
,
Montmayeur
JP
,
Besnard
P.
2008
.
The gustatory pathway is involved in CD36-mediated orosensory perception of long-chain fatty acids in the mouse.
FASEB J
.
22
(
5
):
1458
1468
.

Gentleman
RC
,
Carey
VJ
,
Bates
DM
,
Bolstad
B
,
Dettling
M
,
Dudoit
S
,
Ellis
B
,
Gautier
L
,
Ge
Y
,
Gentry
J
, et al.
2004
.
Bioconductor: open software development for computational biology and bioinformatics.
Genome Biol
.
5
(
10
):
R80
.

Ghosh
A
,
Murugesan
G
,
Chen
K
,
Zhang
L
,
Wang
Q
,
Febbraio
M
,
Anselmo
RM
,
Marchant
K
,
Barnard
J
,
Silverstein
RL.
2011
.
Platelet CD36 surface expression levels affect functional responses to oxidized LDL and are associated with inheritance of specific genetic polymorphisms.
Blood
.
117
:
6355
6366
.

Goudet
J.
2004
.
hierfstat, a package for R to compute and test hierarchical F-statistics.
Mol Ecol Notes
.
5
:
184
186
. doi:10.18637/jss.v016.c0.

Graham
J
,
McNeney
B
,
Blay
S
,
Shin
J-H.
2006
.
LDheatmap: an R function for graphical display of pairwise linkage disequilibria between single nucleotide polymorphisms.
J Stat Softw
.
16
:
c03
. doi:10.18637/jss.v016.c03.

Greene
LS.
1974
.
Physical growth and development, neurological maturation, and behavioral functioning in two Ecuadorian Andean communities in which goiter is endemic. II. PTC taste sensitivity and neurological maturation.
Am J Phys Anthropol
.
41
(
1
):
139
151
.

Hancock
AM
,
Di Rienzo
A.
2008
.
Detecting the genetic signature of natural selection in human populations: models, methods, and data.
Annu Rev Anthropol
.
37
:
197
217
. doi:10.1146/annurev.anthro.37.081407.085141.

Hayes
JE
,
Feeney
EL
,
Nolden
AA
,
McGeary
JE.
2015
.
Quinine bitterness and grapefruit liking associate with allelic variants in TAS2R31.
Chem Senses
.
40
(
6
):
437
443
.

Heni
M
,
Mussig
K
,
Machicao
F
,
Machann
J
,
Schick
F
,
Claussen
CD
,
Stefan
N
,
Fritsche
A
,
Haring
HU
,
Staiger
H.
2011
.
Variants in the CD36 gene locus determine whole-body adiposity, but have no independent effect on insulin sensitivity.
Obesity (Silver Spring)
.
19
:
1004
1009
.

Henn
BM
,
Cavalli-Sforza
LL
,
Feldman
MW.
2012
.
The great human expansion.
Proc Natl Acad Sci USA
.
109
(
44
):
17758
17764
.

Holsinger
KE
,
Weir
BS.
2009
.
Genetics in geographically structured populations: defining, estimating and interpreting FST.
Nat Rev Genet
.
10
(9):
639
650
.

Ihaka
R
,
Gentleman
R.
1996
.
R: a language for data analysis and graphics.
J Comput Graph Stat
.
5
(3):
299
314
.

Jayewardene
AF
,
Mavros
Y
,
Gwinn
T
,
Hancock
DP
,
Rooney
KB.
2016
.
Associations between CD36 gene polymorphisms and metabolic response to a short-term endurance-training program in a young-adult population.
Appl Physiol Nutr Metab
.
41
:
157
167
.

Jiang
P
,
Josue
J
,
Li
X
,
Glaser
D
,
Li
W
,
Brand
JG
,
Margolskee
RF
,
Reed
DR
,
Beauchamp
GK.
2012
.
Major taste loss in carnivorous mammals.
Proc Natl Acad Sci USA
.
109
(
13
):
4956
4961
.

Jombart
T
,
Ahmed
I.
2011
.
adegenet 1.3-1: new tools for the analysis of genome-wide SNP data.
Bioinformatics
.
27
(
21
):
3070
3071
.

Keller
KL
,
Liang
LC
,
Sakimura
J
,
May
D
,
van Belle
C
,
Breen
C
,
Driggin
E
,
Tepper
BJ
,
Lanzano
PC
,
Deng
L
, et al.
2012
.
Common variants in the CD36 gene are associated with oral fat perception, fat preferences, and obesity in African Americans.
Obesity (Silver Spring)
.
20
(
5
):
1066
1073
.

Kim
UK
,
Breslin
PA
,
Reed
D
,
Drayna
D.
2004
.
Genetics of human taste perception.
J Dent Res
.
83
(
6
):
448
453
.

Kim
UK
,
Wooding
S
,
Riaz
N
,
Jorde
LB
,
Drayna
D.
2006
.
Variation in the human TAS1R taste receptor genes.
Chem Senses
.
31
(
7
):
599
611
.

Kim
U
,
Wooding
S
,
Ricci
D
,
Jorde
LB
,
Drayna
D.
2005
.
Worldwide haplotype diversity and coding sequence variation at human bitter taste receptor loci.
Hum Mutat
.
26
(
3
):
199
204
.

Kumar
P
,
Henikoff
S
,
Ng
PC.
2009
.
Predicting the effects of coding non-synonymous variants on protein function using the SIFT algorithm.
Nat Protoc
.
4
(
7
):
1073
1081
.

Laugerette
F
,
Passilly-Degrace
P
,
Patris
B
,
Niot
I
,
Febbraio
M
,
Montmayeur
JP
,
Besnard
P.
2005
.
CD36 involvement in orosensory detection of dietary lipids, spontaneous fat preference, and digestive secretions.
J Clin Invest
.
115
(
11
):
3177
3184
.

Lecompte
S
,
Szabo de Edelenyi
F
,
Goumidi
L
,
Maiani
G
,
Moschonis
G
,
Widhalm
K
,
Molnar
D
,
Kafatos
A
,
Spinneker
A
,
Breidenassel
C
, et al.
2011
.
Polymorphisms in the CD36/FAT gene are associated with plasma vitamin E concentrations in humans.
Am J Clin Nutr
.
93
:
644
651
.

Li
H.
2011
.
Tabix: fast retrieval of sequence features from generic TAB-delimited files.
Bioinformatics
.
27
(
5
):
718
719
.

Li
D
,
Zhang
J.
2014
.
Diet shapes the evolution of the vertebrate bitter taste receptor gene repertoire.
Mol Biol Evol
.
31
(
2
):
303
309
.

Lindemann
B.
2001
.
Receptors and transduction in taste.
Nature
.
413
(
6852
):
219
225
.

Love-Gregory
L
,
Sherva
R
,
Schappe
T
,
Qi
JS
,
McCrea
J
,
Klein
S
,
Connelly
MA
,
Abumrad
NA.
2011
.
Common CD36 SNPs reduce protein expression and may contribute to a protective atherogenic profile.
Hum Mol Genet
.
20
(
1
):
193
201
.

Love-Gregory
L
,
Sherva
R
,
Sun
L
,
Wasson
J
,
Schappe
T
,
Doria
A
,
Rao
DC
,
Hunt
SC
,
Klein
S
,
Neuman
RJ
, et al.
2008
.
Variants in the CD36 gene associate with the metabolic syndrome and high-density lipoprotein cholesterol.
Hum Mol Genet
.
17
(
11
):
1695
1704
.

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
(
D1
):
D896
D901
.

Marth
G
,
Schuler
G
,
Yeh
R
,
Davenport
R
,
Agarwala
R
,
Church
D
,
Wheelan
S
,
Baker
J
,
Ward
M
,
Kholodov
M
, et al.
2003
.
Sequence variations in the public human genome data reflect a bottlenecked population history.
Proc Natl Acad Sci USA
.
100
(
1
):
376
381
.

Mattes
RD.
2011
.
Accumulating evidence supports a taste component for free fatty acids in humans.
Physiol Behav
.
104
(
4
):
624
631
.

McLaren
W
,
Gil
L
,
Hunt
SE
,
Riat
HS
,
Ritchie
GR
,
Thormann
A
,
Flicek
P
,
Cunningham
F.
2016
.
The Ensembl variant effect predictor.
Genome Biol
.
17
(
1
):
122
.

Mueller
JC.
2004
.
Linkage disequilibrium for different scales and applications.
Brief Bioinform
.
5
(
4
):
355
364
.

Neel
JV.
1962
.
Diabetes mellitus: a “thrifty” genotype rendered detrimental by “progress”?
Am J Hum Genet
.
14
(4):
353
362
.

Obenchain
V
,
Lawrence
M
,
Carey
V
,
Gogarten
S
,
Shannon
P
,
Morgan
M.
2014
.
VariantAnnotation: a Bioconductor package for exploration and annotation of genetic variants.
Bioinformatics
.
30
(
14
):
2076
2078
.

Osada
N.
2015
.
Genetic diversity in humans and non-human primates and its evolutionary consequences.
Genes Genet Syst
.
90
(
3
):
133
145
.

Paradis
E.
2010
.
pegas: an R package for population genetics with an integrated-modular approach.
Bioinformatics
.
26
(
3
):
419
420
.

Pepino
MY
,
Kuda
O
,
Samovski
D
,
Abumrad
NA.
2014
.
Structure-function of CD36 and importance of fatty acid signal transduction in fat metabolism.
Annu Rev Nutr
.
34
:
281
303
. doi:10.1146/annurev-nutr-071812-161220.

Pepino
MY
,
Love-Gregory
L
,
Klein
S
,
Abumrad
NA.
2012
.
The fatty acid translocase gene CD36 and lingual lipase influence oral sensitivity to fat in obese subjects.
J Lipid Res
.
53
(
3
):
561
566
.

Pfeifer
B
,
Wittelsbürger
U
,
Ramos-Onsins
SE
,
Lercher
MJ.
2014
.
PopGenome: an efficient Swiss army knife for population genomic analyses in R.
Mol Biol Evol
.
31
(
7
):
1929
1936
.

Reales
G
,
Rovaris
DL
,
Jacovas
VC
,
Hünemeier
T
,
Sandoval
JR
,
Salazar-Granara
A
,
Demarchi
DA
,
Tarazona-Santos
E
,
Felkl
AB
,
Serafini
MA
, et al.
2017
.
A tale of agriculturalists and hunter-gatherers: exploring the thrifty genotype hypothesis in native South Americans.
Am J Phys Anthropol
.
163
(
3
):
591
601
.

Reed
DR
,
Knaapila
A.
2010
.
Genetics of taste and smell: poisons and pleasures.
Prog Mol Biol Transl Sci
.
94
:
213
240
. doi:10.1016/B978-0-12-375003-7.00008-X.

Risso
DS
,
Mezzavilla
M
,
Pagani
L
,
Robino
A
,
Morini
G
,
Tofanelli
S
,
Carrai
M
,
Campa
D
,
Barale
R
,
Caradonna
F
, et al.
2016
.
Global diversity in the TAS2R38 bitter taste receptor: revisiting a classic evolutionary PROPosal.
Sci Rep
.
6
(10):
25506
.

Risso
D
,
Sainz
E
,
Morini
G
,
Tofanelli
S
,
Drayna
D.
2018
.
Taste perception of Antidesma bunius fruit and its relationships to bitter taste receptor gene haplotypes.
Chem Senses
.
43
(
7
):
463
468
.

Rogers
AR.
1995
.
Genetic evidence for a Pleistocene population explosion.
Evolution
.
49
(
4
):
608
615
.

Roper
SD
,
Chaudhari
N.
2017
.
Taste buds: cells, signals and synapses.
Nat Rev Neurosci
.
18
(
8
):
485
497
.

Silverstein
RL
,
Febbraio
M.
2009
.
CD36, a scavenger receptor involved in immunity, metabolism, angiogenesis, and behavior.
Sci Signal
.
2
(
72
):
re3
.

Slatkin
M.
2008
.
Linkage disequilibrium—understanding the evolutionary past and mapping the medical future.
Nat Rev Genet
.
9
(
6
):
477
485
.

Tajima
F.
1983
.
Evolutionary relationship of DNA sequences in finite populations.
Genetics
.
105
(
2
):
437
460
.

Tajima
F.
1989
.
Statistical method for testing the neutral mutation hypothesis by DNA polymorphism.
Genetics
.
123
(
3
):
585
595
.

The 1000 Genomes Project Consortium.
2010
.
A map of human genome variation from population-scale sequencing.
Nature
.
467
(
7319
):
1061
1073
.

The 1000 Genomes Project Consortium.
2015
.
A global reference for human genetic variation.
Nature
.
526
(
7571
):
68
74
.

Vander Molen
J
,
Frisse
LM
,
Fullerton
SM
,
Qian
Y
,
Del Bosque-Plata
L
,
Hudson
RR
,
Di Rienzo
A.
2005
.
Population genetics of CAPN10 and GPR35: implications for the evolution of type 2 diabetes variants.
Am J Hum Genet
.
76
(
4
):
548
560
.

Weir
BS
,
Hill
WG.
2002
.
Estimating F-statistics.
Annu Rev Genet
.
36
:
721
750
. doi:10.1146/annurev.genet.36.050802.093940.

Wooding
S.
2003
.
PopHist: inferring population history from the spectrum of allele frequencies.
Bioinformatics
.
19
(
4
):
539
540
. doi:10.1186/1471-2350-13-96.

Wooding
S.
2011
.
Signatures of natural selection in a primate bitter taste receptor.
J Mol Evol
.
73
(
5–6
):
257
265
.

Wooding
SP
,
Atanasova
S
,
Gunn
HC
,
Staneva
R
,
Dimova
I
,
Toncheva
D.
2012
.
Association of a bitter taste receptor mutation with Balkan Endemic Nephropathy (BEN).
BMC Med Genet
.
13
:
96
.

Wooding
S
,
Bufe
B
,
Grassi
C
,
Howard
MT
,
Stone
AC
,
Vazquez
M
,
Dunn
DM
,
Meyerhof
W
,
Weiss
RB
,
Bamshad
MJ.
2006
.
Independent evolution of bitter-taste sensitivity in humans and chimpanzees.
Nature
.
440
(
7086
):
930
934
.

Wooding
S
,
Gunn
H
,
Ramos
P
,
Thalmann
S
,
Xing
C
,
Meyerhof
W.
2010
.
Genetics and bitter taste responses to goitrin, a plant toxin found in vegetables.
Chem Senses
.
35
(
8
):
685
692
.

Wooding
S
,
Kim
UK
,
Bamshad
MJ
,
Larsen
J
,
Jorde
LB
,
Drayna
D.
2004
.
Natural selection and molecular evolution in PTC, a bitter-taste receptor gene.
Am J Hum Genet
.
74
(
4
):
637
646
.

Yu
N
,
Chen
FC
,
Ota
S
,
Jorde
LB
,
Pamilo
P
,
Patthy
L
,
Ramsay
M
,
Jenkins
T
,
Shyue
SK
,
Li
WH.
2002
.
Larger genetic differences within Africans than between Africans and Eurasians.
Genetics
.
161
(
1
):
269
274
.

Zhao
H
,
Li
J
,
Zhang
J.
2015
.
Molecular evidence for the loss of three basic tastes in penguins.
Curr Biol
.
25
(
4
):
R141
R142
.

Zhao
H
,
Zhou
Y
,
Pinto
CM
,
Charles-Dominique
P
,
Galindo-González
J
,
Zhang
S
,
Zhang
J.
2010
.
Evolution of the sweet taste receptor gene Tas1r2 in bats.
Mol Biol Evol
.
27
(
11
):
2642
2650
.

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)