You are on page 1of 59

The official journal of

INTERNATIONAL FEDERATION OF PIGMENT CELL SOCIETIES · SOCIETY FOR MELANOMA RESEARCH

PIGMENT CELL & MELANOMA


Research
The evolution of human skin pigmentation
involved the interactions of genetic,
environmental, and cultural variables
Nina G. Jablonski

DOI: 10.1111/pcmr.12976
Volume 34, Issue 4, Pages 707–729

If you wish to order reprints of this article,


please see the guidelines here

EMAIL ALERTS
Receive free email alerts and stay up-to-date on what is published
in Pigment Cell & Melanoma Research – click here

Submit your next paper to PCMR online at http://mc.manuscriptcentral.com/pcmr

Subscribe to PCMR and stay up-to-date with the only journal committed to publishing
basic research in melanoma and pigment cell biology

As a member of the IFPCS or the SMR you automatically get online access to PCMR. Sign up as
a member today at www.ifpcs.org or at www.societymelanomaresarch.org

To take out a personal subscription, please click here


More information about Pigment Cell & Melanoma Research at www.pigment.org
Received: 25 February 2021 | Revised: 30 March 2021 | Accepted: 3 April 2021

DOI: 10.1111/pcmr.12976

REVIEW

The evolution of human skin pigmentation involved the


interactions of genetic, environmental, and cultural variables

Nina G. Jablonski

Department of Anthropology, The


Pennsylvania State University, University Abstract
Park, PA, USA The primary biological role of human skin pigmentation is as a mediator of penetra-
Correspondence tion of ultraviolet radiation (UVR) into the deep layers of skin and the cutaneous
Nina G. Jablonski, Department of circulation. Since the origin of Homo sapiens, dark, protective constitutive pigmenta-
Anthropology, 409 Carpenter Building, The
Pennsylvania State University, University tion and strong tanning abilities have been favored under conditions of high UVR and
Park, PA 16802, USA. represent the baseline condition for modern humans. The evolution of partly depig-
Email: ngj2@psu.edu
mented skin and variable tanning abilities has occurred multiple times in prehistory,
as populations have dispersed into environments with lower and more seasonal UVR
regimes, with unique complements of genes and cultural practices. The evolution of
extremes of dark pigmentation and depigmentation has been rare and occurred only
under conditions of extremely high or low environmental UVR, promoted by positive
selection on variant pigmentation genes followed by limited gene flow. Over time,
the evolution of human skin pigmentation has been influenced by the nature and
course of human dispersals and modifications of cultural practices, which have modi-
fied the nature and actions of skin pigmentation genes. Throughout most of prehis-
tory and history, the evolution of human skin pigmentation has been a contingent and
non-­deterministic process.

KEYWORDS

bottleneck, depigmentation, dispersal, eumelanin, folate, tanning, ultraviolet radiation, UVB,


vasodilation, vitamin D

1 | I NTRO D U C TI O N was a breakthrough in evolution that contributed to the ability of


hominins to occupy, adapt to, and modify diverse terrestrial environ-
The evolution of mostly naked, darkly pigmented skin in the genus ments. Significant expansion of the knowledge bases in paleontology,
Homo is often not accorded the same importance in human prehis- archeology, genetics, genomics, and environmental physiology in re-
tory as the evolution of a relatively large brain, a musculoskeletal cent decades has made possible a new focus on the central role of
system adapted for sustained, striding bipedal locomotion, and the human skin and skin pigmentation in maintaining homeostasis and in
manufacture and use of stone tools. This is mostly because, until re- ensuring individual survival and reproductive success. The lability of
cently, details of the evolution of the human integument were ignored skin pigmentation, in particular, in relation to diverse environmental
or downplayed in the absence of direct fossil evidence. We now un- conditions has contributed to the ability of hominins to disperse and
derstand that the evolution of a skin interface, which facilitated sus- adapt to local circumstances over the long temporal span of human
tained high levels of physical activity in a sunny and hot environment, evolution. This facility was critical during most of prehistory when, in

This is an open access article under the terms of the Creative Commons Attribution-NonCommercial License, which permits use, distribution and reproduction
in any medium, provided the original work is properly cited and is not used for commercial purposes.
© 2021 The Authors. Pigment Cell & Melanoma Research published by John Wiley & Sons Ltd.

Pigment Cell Melanoma Res. 2021;34:707–729.  wileyonlinelibrary.com/journal/pcmr | 707


708 | JABLONSKI

the absence of body coverings or built shelters, naked skin was the the nature and expression of skin pigmentation genes, the role of
primary interface between the hominin body and the environment. genetic epistasis, population genetics processes, and mechanisms of
This review is focused primarily on the evolution of skin pig- melanin production (Crawford et al., 2017; Feng et al., 2021; Hanel
mentation during the history of anatomically modern people, Homo & Carlberg, 2020a; Lona-­Durazo et al., 2019; Norton, 2019; Pavan &
sapiens. Inferences about the characteristics of the skin and skin pig- Sturm, 2019; Quillen et al., 2019; Rocha, 2020) and attempts to show
mentation in pre-­sapiens Homo species are provided only in instances how human skin pigmentation evolved in a complex and continually
where interbreeding of these species with H. sapiens appears to have contingent manner in relation to changing complements of genes
clearly influenced skin pigmentation in descendant populations of and changing backgrounds of geography and culture.
modern people. I have taken an historical and geographic approach
here so that readers will achieve a comprehensive understanding
of how skin pigmentation evolved in groups of people dispersing at 2 | N AT U R A L S E LEC TI O N A N D TH E S K I N
different times into different environments, with different comple- O F E A R LY H O M O SA PI EN S
ments of genes, culturally determined behaviors, and material cul-
tures. Skin pigmentation is one of the most thoroughly researched of The genus Homo originated about 2.8 million years ago (mya) in
complex phenotypic traits in humans, but this expanded knowledge Africa from a representative of the genus Australopithecus (Villmoare
base has allowed us to see just how much more needs to be done, et al., 2015). The emerging consensus view is that the evolution of
especially in understanding the factors affecting the expression of mostly naked skin in the human lineage probably occurred quite
skin pigmentation genes. The importance of skin pigmentation as a early in the history of the genus Homo in order to facilitate the evap-
human phenotypic trait stems from its biological role in prehistory orative cooling of eccrine sweat during extended periods of physi-
and its diverse social meanings in history; this review will focus on cal exertion in hot environments (Jablonski, 2004; Lieberman, 2011;
the former, not the latter. Zihlman & Cohn, 1988). The onset and duration of this process are
Studying the evolution of human skin pigmentation is challenging not known with certainty, but evidence provided by comparative
because the trait has been influenced by a changing array of multi- study of patterns of neutral variation in the MC!R locus argues for a
ple variables over time. Identifying the genes and population genetic selective sweep having occurred approximately 1.2 mya ago that fa-
processes influencing human skin pigmentation is important, but vored MC1R variants producing eumelanin-­rich protective pigmen-
not sufficient, for achieving a comprehensive understanding of the tation on effectively naked skin (Rogers et al., 2004). Polymorphisms
evolution of the trait. This is because human skin pigmentation is in other genes favoring enhanced eumelanin production may have
a highly contingent trait, which has changed under different com- been selected for early in the history of the genus Homo (Feng
binations of genetic and cultural conditions over time and through et al., 2021), but the time of origin of these variants cannot be
space. It has been influenced by natural selection, population ad- fixed with certainty. The evolution of functionally naked skin was
mixture, and population bottlenecks as hominins have dispersed into also associated with changes in epidermal barrier functions and the
different biomes at various points in the past, but it has also been genes of the epidermal differentiation complex (Elias, 2005; Elias &
influenced by diverse cultural practices in different places and at Friend, 1975; Goodwin & de Guzman Strong, 2017).
different times in the past. In the last 20,000 years (kya), especially, The evolution of dark pigmentation early in the evolution of
skin pigmentation evolved under different combinations of genetic the genus Homo has been studied primarily from the perspective
and cultural factors, as humans have dispersed relatively rapidly into of identifying the nature and action of putative selective factors
diverse—­including extreme—­environments. Changes in skin pigmen- that accounted for the phenotype. The association of strong sun-
tation probably occurred rapidly in some cases as relatively small light with dark skin was appreciated by scholars of Greek antiquity
populations with restricted gene pools moved into extreme high lat- (Hippocrates, 1849) and examined as a likely cause-­and-­effect re-
itude or island environments where selective pressures were strong. lationship in the seventeenth century by Robert Boyle, Benjamin
The skin pigmentation of some groups was also probably influenced Franklin, and John Hunter (Klaus, 1973). Recognition of the role
by repeated bottlenecks and movements over many generations into played by UVR in this relationship owes to key insights made by
diverse environments with different selective pressures, especially Walter (1958), who was the first to demonstrate a high correlation
with respect to the nature and intensity of ultraviolet radiation (UVR). between skin pigmentation and latitude as a proxy for UVR. This
And, unlike other mammals, modern humans at specific places and was followed up by studies that established that UVR had a higher
times possessed unique patterns of diet, food procurement prac- correlation to skin pigmentation than environmental temperature,
tices, body coverings, and use of the built environment that would humidity, or other physical environmental parameters (Chaplin &
have affected skin pigmentation phenotypes differentially to cre- Jablonski, 2002; Roberts & Kahlon, 1976). Critical consideration of
ate unique and frequently highly localized biocultural adaptations. selective factors responsible for the UVR–­skin pigmentation rela-
While making for considerable complexity, these phenomena render tionship began with Blum's rejection of skin cancer as the likely pri-
the study of the evolution of human skin pigmentation fascinating. mary driver of increased melanin content of the skin because of its
Because of these considerations, this review differs from excellent limited effect on mortality (Blum, 1961). The most important result
recent reviews of this subject because it builds on information on of Blum's contribution was his focus on the importance of survival,
JABLONSKI | 709

reproductive success, and mechanisms of natural selection in de- of widespread longevity and the importance of grandparental
fining adaptive traits, as opposed to identification of factors with care of offspring needs thus needs to be questioned (Osborne &
negligible or unknowable selective value. Most of the hypotheses Hames, 2014). It is also important to choose suitable modern human
advanced for the evolution of skin pigmentation in the last 70 years models or model systems when framing theoretical discussions of
have failed in this regard, for reasons fully described and reviewed the evolution of skin pigmentation because choices of inappropriate
elsewhere (Jablonski, 2004; Jablonski & Chaplin, 2000). model systems result can result in misleading conclusions. An import-
Two sets of hypotheses for the evolution of skin pigmentation ant case in point is another hypothesis proposing protection against
that are not focused on the primacy of UVR intensity warrant men- skin cancer as the primary selective agent favoring the evolution of
tion here because they have stressed the importance of eumelanin permanent dark skin pigmentation. In this study, high rates of fatal
in enhancing specific barrier functions of the skin. One set focused skin cancers in depigmented people with albinism living in high UVR
on the antimicrobial action of eumelanin in the epidermis and its environments were presented as evidence that the most important
heightened importance in combating parasitic and waterborne function of eumelanin was protection against UVR-­induced damage
disease agents attacking the skin as being the likely agent for se- to DNA and its connection to diminished reproductive success and
lection of darkly pigmented skin in the tropics (Mackintosh, 2001; premature death (Greaves, 2014). This hypothesis was based on the
Wassermann, 1965). Another set of hypotheses, championed by Elias assumption that ancestral hominins had lightly pigmented or depig-
and colleagues, has stressed the importance of increased melaniza- mented skin and that modern people mostly or entirely lacking eu-
tion of the skin in enhancing the competence of the epidermal melanin in their integument might a suitable model for the ancestral
permeability barrier in desiccating terrestrial environments (Elias, human state. The notion that very lightly pigmented people or indi-
Menon, Wetzel, & Williams, 2009, 2010; Elias & Williams, 2013). viduals with albinism could serve as models for understanding the
The explanatory power of these hypotheses to account for known selective forces operating on the skin of early hominins is wrong. All
patterns of human skin pigmentation and, in particular, the distribu- living catarrhine primates, including the closest living ape relatives
tion of eumelanin-­rich skin has been questioned on several grounds, of humans, have intact pigmentary systems and can develop faculta-
including the fact that the skin of the lips and the volar surfaces of tive pigmentation through tanning of non-­hairy exposed skin or have
the hands and feet exhibits a near absence of eumelanin pigmenta- permanently melanized exposed skin (Jablonski & Chaplin, 2014b).
tion in all people despite being subjected to some of the most rigor- In the human lineage, lightly pigmented skin and depigmented skin
ous physical and antimicrobial challenges of any skin on the body's due to albinism are highly derived, not ancestral, conditions. The
surface (Jablonski & Chaplin, 2013). The weight of current evidence manifold protective effects of eumelanin in the skin, especially with
thus supports the theory that UVR is the primary selective agent respect to mitigation of UVR damage and prevention of carcino-
which has influenced the evolution of human skin pigmentation. genesis, have been thoroughly researched and reviewed elsewhere
The distribution and intensity of the different wavelengths of (Abdel-­Malek, 2009; Hennessy, Oh, Diffey, et al., 2005; Swope &
UVR vary over terrestrial surfaces according to time of day, sea- Abdel-­Malek, 2018; Tadokoro et al., 2003; Young et al., 2017). More
son, geographical latitude, altitude, cloud cover, and other factors germane in the context of this review is discussion of the likely selec-
(Diffey, 1991). Individual exposure to UVR depends on the intensity tive value of different degrees of eumelanin pigmentation on health
of the ambient UVR in a given location, the fraction of ambient ex- and potential reproductive success. In this context, we cannot as-
posure received at a given anatomical site, and the timing and nature sume that the appearance, genetic composition, sun exposure hab-
of outdoor activity (Diffey, 1999). In assessing the effects of UVR on its, or UVR skin reactions of humans today are the same as those of
the skin and weighing the relative importance of these effects on the our ancestors in prehistory. Choice of appropriate models for ances-
evolution of skin pigmentation, it is important to appreciate that hu- tral or ancient hominin appearance, physiology, or behavior must be
mans in prehistory exhibited different lifestyles and patterns of UVR done with great care and will always involve uncertainty because of
exposure than modern people. While this may seem self-­evident, it the many kinds of variables involved. Consideration of the evolution-
is rarely taken into account. ary significance of eumelanin-­rich skin pigmentation in the human
The relative importance of different selective factors in the evo- lineage must be informed by this principle.
lution of human skin pigmentation depends on their potential effects
on reproductive success. A potential selective agent like UVR may
damage connective tissue and DNA, for instance, but if the effects 2.1 | Folate-­dependent processes and their role
of the damage do not adversely affect the individual's reproductive in the evolution of eumelalin-­rich pigmentation
success, or if the number of individuals adversely affected is small,
then its evolutionary impact is diminished. In this connection, the Research conducted first in the 1970s suggested that eumelanin in
patterns of life history, longevity, and demographics observed in the skin may be protective against degradation of folate (Branda &
modern people cannot be assumed for early members of the genus Eaton, 1978). Appreciation of the centrality of folate in regulating
Homo or even prehistoric H. sapiens. A recent proposal advancing DNA synthesis and repair grew rapidly in subsequent years, espe-
the importance of skin cancer as a selective agent in the evolu- cially after the connection between folate deficiencies and the then
tion of human skin pigmentation that was based on an assumption common class of birth defects, neural tube defects, became widely
710 | JABLONSKI

recognized on the basis of epidemiological and experimental studies UVR absorptive abilities, which slow the formation of highly muta-
(Bower & Stanley, 1989; Elwood, 1983; Fleming & Copp, 1998; MRC genic cyclobutane pyrimidine dimers (CPD) (Fajuyigbe et al., 2018).
Vitamin Study Research Group, 1991). Folate, primarily in the form The localization of most eumelanin in the stratum basale also serves
of 5-­methyltetrahydrofolate (5-­MTHF), is used at the cellular level to protect the capillary bed of the dermis from UVR, especially UVB.
for DNA production, the cysteine cycle, and regulation of homocyst- Penetration of UVA into the stratum basale is considerably greater
eine. Knowledge of the specific actions, enzymatic and genetic regu- and potentially more harmful in lightly pigmented skin (Meinhardt
lation, and environmental sensitivity of folate grew quickly in light of et al., 2008). Attention in this connection has been paid primarily to
increasing appreciation of the vitamin's extensive clinical significance the cosmetic effects of UVA on the breakdown of connective tissues
(Branda & Blickensderfer, 1993; Giovannucci et al., 1995; Lucock & in the dermis (Battie et al., 2014), but the demonstrated effects of
Daskalakis, 2000; Lucock et al., 2001; Mastropaolo & Wilson, 1993). UVA on vasodilation of arteries in the capillary bed of the dermis (Liu
Appreciation of the importance of folate to DNA production and et al., 2014) suggest that control of blood pressure may be of more
cell proliferation during embryogenesis and spermatogenesis and fundamental importance.
its likely sensitivity to UVR lead to initial formulations of a hypoth- Evidence for the hypothesis that permanent dark pigmentation
esis about the protective effect of eumelanin pigmentation with in hominin skin was an adaptation to protect against UVR-­induced
respect to successful pregnancy outcomes (Jablonski, 1992, 1999). degradation of folate in the skin was originally based on direct
This work was further elaborated when it became possible to exam- effects on fertility such as potential embryo loss or faulty sper-
ine patterns of human skin pigmentation (as provided in reports of matogenesis (Jablonski & Chaplin, 2000). These effects are hard
standardized skin reflectance measurements taken from indigenous to demonstrate in living human subjects or through retrospective
peoples) relative to levels of remotely sensed UVR as measured by epidemiological studies. Folate levels and folate metabolism are
the NASA Total Ozone Mapping Spectrometer 7 (TOMS-­7) satellite affected by primary folate deficiency, non-­folate B-­vitamin nutri-
(McPeters et al., 1996). The TOMS-­7 and subsequent NASA satellites ent deficiencies, and genetic variations that influence cellular folate
revolutionized the study of the effects of UVR on biological systems accumulation and utilization (Perry et al., 2004; Stover, 2009), and
because they provided standardized, global, spatially continuous these factors were generally not considered in 20th-­century sur-
and gridded data that made it possible to assess the potential for veys of folate levels within and between populations. The clearest
biological damage due to solar irradiation, given the column ozone epidemiological demonstration of the possible folate-­
sparing ef-
amount and cloud conditions on any day, for most places on earth fect of dark pigmentation against UVR challenge came from a study
(Langston et al., 2017). The demonstration that human skin pigmen- demonstrating the lower prevalence of NTDs in darkly pigmented
tation was more highly correlated to UVB than to latitude (Jablonski as opposed to lightly pigmented South Africans, despite the greater
& Chaplin, 2000) or environmental parameters such as total solar ir- affluence and presumed better diet of the latter group (Buccimazza
radiance, temperature, humidity, or rainfall. Chaplin (2004) indicated et al., 1994). Experimental studies of the effects of UVA and UVB on
that UVR was the most important environmental determinant of skin serum and red cell folate conducted in vitro and in vivo have yielded
pigmentation. mixed results and have been difficult to interpret because of incon-
The nature of the specific protective effect of eumelanin in sistencies in experimental conditions and folate assay standardiza-
the epidermis and its effects of evolutionary fitness depends fun- tion (Borradale et al., 2014; Borradale & Kimlin, 2012; Fukuwatari
damentally on where the pigment is localized in the skin. Different et al., 2009; Gambichler et al., 2001; Juzeniene et al., 2009; Wolf
wavelengths of UVR penetrate the skin to different depths of the & Kenney, 2019), but most studies have found depletion of folate
epidermis and dermis, depending on the skin site, the amount and to varying degrees, depending on the wavelength and duration of
location of eumelanin, and the degree of keratinization of the skin UVR exposure, the pre-­exposure folate level, and the species of fo-
(Meinhardt et al., 2008). Eumelanin is concentrated within keratino- late being assayed. Experimental demonstration of a decline in epi-
cytes in the stratum basale of the epidermis and in this position can dermal 5-­MTHF following UVA exposure in individuals with lightly
afford considerable protection against damage by UVR, especially pigmented but not darkly pigmented skin (Hasoun et al., 2015) is
UVB (Fajuyigbe et al., 2018; Fajuyigbe & Young, 2016). The stratum potentially more significant because it indicates the importance of
corneum itself affords some protection against UVB, but the degree UVR-­induced folate depletion in the skin after a short UVR exposure
of protection afforded depends on the thickness of individual layers, and possible systemic depletion following prolonged exposure.
the number of layers, the degree of hydration of epidermis (Bruls Appreciation of the central importance of folate on human health
et al., 1984). The protective effect of stratum corneum itself is more through its effects of DNA synthesis and methylation (Stover, 2009)
significant with respect to UVB shielding in skin with less eumela- has been augmented in the last 20 years by research demonstrat-
nin content, but the magnitude of this effect is variable and, overall, ing direct effects of folate on vascular function. These effects are
minor compared to that of eumelanin itself (Hennessy et al., 2005; pertinent to the evolution of skin pigmentation because of the role
Kaidbey et al., 1979; Olivarius et al., 1997). played by folate (as 5-­MTHF and as its further derivative, tetrahydro-
Most of the attention paid to the UVB shielding effects of eu- biopterin, or BH4) in improving nitric oxide (NO)-­mediated endothelial
melanin has been concerned with the sparing of damage to DNA function and affecting vasodilation, blood pressure, and thermoregu-
and mitigation of carcinogenesis because of the pigment's significant lation (Alexander et al., 2013; Ng et al., 2009; Stanhewicz et al., ,2012,
JABLONSKI | 711

2015). UVA triggers cutaneous vasodilation, probably by releasing to maximize cutaneous photoconversion of 7-­dehydrocholesterol (7-­
NO from preexisting stores of nitrite in the dermis (Liu et al., 2014). DHC) into vitamin D under low UVB conditions at high latitudes and
The discovery that lower NO levels contributed to attenuated vaso- that “black skins” had evolved to protect against vitamin D toxicity
dilation in response to local heating of the skin of older individuals due to potential overproduction of vitamin D at low latitudes under
and that these responses could be restored by folate administra- high UVB conditions (Loomis, 1967). Research into the mechanisms
tion (Kenney, 2017) led to studies aimed at discovering the likely of cutaneous vitamin D synthesis by Holick and colleagues provided
molecular mechanisms whereby folate improved NO bioavailability evidence that the process was tightly photochemically controlled
and protected endothelial function (Stanhewicz & Kenney, 2017). and that the eumelanin content of the skin determined the rate of
Recent findings indicating that acute UVB exposure attenuated NO-­ conversion of 7-­DHC to previtamin D3 (pre-­D). (Clemens et al., 1982;
mediated vasodilation of the cutaneous microvasculature via degra- Holick et al., 1980, 1981). This body of research demonstrated that
dation of 5-­MTHF led to the pursuit of studies examining the effects increased eumelanin content of the skin could not be implicated in
of skin pigmentation, age, and folate status on vasodilation and ther- preventing vitamin D toxicity, but that depigmented skin facilitated
moregulation (Wolf & Kenney, 2019). Available evidence indicates cutaneous vitamin D production because of the fact that eumelanin
that folate degradation in response to UVR exposure is likely dose-­ competes with 7-­DHC for UVB photons (Holick et al., 1981). This
dependent and requires relatively large doses, repeated prolonged research laid the groundwork for subsequent studies establishing
exposures, or both to observe reductions in serum or red blood cell the nature and genetic basis of the selective forces acting on the
folate (Wolf & Kenney, 2019). These environmental conditions would evolution of skin pigmentation in populations dispersing into and
have been met in the early history of the genus Homo when hominins inhabiting regions of lower and more highly seasonal UVB (Beleza,
were active in the sun or at least exposed to strong, equatorial sun- Santos, et al., 2013; Gozdzik et al., 2008; Hanel & Carlberg, 2020a;
light for most daylight hours without the benefit of body coverings. Izagirre et al., 2006; Jablonski & Chaplin, 2000, 2010; Makova &
The sum of the effects of UVR-­induced folate loss on hominins Norton, 2005).
cannot be accurately estimated, but there is little doubt that low- The body of research pertaining to the importance of vitamin D
ered fertility due to embryo loss and impaired spermatogenesis in bone metabolism and many cellular and immunological processes
combined with impairments of vasodilation and temperature regu- is now immense and is reviewed elsewhere (Lips, 2006, 2007; Lips
lation would have represented direct and severe selective pressures et al., 2014; Sassi et al., 2018). Issues raised about the strength of
favoring conservation of folate through increased eumelanin in the natural selection acting to reduce eumelanin pigmentation in order
functionally naked skin of early Homo. This does not mean that early to enhance cutaneous vitamin D synthesis (Elias et al., 2009; Elias &
Homo had “dark skin,” but that selection favoring enhanced eumel- Williams, 2013, 2016; Robins, 2009) have been countered by con-
anin pigmentation in the skin was evolving early in the evolution siderable evidence. The claim that loss of eumelanin and enhance-
of the genus, along with polymorphisms in genes affecting folate ment of cutaneous vitamin D production did not significantly improve
metabolism. Because folate metabolism is regulated by several en- the reproductive success of hominins dispersing to high latitudes
zymes, including methylenetetrahydrofolate reductase (MTHFR), (Robins, 2009) was called into question by challenging incorrect as-
methionine synthase (MTR), and methionine synthase reductase sertions, including the claim that, even at the highest latitudes, stored
(MTRR), and because the activity of these enzymes varies according vitamin D alone is sufficient to meet physiological needs during
to different functional polymorphisms, these polymorphisms appear months when the UVB in sunlight is insufficient to catalyze cutaneous
to have evolved in response to seasonal availability of folate and lev- vitamin D production (Chaplin & Jablonski, 2009, 2013). Other claims
els of UVR (Jones et al., 2018; Lucock et al., ,2010, 2017). It is thus centering around reduction in epidermal eumelanin having been se-
probable that multiple mechanisms, including enhanced eumelanin lected for altered epidermal barrier functions have also been found
pigmentation and conservation of folate through modification of en- deficient. These have focused on the significance of reduced levels of
zymatically controlled metabolism, evolved in the face of seasonally the protein filaggrin in the stratum corneum, leading to enhanced cu-
dependent variations in folate availability and UVR levels. taneous synthesis of pre-­D and also on the argument that reduction
in eumelanin production was favored by natural selection because of
the importance of conserving energy (Elias & Williams, 2013). These
2.2 | Vitamin D and skin pigmentation explanations failed to demonstrate any significant evolutionary ad-
vantage for reduction of filaggrin (Jablonski & Chaplin, 2013), and
The importance of vitamin D in connection with the evolu- further study of variation in loss of function genes affecting filaggrin
tion of human skin pigmentation was first introduced by Murray in the epidermis failed to demonstrate a relationship between genetic
(Murray, 1934) and later elaborated by Loomis (Loomis, 1967). These variants and latitude (as a surrogate for UVB and cutaneous vitamin D
were based, at that time, on the recognition of the physiological im- synthesis potential) (Eaaswarkhanth et al., 2016).
portance of vitamin D in the growth and maintenance of the skeleton A key consideration in understanding the evolution of skin pig-
and understanding that production of vitamin D could be catalyzed mentation in people living under high and less seasonally variable
only by specific wavelengths of UVR, from about 290–­320 nm, in the UVR conditions—­generally within the tropics—­is the degree to which
UVB range. Loomis originally opined that “white skins” had evolved eumelanin pigmentation actually slows the cutaneous biosynthesis
712 | JABLONSKI

of vitamin D. Recent experimental work done by Young and col- with skin pigmentation. As modern populations have dispersed and
leagues on human subjects has demonstrated that darkly pigmented undergone bottlenecks and admixture, variant genes encoding for
individuals (Fitzpatrick type VI according to study protocols) exhib- proteins responsible for photoconversion of 7-­DHC into pre-­D (Kuan
ited a melanin vitamin D inhibition factor of 1.3–­1.4 relative to lightly et al., 2013) and governing transport, metabolism, and signaling of
pigmented individuals (Fitzpatrick type II) (Young et al., 2020). They vitamin D, including DHCR7, GC, CYP2R1, and CYP24A1 (Hanel &
reasoned that this significant, but relatively modest, inhibitory ef- Carlberg, 2020a, 2020b), have undergone changes in frequency
fect may have been due to the fact that photoconversion of 7-­DHC appeared and become common. Genetic changes affecting vitamin
into pre-­D occurs mostly in the stratum granulosum and stratum spi- D metabolism rather than skin pigmentation have been particularly
nosum of the epidermis, above the heavily melanized layers of the important (Hanel & Carlberg, 2020a, 2020b), especially at extreme
basal epidermis (Young et al., 2020). northern high latitudes where levels of UVB are low and highly sea-
Vitamin D serves many critical functions in the body, and de- sonal, creating conditions for only short and sporadic cutaneous
ficiency in the vitamin increases susceptibility to a range of devel- photosynthesis of vitamin D (Chaplin & Jablonski, 2013; Jablonski
opmental, chronic, and infectious diseases via its effects on the & Chaplin, 2010). Lastly, the nature of the genetic changes occur-
epigenome and on the expression of many genes on nearly all organs ring was being mediated by numerous lifestyle variables (includ-
and tissue types in the body (Bora & Cantorna, 2017; Bustamante ing diet, typical body coverings and kinds of shelter, and patterns
et al., 2020; Caccamo et al., 2018; Carlberg, 2019; Sassi et al., 2018). of daily activity) which affected vitamin D status and would have
The recognition that vitamin D improves endothelial function by contributed to individual survival and reproductive success. Thus,
signaling for the transcription of endothelial nitric oxide synthase consideration of the contingent nature of biocultural compromises in
(eNOS) and thereby acting to preserve production of NO and different places and at different times is essential to the study of the
healthy peripheral vascular function augments our understanding of evolution of human skin pigmentation and human health (Chaplin
the vitamin's consummate importance in human physiology and has & Jablonski, 2013). Genetically driven changes in skin pigmentation
potentially great consequences for our understanding of the evolu- favoring less integumental eumelanin were only one factor in the
tion of skin pigmentation (Wolf et al., 2020; Wolf & Kenney, 2019). “adaptive equation” contributing to healthy vitamin D status.
Vitamin D is produced in the upper epidermis through UVB-­induced
catalysis of 7-­DHC into pre-­D, and it is required in the dermis to
preserve NO production required for healthy vasodilation. This im- 2.3 | Likely skin pigmentation of early Homo sapiens
portant function of vitamin D probably accounts for the intriguing
finding of reduced all-­cause mortality in a large cohort of lightly pig- The probable loss of most body hair during the early evolution of the
mented Scandinavian women reported by Lindqvist and colleagues genus Homo left the body highly vulnerable to the effects of UVR,
(Lindqvist et al., 2020). The role of vitamin D’s role in eNOS signaling, as well as to many other potentially harmful physical, chemical,
control of vasodilation, and regulation of blood pressure warrants and biological agents. The evolution of protective eumelanin-­rich
further investigation in an evolutionary context with respect to the constitutive pigmentation in the tropical-­dwelling, African ances-
evolution of skin pigmentation under low UVR conditions (Wolf tors of all modern people was probably a step-­by-­s tep process in-
et al., 2020). It is also of potentially great relevance in clinical con- volving many genetic loci. Early in the history of skin pigmentation
texts with respect to the importance of low and carefully controlled genomics, considerable attention was focused on the primacy of
doses of natural UVR sufficient to produce vitamin D and maintain a selective sweep affecting the MC1R locus to eliminate variation
healthy vasodilation and blood pressure (Alfredsson et al., 2020; and establish permanent dark pigmentation in the human lineage
Wolf et al., 2020). This consideration is of particular concern in ha- (Rana et al., 1999; Rogers et al., 2004). Since those early studies,
bitually indoor-­dwelling people, older people with attenuated cuta- the field of skin pigmentation genomics has been revolutionized
neous vitamin D production abilities, and darkly pigmented people by the development of high-­throughput sequencing technologies
living in low UVR environments. coupled with the application of genome-­wide association studies
The manifest importance of vitamin D in human physiology im- (GWAS) for identifying multiple genes associated with human skin
plies that complex, genetically based mechanisms for establishing pigmentation and the execution of genome-­wide scans of natural
and maintaining vitamin D sufficiency have evolved in the course selection. These developments have made possible the identifica-
of human evolution. Modification of the production and packaging tion of many genes contributing to skin pigmentation, and the de-
of eumelanin in the skin affects the penetration of UVB into the gree to which they have been acted upon by natural selection (Feng
epidermis and the physical potential for photoconversion of 7-­DHC et al., 2021). For many loci of interest, application of coalescent
into pre-­D, but levels of vitamin D can and are affected by many studies on specific derived alleles has made possible estimation of
other processes and pathways. Advances in genomics and archeog- when was the most recent common ancestor of existing derived
enomics have made possible a significant shift in our understanding alleles first occurred.
of these processes in recent years, to the extent that there is now Based on these studies, it is possible to estimate that multiple
greater appreciation of how the evolution of genetic variants associ- derived alleles contributed to the evolution of dark pigmentation in
ated with vitamin D metabolism and signaling has evolved in concert ancestral H. sapiens. These include two derived alleles of the major
JABLONSKI | 713

facilitator superfamily domain-­


containing protein 12 (MFSD12), 3 | S K I N PI G M E NTATI O N E VO LU TI O N I N
which appeared at or near the time of origin of H. sapiens, about R E L ATI O N TO D I S PE R SA L S , G E N E S , A N D
0.3 mya (Crawford et al., 2017; Feng et al., 2021). Other candidate CU LT U R E I N PR E H I S TO RY
loci, including multiple alleles of the DNA damage-­binding protein
1 (DDB1), may also have been present in early H. sapiens because The processes affecting the evolution of human skin pigmenta-
of their presence at high frequencies in modern African and in tion have changed in relative importance over time. In the last
some non-­European populations today (Crawford et al., 2017; Feng 100,000 years (100 kya), H. sapiens has gone from being an exclu-
et al., 2021) (Figure 1), thus suggesting that derived alleles were sively African species to a global one. The nature and speed of major
carried in one of the populations that dispersed along the southern dispersals of modern humans varied according to many geographic,
and southeast Asian coasts and into Melanesia. Detailed examina- demographic, technological, and cultural variables. With gains in tech-
tion of the geographical pattern of variation in one of these alleles, nological competence and, especially, with the domestication of ani-
rs7948623 (T), further suggests that it may have been under positive mals used for food and transportation, humans became more mobile
selection, resulting in further skin darkening, in regions of the world and had better abilities to buffer themselves against the exigencies
including East Africa, parts of South Asia, and Melanesia with ex- of the environment because they could control more aspects of per-
tremely high environmental UVR (Feng et al., 2021). The key point sonal nutrition and bodily protection through technology and cultural
is that the ancestor of all H. sapiens had darkly pigmented, but not practices. Through time, and especially since the beginning of history
necessarily maximally darkly pigmented, skin (Chaplin, 2004; Hanel sensu stricto about five thousand years ago, culturally based prefer-
& Carlberg, 2020a). In Africa today, and among many South Asians ences for skin color (including sexual selection and social selection)
and Melanesians, there is great variation in the darkness of “darkly have also influenced regional trends in the skin pigmentation. For all
pigmented” skin that appears to be due to variations in UVR inten- of these reasons, there has probably never been a time or place in the
sity (Chaplin, 2004; Jablonski & Chaplin, 2014a). Genomic studies history of our species when skin pigmentation has been in equilibrium.
of the last two decades have revealed that the genetic “palette” of The following review of the history of skin pigmentation in
skin pigmentation gene variants is great and that many combinations the major regions of the world is, perforce, speculative because of
of multiple variant genes have contributed to the complex pattern the multiplicity of variables being considered. The level of detail
of skin pigmentation phenotypes and genotypes observed today. presented here falls far short of that covered in global reviews of
Under more highly seasonal UVR regimes, the evolution of enhanced human dispersals that have examined the details of human move-
tanning abilities and genes contributing to rapid development of fac- ments from the perspectives of genomics, genetics, and population
ultative pigmentation has been favored (Quillen et al., 2019). Despite genetics (James et al., 2019; Nielsen et al., 2017; Posth et al., 2016).
the manifest importance and relevance of tanning genetics to skin Nonetheless, they form a basis for discussion and future research.
cancer dynamics and skin esthetics, relatively little genomics-­based
research on the topic has been done, especially in African and in-
digenous American populations. Although advances in evolutionary 3.1 | Africa
skin pigmentation genetics have been enormous, there is an urgent
need to validate candidate skin pigmentation variants using GWAS The earliest history of H. sapiens occurred in Africa and has been
and scans of natural selection, using functional experiments in vitro reconstructed through a combination of sparse fossil evidence and
and in vivo (Feng et al., 2021). inferences from archeogenomics (Galway-­Witham & Stringer, 2018;

F I G U R E 1 The “hairy timeline of human evolution” illustrating, schematically, the evolution of body hair and skin color in pre-­Homo
sapiens members of the human lineage. Prior to the evolution of early members of the genus Homo, about 2.8 mya, hominin skin, was lightly
pigmented (but tannable) and covered with dark hair. The evolution of mostly hairless bodies beginning about 2 mya inaugurated a selective
sweep eliminating MC1R polymorphism, after which time selection for dark pigmentation variants of DDB1 and MFSD12 persists through the
time of origin of H. sapiens, about 0.3 mya
714 | JABLONSKI

F I G U R E 2 Generalized timeline of
main events in the evolution of skin
pigmentation in Africa, by region. After
originating around 0.3 mya, H. sapiens
evolved in Africa only until about 80 kya
when some dispersals from northeastern
Africa into the Afro-­Arabian Peninsula
occur. (a) In East Africa, continued
purifying selection on MC1R variants is
accompanied by positive selection on
variants of DDB1 and MFSD12 leads to
darker pigmentation. The introduction
of the classic depigmentation SLC24A5
variant around 9 kya via the Afro-­
Arabian Peninsula sees its rapid dispersal
southward. (b) In West Africa, purifying
selection on MC1R variants continues,
accompanied by positive selection on
variants of DDB1 and MFSD12 producing
darker pigmentation. (c) In South Africa,
among Khoe-­san hunter-­gatherers,
relaxation of selection on MC1R is
accompanied by positive selection on the
SLC24A5 introduced from northeastern
Africa. The dispersal of darkly pigmented
Bantu-­language speaking agriculturalists
from western and central Africa from
5 ka onward and their admixture with
Khoe-­san hunter-­gatherers leads to
the evolution of complexly mixed
pigmentation genotypes and a wide range
of moderately pigmented phenotypes

Henn et al., 2018; Hublin et al., 2017; Stringer, 2016; White of Africa are considered. Other reasons are demographic. Until the
et al., 2003). The skeletal traits which define H. sapiens emerged in advent of agriculture and animal husbandry in Africa about 5 kya,
Africa over the course of about 100,000 years, from 0.3 to 0.2 mya most populations of hunter-­gatherers were relatively small, scat-
(Galway-­Witham & Stringer, 2018). By 0.2 mya, diverse forms of tered, and probably experienced repeated bottlenecks and periods
anatomically modern humans lived in Africa, mostly in dispersed of expansion due to sometimes dramatic environmental perturba-
regional and environmental enclaves, and pursued technologi- tions (Ambrose, 1998; Hsieh et al., 2016; Powell et al., 2009).
cally and artistically complex but locally distinctive cultures (Scerri In the specific context of human skin pigmentation evolution in
et al., 2018). The genetic history of modern people in Africa thus Africa, the diversity of UVR regimes in the African continent, includ-
reflects deep roots, and complex patterns of isolation and reticula- ing the highly seasonal pattern of UVB in the far south, is one of many
tion. Some of the reasons for this can be readily appreciated when factors that must be appreciated (Coussens et al., 2015; Jablonski &
the vastness, ecological diversity, and remarkable latitudinal extent Chaplin, 2014a). These patterns are reflected in the complex mosaic
JABLONSKI | 715

of genes that contribute to skin pigmentation across Africa today Linguistic, archeological, and genetic evidence for the movement
(Crawford et al., 2017; Feng et al., 2021; Lin et al., 2018; Martin of Bantu-­speaking agriculturalists attests that this was one of the
et al., 2017; Rocha, 2020). The strong and relatively seasonally in- largest and most genetically impactful dispersals in human history
variant UVR regimes of East Africa, including the Horn of Africa and because of its effects on preexisting populations of rainforest-­and
the East African coast, have continued to drive selection for increas- desert-­dwelling hunter-­gatherers in western, central, and southern
ingly dark skin pigmentation in these regions (Chaplin, 2004; Feng Africa (Patin et al., 2017). With respect to skin pigmentation, the in-
et al., 2021; Jablonski & Chaplin, 2000, 2014a). This trend has also gress of very darkly pigmented people into southern Africa in recent
been influenced by the wearing of relatively little clothing during millennia appears to have adversely affected vitamin D status and
daylight hours and while active, up to and including historical times, health, especially in modern, mostly urban populations in southern-
because of extreme environmental heat. A trend toward increas- most Africa (Coussens et al., 2015). A hypothetical skin pigmentation
ing pigmentation darkening also appears to have occurred in West timeline for the major regions of Africa is presented in Figure 2.
Africa, but the combination of genetic variants contributing to this
is less clear (Crawford et al., 2017; McEvoy et al., 2006). The nature
of the genes and genetic epistasis contributing to the enhancement 3.2 | Coastal south and southeast Asia,
of dark constitutive pigmentation and tanning abilities in equatorial Melanesia, and Australia
African people (and elsewhere) is still poorly understood, but it is
likely that insights will come from functional genomic studies and The timing and routes of the dispersal of H. sapiens into Eurasia, and
those focused on the study of recently admixed populations in order the population sizes of dispersing populations, have been studied in-
to determine causal effects (Beleza, Johnson, et al., 2013; Lona-­ tensively by geneticists and genomicists in the last 20 years and are
Durazo et al., 2019; Rocha, 2020). now reasonably well understood (Henn et al., ,2012, 2016; Nielsen
The other major phenomena that have influenced the evolu- et al., 2017; Pagani et al., 2016; Reyes-­Centeno et al., 2014). The
tion of skin pigmentation in Africa were recent human dispersals, number, timing, and exact routes of these dispersals are still de-
but not those associated with European colonialism. The first of bated, but genetic and paleontological evidence indicates that the
these was the dispersal of people from Eurasia via the Afro-­Arabian first groups of H. sapiens dispersed out of Africa beginning about
Peninsula into eastern and southern Africa. These people carried 70 kya, following a primarily coastal route from the Afro-­Arabian
derived variants of SLC24A5 (solute carrier family 24 [sodium/potas- Peninsula into south and southeast Asia, thence into Melanesia and,
sium/calcium exchanger], member 5) associated with depigmented eventually, into Australia (Groucutt et al., 2015; James et al., 2019;
skin (Martin et al., 2017). These variants spread relatively quickly Malaspinas et al., 2016; Pagani et al., 2016; Posth et al., 2016; Reyes-­
southward, beginning around 5,000 yr (Crawford et al., 2017), and Centeno, 2016; Reyes-­Centeno et al., 2014). Divergence of popula-
were favored under positive selection in the far southern latitudes tions ancestral to modern Papuans and Aboriginal Australian from
of Africa among Khoe-­san hunter-­gatherers and pastoralists, espe- Eurasians is estimated from genomic evidence to have occurred
cially in the last 2,000 yr (Lin et al., 2018; Martin et al., 2017). The 72–­51 kya, after which time admixture of ancient Denisovans with
evidence for positive selection for SLC24A5 variants that conferred an ancestral Papuan-­Australian population occurred (Malaspinas
some skin lightening in the indigenous Khoe-­san peoples inhabiting et al., 2016).
the lower and more seasonal UVR conditions of southern Africa is Populations dispersing and living along the coasts of southern
significant. It complements evidence that non-­
synonymous mu- and southeastern Asia in the Late Pleistocene experienced strong
tations in the MC1R locus exist in the Khoe-­san (John et al., 2003) and seasonally relatively invariant UVR levels, favoring maintenance
and that other variant forms of MC1R may have arisen in southern of dark pigmentation, probably involving selection on ancestral gene
African under relaxed selection. It also illustrates that a new genetic variants MFSD12 and DDB1 (Feng et al., 2021) and variant forms of
variant can achieve high frequency in human populations after being SLC24A5 and OCA2 (Jinam et al., 2017; Stokowski et al., 2007). The
introduced if it affects a trait that positively impacts fitness. This contribution of Denisovan genes to these populations is recognized
result is also important because it puts paid to arguments that the (Gittelman et al., 2016; Jinam et al., 2017) but the genetic contri-
relatively light skin pigmentation of the Khoe-­san was due to genetic bution of Denisovans to skin pigmentation is not yet clearly estab-
admixture with Europeans early in colonial history. lished. In considering the evolution of skin pigmentation in these
The second major dispersal event within Africa that has affected populations, the effects of population size, diet, and habitual raiment
the distribution of skin pigmentation phenotypes and the continu- must be considered, even if the strength of these influences cannot
ing evolution of skin pigmentation by population admixture is the be quantified easily. The populations dispersing along these coasts
expansion of Bantu-­language speakers from western Africa into were hunter-­gatherers and probably of small population size. Some
central, eastern, and southern Africa, beginning about 5.6 kya (Li of these, notably several so-­called “Negrito” populations, manifest
et al., 2014; Patin et al., 2017). The timing and nature of dispersals genetic evidence of severe population bottlenecks as well as recent
of Bantu-­speaking peoples have been characterized using genome-­ admixture (Jinam et al., 2017). Based on ethnographic accounts of
wide microsatellite markers and are now understood to have, in gen- living Andamanese, these people ate traditional diets of fish, wild
eral, proceeded from west to east, and then south (Li et al., 2014). boar, shellfish, turtle, and turtle eggs, along with foraged tubers and
716 | JABLONSKI

fruits (Headland, 1989; Sahani, 2010). They also wore few, if any, extremes of skin pigmentation evolution, occurring as the result of
body coverings. This lifestyle conduced to a high level of vitamin people being isolated in geographic dead-­ends, while simultaneously
D sufficiency, thereby rendering unlikely the development of any experiencing little admixture and positive selection. In these cases,
level of vitamin D insufficiency as long as a traditional diet was being diets replete in vitamin D may have worked to further promote the
consumed. This may have released a selective constraint on further evolution of increasingly dark, protective eumelanin pigmentation
pigment darkening. up to a threshold level. This hypothesis invites further study of genes
Knowledge of the skin pigmentation phenotypes and pigmen- affecting skin pigmentation, vitamin D metabolism, and the eumel-
tation genes of Melanesians and Australian Aboriginals is limited, anin threshold effect. A hypothetical skin pigmentation timeline for
but the available data indicate that skin pigmentation is very dark, coastal South and Southeast Asia, Melanesia, and Australia is pre-
with Bougainvilleans exhibiting the darkest levels of pigmentation sented in Figure 3.
measured by reflectometry among people living today (Jablonski
& Chaplin, 2000; Norton et al., 2006). Similar, extremely dark
pigmentation reportedly existed among Aboriginal Tasmanians 3.3 | Hinterland Eurasia and the Americas
(Diamond, 2005). Melanesian populations notably lack evidence
for purifying selection at the MC1R locus (Norton et al., 2015) and, The evolution of skin pigmentation in Eurasia, especially in Europe,
thus, their extremely dark pigmentation must be accounted for by has been the focus of considerable attention (Norton, 2019; Quillen
enhanced eumelanin production and tanning abilities controlled by et al., 2019), especially with the advent of archeogenomics and
other genes, possibly MFSD12 and DDB1, and others. Under these the functional genomic interpretation of skin pigmentation genes
conditions, dark skin pigmentation evolved to what appears to be in ancient Eurasian populations. New and comprehensive reviews
a threshold level, past which no further darkening appears to be of Eurasian skin pigmentation genomics in relation to historical
possible (Chaplin, 2004). It is noteworthy that the people who ex- genomics should be consulted for details, especially with regard
hibit some of the darkest skin pigmentation observed are not only to the movements of peoples and genes in historic times and their
exposed to extremely high environmental UVR, but, traditionally, likely effects on skin pigmentation (Allentoft et al., 2015; Hanel &
wear few clothes and eat diets rich in fish. These are also mostly Carlberg, 2020a; Ju & Mathieson, 2020; Røyrvik et al., 2018). Here,
coastal and island populations that have also been relatively isolated I reiterate salient points of those reviews and other works and sup-
throughout most of history because of lack of accessibility over plement them sparingly with some additional insights.
land. Thus, they were not subject to high levels of admixture from Inferences about the likely evolutionary factors affecting human
mainland groups. These populations are exemplars of one of the skin pigmentation among people living at high latitudes in Eurasia

F I G U R E 3 Generalized timeline of
main events in the evolution of skin
pigmentation in Coastal South and
Southeast Asia, Melanesia, and Australia.
(a) Beginning around 70 kya, small
populations of coastal dispersers make
their way along the coasts of southern
and southeastern Asia. Introgression from
Denisovans occurs about 43 kya, possibly
involving introduction of variants favoring
darker pigmentation. (b) In Melanesia
and Australia, following the divergence
from mainland coastal groups, continued
positive selection on variants of DDB1,
MFSD12, and other loci promotes in
the absence of purifying selection on
MC1R produces dark and very dark skin
pigmentation phenotypes
JABLONSKI | 717

have been mooted in the literature for nearly a century (Loomis, 1967; pressure for depigmentation. A further consideration is that in-
Murray, 1934). Emphasis on loss of pigmentation in order to promote creased thickening of the stratum corneum in response to repeated
cutaneous vitamin D photosynthesis at high latitudes has been em- outdoor conditions and UVR (Oh et al., 2004) would also have
phasized, but so too has the importance of consumption of vitamin served to attenuate any possible cutaneous vitamin D production
D-­rich foods in regions where cutaneous biosynthesis is insufficient under these conditions (Young et al., 2020).
to fulfill year-­round needs for the vitamin (Chaplin & Jablonski, 2013; In far northwestern Europe, the extreme physical isolation of
Jablonski & Chaplin, 2000, 2010). Elucidation of the genetic basis for people in northernmost Britain at the end of the Pleistocene and
depigmentation began with the landmark study in which a zebrafish early Holocene favored maximal skin depigmentation and a near ab-
model was used to demonstrate the action of the SLC24A5 variant sence of eumelanin in the skin. This was made possible by positive
common to northwestern European people (Lamason et al., 2005). selection for the classic depigmentation variant of SLC24A5, and by
The absence of the same variant in East Asian peoples living at sim- relaxation of selection pressure on the MC1R locus, resulting in high
ilar latitudes (Lamason et al., 2005; Norton et al., 2007) inaugurated levels of polymorphism at the locus (Harding et al., 2000; Latreille
an intensive search for the genes responsible for depigmentation in et al., 2009; Rees, 2000). Under these circumstances, cutaneous
East Asians. Subsequent historical genomic studies of modern and production of vitamin D could be maximized during the few months
ancient populations have shown that depigmentation in Europeans, when photocatalytic UVB wavelengths were present in the sunlight
especially among far northern-­dwelling peoples, occurred in a step-­ (Chaplin & Jablonski, 2013; Jablonski & Chaplin, 2000, 2010). During
wise fashion, beginning with a shared variant of the Kit ligand (KITLG) the millennia in prehistoric and early historic times when people did
gene in the common ancestor of western European and east Asian not travel widely or migrate to other climes, this highly depigmented
people (Hanel & Carlberg, 2020a; Lao et al., 2007; Sulem et al., 2007). skin phenotype worked well. Because their UVR regimes were mark-
Extreme depigmentation in northwestern Europeans involved mul- edly different from those of people in the tropics, they experienced
tiple skin pigmentation genes and appears to have been further pro- only short periods of UVB during the height of the summer and sun-
moted by the introduction of agriculture (Brace et al., 2019). At the burns would have been rare (Chaplin & Jablonski, 2013; Jablonski
highest European latitudes, including Scotland, genetic variants con- & Chaplin, 2010). Outdoor lifestyles meant that people adapted
tributing to mostly depigmented or lightly pigmented skin comprise to gradually changing durations and intensities of UVR—­
mostly
one component of the biocultural compromise necessary for survival UVA—­exposure throughout the year, to which their skin adapted
and reproductive success under low and markedly seasonal UVB at primarily by thickening of the stratum corneum; the fact that they
extreme high latitudes (Chaplin & Jablonski, 2013). Modifications could develop only negligible protection from melanin production
of genes affecting the production and metabolism of vitamin D did not matter because the challenge from UVB was not severe
constitute another part of this compromise, in some cases offset- or prolonged (Bech-­Thomsen & Wulf, 1995; Hennessy, Oh, Rees,
ting extreme depigmentation in high-­latitude Eurasian populations et al., 2005; Sheehan et al., 1998). Under these conditions, mostly
(Hanel & Carlberg, 2020a). The last component of the high-­latitude depigmented skin was not a liability because it did not detract from
biocultural compromise is a vitamin D-­rich diet, focused on oily fish health and well-­being and did not affect reproductive success. This
and, depending on the location, also including marine mammals and situation changed markedly when people from northwestern and
wild or domesticated reindeer (Chaplin & Jablonski, 2013; de Barros northern Europe began to travel and migrate to sunnier places in
Damgaard et al., 2018; Ross et al., 2006). The fragility of the bio- colonial times, and experience high episodic or sustained loads of
cultural compromise at high latitudes in Eurasia is illustrated by the UVR (Jablonski, 2012). Historical migrations of people from north-
high prevalence of chronic lifestyle diseases, including cardiovascu- ern Europe to far southern Europe, Africa, Australia, and tropical lat-
lar disease and metabolic syndrome, in populations that mostly or itudes of the Americas in the last 200 years, along with the increased
completely abandon vitamin D-­rich diets (Chaplin & Jablonski, 2013; popularity of holiday travel, caused dramatic shifts in the nature of
Ross et al., 2006). UVR exposure, and increased prevalence of all skin cancers (Latreille
The evolutionary trend toward depigmented skin in high-­latitude et al., 2009; Rees, 2003; Sturm et al., 2003).
Eurasian populations involved numerous genes affecting skin pig- The evolution of skin lightening in northeastern Asia oc-
mentation and vitamin D metabolism and did not result in uniform curred under similar, but not identical, conditions of generally low
adaptive strategies across the great expanse of northern Eurasia fa- and seasonal UVB as those of northwestern Europe (Chaplin &
voring extreme depigmentation (Hanel & Carlberg, 2020a, 2020b). Jablonski, 2013; Jablonski & Chaplin, 2000). Among the most inter-
Rather, what we observe is that skin pigmentation has evolved as esting facts to emerge in the study of the evolution of human skin pig-
part of a larger genetic and cultural package favoring enhanced mentation evolution is that different suites of genes have influenced
availability and economical utilization of dietary and cutaneously depigmentation in northern Europeans and East Asians. Recent re-
synthesized vitamin D. Because habitation of far-­northern latitudes search on the function of the upstream region of KITLG shared by the
in Eurasia was predicated on the development of material culture, common ancestor of all Eurasians shows that the region was under
including sewn clothing, which afforded protection against extreme stronger selective pressure in East Asians, possibly because it con-
cold and wind, the surface area of skin available for cutaneous vita- ferred depigmentation along with greater resistance to cold (Yang
min D production was reduced, thereby intensifying the selective et al., 2018). This appears to have complemented selection on an
718 | JABLONSKI

OCA2 (oculocutaneous albinism, type 2) variant absent in Europeans this context, the much-­discussed dark or “dark to black” skin re-
(Yang et al., 2016) and a variant of MFSD12 (Adhikari et al., 2019) to constructed for “Cheddar Man,” a Mesolithic inhabitant of Cheddar
create a comparable level of depigmentation conducive to cutane- Gorge in Somerset, England (Brace et al., 2019), becomes more
ous vitamin D production. Thus, depigmentation in the ancestors of clearly understandable and less sensational. Traditional hunter-­
modern western Europeans and East Asians provides an excellent of gatherer diets centered around coastal marine sources, hunted
convergent evolution (Norton et al., 2007; Yang et al., 2016). Natural terrestrial game, and foraged plant foods are rich in vitamin D and
selection has acted upon on different genes and gene variants in re- make it possible for people to continue to live healthy reproductive
sponse to comparable environmental forces to produce comparable lives while maintaining “dark” and highly tannable skin. Skin depig-
physiological solutions to a problem affecting health and reproduc- mentation in northwestern Europe occurred gradually, over many
tive success. millennia, with the introduction of the classic SLC24A5 variant being
East Asians and western Europeans share visibly similar “light” the last and most dramatic step in the process after the introduction
skin, but their responses to UVR are different. The skin of East of agriculture.
Asians has a higher density of melanosomes and produces more eu-
melanin and pheomelanin in response to UVR exposure than does
the skin or western Europeans (Hennessy, Oh, Diffey, et al., 2005; 3.3.1 | Beringia and the Americas
Hennessy, Oh, Rees, et al., 2005). Thus, most East Asian people can
tan, whereas many Europeans, especially northern Europeans, can Discussion of the evolution of human skin pigmentation among in-
tan only slightly or not at all. The development and persistence of digenous populations of the Americas follows naturally from that
tanning abilities are conferred by many genes operating on mela- of northern and northeast Asians because northern Asia was the
nin production, distribution, and breakdown in the skin (Del Bino area of origin for the earliest peoples entering Beringia and North
et al., 2018) and tanning abilities, like depigmentation, have evolved America via coastal and overland routes (Goebel et al., 2008;
multiple times independently in human history under conditions of Moreno-­Mayar et al., 2018; Nielsen et al., 2017). Considerable re-
seasonally strong UVR (Martínez-­C adenas et al., 2013; Quillen, 2015; search and debate over the number, nature, timing, and routing of
Quillen et al., 2019). The moderate tanning abilities among some these dispersals over the last century have resulted in general con-
Scandinavians and northern Europeans appear to have been con- sensus that the first humans to disperse into the Americas trave-
ferred by recent genetic admixture across northern Eurasia (Hanel led via a coastal route beginning around 15 kya, having genetically
& Carlberg, 2020a), but other factors, including altered epistatic in- diverged from ancient north Asians beginning around 36 kya with
teractions among pigmentation genes, may have also contributed. gene flow continuing until about 25 kya (Moreno-­Mayar et al., 2018;
Following the amelioration of climatic conditions at the end Nielsen et al., 2017). Details of the population movements are be-
of the Pleistocene, the development of extensive steppe grass- yond the scope of this paper, but a key feature of this dispersal event
lands across much of hinterland Eurasia created a theater for rapid was that it involved a prolonged “Beringian standstill” lasting about
human population growth associated with animal husbandry, and 15,000 years (Moreno-­Mayar et al., 2018), during which an isolated
relatively rapid, bidirectional east–­
west movements of people population lived under conditions of limited and highly seasonal sun-
(Allentoft et al., 2015; Damgaard et al., 2018). The results of ge- light, including virtually no UVB (Hlusko et al., 2018). In the absence
nomic, linguistic, and archeological studies show a gradual transi- of UVB and cutaneous biosynthesis of vitamin D, the consumption
tion from Bronze Age pastoralists of West Eurasian ancestry toward of vitamin D-­rich foods was necessary for survival and reproduc-
horse-­mounted warrior peoples of increased East Asian ancestry tive success. Because normal human development in utero and in
(Damgaard et al., 2018). The rise and expansion of agriculture from early postnatal months depends on maternal vitamin D stores, the
Anatolia and Iran saw incursions of agriculturalists into southern evolution of efficient transmission of vitamin D in breastmilk was
and central Europe, increasingly displacing earlier hunter-­gatherer critical for infant survival. This need appears to have been met by
populations (Skoglund & Mathieson, 2018). The evolution of skin positive selection on the human-­
specific EDAR (Ectodysplasin A
pigmentation genes and phenotypes of the region reflects these receptor) variant V370A associated with mammary ductal branch-
dramatic movements, and the influence of relatively few genes of ing, which appears to have amplified the transfer of vitamin D to
major effect transmitted over long distances by admixture (Ju & infants via mothers’ milk (Hlusko et al., 2018). Evidence for the rapid
Mathieson, 2020). Of these, the effect of the classic depigmentation spread and positive selection of this variant back into northern Asia
variant of SLC24A5 is greatest, with recent evidence showing that and Scandinavia (Mathieson et al., 2015) suggests that it conferred
this variant reached western Europe by admixture from Anatolian an evolutionary advantage to people inhabiting regions in which the
agriculturalists followed by continued positive selection (Ju & potential for cutaneous production of vitamin D from UVB was se-
Mathieson, 2020). Recent genomic evidence indicates that some verely limited and the vitamin could be derived only from dietary
of the hunter-­gatherer populations of Mesolithic western Europe sources. This example illustrates the importance of understanding
and Britain had moved along the coast from the Iberian Peninsula the evolution of skin pigmentation in the broader context of human
into northern France, and thence up the Atlantic coast, making evolution and dispersals, and recognizing the nature of life events
use of marine food resources as they went (Brace et al., 2019). In most likely to be influenced by natural selection.
JABLONSKI | 719

Skin pigmentation among Inuit people is often discussed, but has indigenous Americans (Quillen et al., 2012, 2019). One of the most
been inadequately measured or genetically characterized. Ancestral biologically significant facts about the evolution of skin pigmentation
Inuit traversed the Bering Land Bridge in the mid-­Holocene, dispers- in the Americas is that it has involved selection for genes that confer
ing into Americas from northern Siberia about 10,000 years after enhanced tanning (facultative pigmentation) rather than darker con-
first coastal dispersal about 15 kya (Nielsen et al., 2017; Skoglund & stitutive pigmentation. Most indigenous Americans, including those
Mathieson, 2018). Their constitutive pigmentation is light to mod- that inhabit areas of high and relatively nonseasonal UVR, exhibit
erate, but they have remarkable tanning abilities, as illustrated by “moderate” skin reflectance values for unexposed skin, but strong
the “tan lines” evident on traditional hunters when their animal skin tanning abilities (Jablonski & Chaplin, 2000; Quillen et al., 2019). The
parkas are removed (Jablonski, 2012). This is another case of bio- genetic basis of these abilities is almost completely unknown and
cultural compromise involving skin pigmentation, dietary vitamin warrants considerable study (Quillen et al., 2019). Looking in general
D, and regulation of vitamin D availability at critical points in the at the evolution of skin pigmentation in the Americas, it is important
lifespan. Inuit and their ancestral populations were, traditionally, to view the range of factors that have contributed to the phenom-
hunter-­
gatherers who, depending on location, subsisted primar- enon: population bottlenecks limiting the pool of potential variant
ily on marine mammals, oily fish, and caribou. Living near or above pigmentation genes, a relatively short history of human habitation,
the Arctic Circle, they experienced markedly seasonal patterns of and the fact that the Late Pleistocene populations dispersing into the
daylight and UVR exposure, and negligible UVB; they did, however, Americas brought with them many means to buffer themselves and
experience seasonally strong direct and reflected UVA (Chaplin & their skin against the exigencies of the environment and seasonal
Jablonski, 2013; Jablonski, 2012; Jablonski & Chaplin, 2000, 2010). changes in food availability through sewn clothing, constructed shel-
Modest cutaneous production of vitamin D occurs in some popula- ters, and food storage technologies (Jablonski, 2012).
tions, but the physiological significance of this is unclear (Andersen
et al., 2012), In the presence of a traditional diet replete in vitamin
D, and a physiological mechanism to ensure ample vitamin D to rap- 3.3.2 | Indian subcontinent
idly developing neonates (Hlusko et al., 2018), there was little se-
lective pressure for further depigmentation; if anything, there may The evolution of skin pigmentation on the Indian subcontinent can
have been selective pressure to enhance facultative pigmentation. only be understood in light of the complex demographic and social
The fragility of the biocultural compromise here is witnessed by the histories of the region. Because of this complexity and the practical
high levels of vitamin D deficiency and high prevalence of rickets and difficulties of studying living populations on the subcontinent and
metabolic syndrome in people no longer eating traditional diets year obtaining biological samples therefrom, studies of skin pigmentation
round (Andersen et al., 2013; El Hayek et al., 2010). diversity and evolution are still in their infancy. The history of dis-
The evolution of skin pigmentation in the Americas has been persals and genetic admixture within the subcontinent is complex
studied relatively little, with the exception of a few pioneering stud- and still poorly understood (Dennell & Petraglia, 2012; Majumder &
ies in the last decade (Adhikari et al., 2019; Quillen et al., 2012). The Basu, 2014; Moorjani et al., 2013). Genetic evidence attests that the
peopling of the Americas from people of northern and northeastern southern coast of the subcontinent received H. sapiens populations
Asian ancestry at the end of the Pleistocene and during the Holocene early in the history of dispersal along the coast of the Afro-­Arabian
meant that indigenous populations carried novel mixtures of skin Peninsula (Majumder & Basu, 2014). Since then, populations dispers-
pigmentation genes, which reflect histories of extreme natural se- ing into the subcontinent have followed mostly hinterland routes,
lection and population bottlenecks. Study of the evolution of skin from central and western Asia, primarily via a northwestern corridor
pigmentation in the pre-­Columbian Americas has been made difficult (Damgaard et al., 2018; Majumder & Basu, 2014), southward. Tracing
by the facts that many indigenous populations have not been stud- the fate and routes of these populations has been difficult. The sub-
ied comprehensively for both their skin pigmentation phenotypes continent contains people hailing from four distinct language groups
and genetics, and because most of the populations that have been who live in mostly non-­overlapping regions and who reached their
studied exhibit evidence of considerable admixture from colonial current locations at different times in late prehistory and early his-
European and enslaved sub-­Saharan African populations (Quillen tory (Majumder & Basu, 2014). Some populations belong to groups
et al., 2019). In this context, studies in which the effects of genetic designated as tribal, while most belong to caste societies. One of the
admixture have been “dissected” to reveal novel skin pigmentation most distinctive aspects of population structure in the subcontinent
genes have been of key importance (Bonilla et al., 2005; Quillen is the pattern of endogamy within groups, including within the hierar-
et al., 2012). Because of the northern and northeastern Asian origin chically arranged Hindu castes (Majumder & Basu, 2014). Because of
of New World populations, the variants of OCA2 and MFSD12 found the complex pattern of dispersal and the prevalence of within-­group
in Asian populations are also present among indigenous Americans marriage, the subcontinent is more accurately described as a genetic
(Adhikari et al., 2019; Yang et al., 2016). In addition, other genes not mosaic and, not surprisingly, social factors and population struc-
previously implicated in skin pigmentation, including EGFR (epider- ture have played a stronger role than natural selection in shaping
mal growth factor receptor) and OPRM1 (opioid receptor, mu-­1), skin color diversity across the region (Iliescu et al., 2018; Stokowski
have emerged as important contributors to skin pigmentation among et al., 2007). Two striking findings have emerged from studies of skin
720 | JABLONSKI

F I G U R E 4 Generalized timeline of
main events in the evolution of skin
pigmentation in hinterland Eurasia,
Beringia, and the Americas. (a) In central
Eurasia, beginning around 70 kya,
moderately pigmented and tannable
phenotypes are favored under continued
relaxation of selection on MC1R and
positive selection of a KITLG producing
lighter skin. After the introduction of
agriculture and animal husbandry around
10 kya, increasing population admixture
with diverse skin pigmentation gene
assemblages producing moderately
pigmented, tannable skin. (b) In
northern and northwestern Europe, the
introduction of the SLC24A5 variant
from Anatolian agriculturalists around
6 kya is followed by positive selection
on the variant; this is favored strongly
in relatively isolated populations living
under the low and seasonal UVB
conditions of northernmost Europe.
(c) In East Asia, intensified selection
on the KITLG variant and other loci
produces further depigmentation. (d)
In Beringia, hunter-­gather populations
experiencing the “Beringian standstill”
from about 25–­15 kya undergo selection
for the EDAR variant favoring enhanced
mammary transmission of vitamin D, while
maintaining or enhancing their tanning
abilities through skin pigmentation genes
not yet identified. (e) In the Americas,
technologically sophisticated hunter-­
gathers disperse southward, carrying
genes capable of producing strong
tans under high UVB conditions. Dark
constitutive pigmentation does not
evolve, probably because of the absence
of appropriate gene variants and the many
cultural buffering mechanisms—­including
food storage—­used by dispersing
populations to mitigate the effects
of different regimes of UVB and food
availability
JABLONSKI | 721

pigmentation diversity in the subcontinent to date. The first is that Chaplin, 2000). Satisfactory tests of this hypothesis in populations
while the range of skin pigmentation phenotypes across the breadth with varying levels of skin color sexual dimorphism have not yet been
and length of the subcontinent is vast (Jablonski & Chaplin, 2000), undertaken. What seems likely is that sexual selection has prob-
the pattern is not strongly correlated to strength of UVR (Iliescu ably played a secondary and complementary role in the evolution
et al., 2018). The second is that the classic skin lightening variant of of skin pigmentation, primarily through manipulating levels of exist-
SLC24A5 is present in high frequencies in most populations but is ing sexual dimorphism (Jablonski, 2012; Jablonski & Chaplin, 2000;
not expressed (Basu Mallick et al., 2013; Iliescu et al., 2018). Thus, Quillen et al., 2019). This apparent directional preference for women
epistasis has affected the expression of the SLC24A5 in such a way as with lighter skin appears to be particularly strongly marked in Island
to prevent the allele's ability to lower eumelanin production (Quillen Melanesia, Japan, and among certain populations in India where
et al., 2019). The reasons behind this can be traced potentially to cultural preferences for lighter females are strongly culturally rein-
the specific and distinct genetic backgrounds, including social selec- forced (Iliescu et al., 2018; Jablonski, 2012; Norton et al., 2006). It
tion and sexual selection, that contributed to the unique reproduc- would be surprising if uniform, global patterns of sexual selection
tive structure of the Indian populations, but the relative importance applied because these would rely on uniform preferences (Quillen
of these phenomena is not clear at present. This result should be et al., 2019). Human skin pigmentation is the product of many cul-
seen as a useful cautionary tale for forensic reconstruction of skin tural forces, including culturally determined mating preferences
pigmentation, viz., specific genotypes are not associated reliably and sexual selection, which have had potent local effects modify-
with specific skin color phenotypes (Iliescu et al., 2018; Quillen ing the influence of UVR-­determined directional selection (Quillen
et al., 2019). et al., 2019).
A hypothetical skin pigmentation timeline for the major regions
of Eurasia and the Americas is presented in Figure 4.
5 | CO N C LU S I O N S

4 | S E X UA L D I M O R PH I S M A N D The evolution of human skin pigmentation has never been a sim-


S E X UA L S E LEC TI O N I N H U M A N S K I N ple deterministic process, but it has been influenced potently by
PI G M E NTATI O N environmental UVR and its effects on vitamin availability. Humans
evolved under the sun. The early evolution of the genus Homo oc-
One of the most interesting and still not fully understood aspects curred in equatorial Africa under conditions of high UVA and high
of human skin pigmentation evolution is the consistent pattern of UVB. Under these conditions, natural selection favored the evolu-
sexual dimorphism observed: In most indigenous groups whose tion of enhanced sweating abilities and enhanced permanent eu-
skin phenotypes have been quantified using reflectometry, female melanin pigmentation, concomitant with the loss of most body hair.
pigmentation is lighter than male, sometimes to a striking degree Enhanced eumelanin constitutive pigmentation provided protection
(Byard, 1981; Jablonski & Chaplin, 2000; Roberts & Kahlon, 1972). against folate degradation and damage to DNA while still permit-
This aspect of human skin pigmentation garnered early attention ting cutaneous vitamin D production. Dark skin in early H. sapiens
from natural historians including Charles Darwin and has been the was made possible by elimination of variation at the MC1R locus and
focus of considerable attention since. Darwin believed that differ- positive selection at the MFSD12, DDB1, and probably other loci.
ences in human skin color between “races” were caused, not by Populations of H. sapiens in equatorial Africa underwent further posi-
natural selection, but by sexual selection; in other words, that skin tive selection for very dark, eumelanin-­rich pigmentation. Habitation
color had been primarily and systematically influenced by deliber- of southern Africa, with its somewhat lower and more seasonal UVR
ate choice of mates (Aoki, 2002; Darwin, 1871). The belief that skin regimes, involved relaxation of purifying selection on MC1R. Further
color is an important determinant of human mate choice became depigmentation in southern African hunter-­gatherers occurred
fixed in the literature and public imagination in the late 20th cen- under the influence of the classic variant of SLC24A5, which en-
tury (Aoki, 2002; Diamond, 1991; Robins, 1991; Van den Berghe & tered Africa through the Afro-­Arabian Peninsula and rapidly spread
Frost, 1986). Further study has revealed, however, that sexual selec- southward. Dispersal of H. sapiens into Eurasia began around 70 kya
tion is not as dominant and omnipresent force as some have con- and involved both rapid coastal and hinterland routes. Dispersing
ceived (Jablonski & Chaplin, 2000; Madrigal & Kelly, 2007). Sexual populations were small, and population bottlenecks affected the na-
dimorphism in human skin pigmentation is real, but its expression ture of pigmentation gene variants they carried. Equipped with so-
varies greatly, with some populations exhibiting marked levels and phisticated food procurement technologies, but not sewn clothing,
others very little (Jablonski & Chaplin, 2000; Norton et al., 2006). agriculture or animal husbandry, people modified their diets and life-
The primary cause of the difference between the sexes may be the styles according to local resources and conditions. Coastal dispersal
importance of increased cutaneous vitamin D production poten- routes in South and Southeast Asia and Europe afforded people the
tial, and therefore lighter skin, in the skin of reproductively aged opportunity to easily harvest fish and other vitamin D-­rich foods,
women, in order to facilitate absorption and redistribution of dietary mitigating selective pressure for depigmentation to facilitate cutane-
calcium to the developing fetus and nursing neonate (Jablonski & ous production of vitamin D. Further darkening of skin pigmentation
722 | JABLONSKI

occurred, involving intensified selection on MFSD12, DDB1, and the contributions of systematic cultural or sexual selection to pro-
probably some variants carried by introgressing Denisovans, in ducing directional evolution of skin pigmentation in recent history
predominantly coastal populations dispersing into equatorial island also requires further, thoughtful study, but such work must be pur-
Southeast Asia, Melanesia, and Australia. Habitation of non-­tropical sued with careful attention to potential investigator bias, based on
hinterland Eurasia with more highly seasonal UVB regimes saw the individual or cultural preoccupations. The evolution of human skin
evolution of lighter constitutive pigmentation under the influence pigmentation has been a complex and contingent process for hun-
of positive selection for a variant of KITLG, which was then shared dreds of thousands of years. This should not discourage us, because
by populations dispersing into northern and western Europe, and we now have the tools to understand the many genetic and cultural
northeastern Asia. Trends toward loss of constitutive eumelanin pig- factors that have contributed to it.
mentation and variable tanning abilities occurred at different times
in these northward diverging populations and involved different AC K N OW L E D G E M E N T S
suites of pigmentation genes. Prior to the introduction of agriculture I am grateful to the editors of the special issue of Pigment Cell and
and the origin of the classic variant of SLC24A5, hinterland Eurasians Melanoma Research for their invitation to contribute. Shosuke Ito is
were moderately pigmented, had some or considerable tanning abili- thanked especially for supervising the review of the manuscript. The
ties, probably conferred by varying configurations of pigmentation ideas in this paper were developed over the course of many years,
genes, and pursued outdoor hunter-­gatherer lifestyles. The evolu- and I am grateful especially to George Chaplin for lengthy and prob-
tion of extremely depigmented skin was a recent novelty in human ing discussions of all of these over the last 25 years. In recent years,
evolution, occurring in northern and northwestern Europe after the I have benefited from clarifying discussions with Roger Bouillon, Tim
introduction of the SLC25A5 variant as the result of admixture with Caro, Keith Cheng, Brenna Henn, Michael Holick, Mircea Iliescu,
agriculturalists. Skin containing little eumelanin and lacking tanning Larry Kenney, Beatriz Marcheco-­Teruel, Heather Norton, Richard
abilities evolved only in relatively isolated populations living under Prum, Ellen Quillen, Mark Shriver, Rick Sturm, Mark Thomas, Sarah
the lowest and most highly seasonal UVB regimes with very limited Tishkoff, Richard Weller, Tony Wolf, and Antony Young. I thank my
opportunities for cutaneous vitamin D synthesis. The extremes of research assistant, Tess Wilson, for her continuing excellence and
human skin pigmentation evolved under extremes of environmen- support in maintaining bibliographic materials, producing figures,
tal UVB and were mitigated by vitamin D-­rich diets. When genetic and providing all-­around logistical support for my research.
variants were available that enhanced reproductive success, they
underwent rapid positive selection, especially in extreme solar en- ORCID
vironments, and quickly brought about changes in skin pigmentation Nina G. Jablonski https://orcid.org/0000-0001-7644-874X
phenotypes. Thus, natural selection for dark pigmentation under
high UVR conditions and for lighter skin capable of tanning under REFERENCES
lower and more seasonal UVR has been the dominant influence, but Abdel-­Malek, Z. (2009). Constitutive pigmentation, melanogenesis, and
skin pigmentation has been modified increasingly by population ge- the UV response. Pigment Cell & Melanoma Research, 22(3), 344.
https://doi.org/10.1111/j.1755-­148X.2009.00569_2.x
netic influences on the genetic composition of dispersing popula-
Adhikari, K., Mendoza-­Revilla, J., Sohail, A., Fuentes-­Guajardo, M.,
tions and by cultural processes, which have mitigated the expression Lampert, J., Chacón-­Duque, J. C., Hurtado, M., Villegas, V., Granja,
of pigmentation genes and modified the impact of the environment V., Acuña-­Alonzo, V., Jaramillo, C., Arias, W., Lozano, R. B., Everardo,
on the human body. In recent millennia, human skin pigmentation P., Gómez-­Valdés, J., Villamil-­Ramírez, H., Silva de Cerqueira, C. C.,
Hunemeier, T., Ramallo, V., … Ruiz-­Linares, A. (2019). A GWAS in
has been influenced increasingly by culturally reinforced processes
Latin Americans highlights the convergent evolution of lighter skin
including, in some places, mating practices and sexual selection. pigmentation in Eurasia. Nature Communications, 10(1), 358. https://
There remains a pressing need for more transdisciplinary re- doi.org/10.1038/s4146​7-­018-­0 8147​- ­0
search on the evolution of human pigmentation diversity. These Alaluf, S., Atkins, D., Barrett, K., Blount, M., Carter, N., & Heath, A.
(2002). Ethnic variation in melanin content and composition in pho-
need to include functional genomic studies aimed at further eluci-
toexposed and photoprotected human skin. Pigment Cell Research,
dating the mechanisms of tanning, as well as studies aimed at un- 15(2), 112–­118. https://doi.org/10.1034/j.1600-­0749.2002.1o071.x
derstanding the effects of genetic admixture and epistasis, and Alaluf, S., Heath, A., Carter, N., Atkins, D., Mahalingam, H., Barrett, K.,
epigenetic influences on melanin production. We also need to bet- Kolb, R., & Smit, N. (2001). Variation in melanin content and com-
ter understand how various combinations of melanin pigments and position in Type V and VI photoexposed and photoprotected human
skin: The dominant role of DHI. Pigment Cell Research, 14(5), 337–­3 47.
the physical packaging of melanosomes contribute to the subtly
https://doi.org/10.1034/j.1600-­0749.2001.140505.x
different colors observed in human skin. It has long been appreci- Alexander, L. M., Kutz, J. L., & Kenney, W. L. (2013). Tetrahydrobiopterin
ated that different forms of melanin in the skin—­DHICA-­eumelanin, increases NO-­ dependent vasodilation in hypercholesterolemic
DHI-­eumelanin, and pheomelanin—­contribute to the chromatic vari- human skin through eNOS-­coupling mechanisms. American Journal of
Physiology -­Regulatory, Integrative and Comparative Physiology, 304(2),
ation in skin color phenotypes (Alaluf et al., ,2001, 2002), but much
R164–­R169. https://doi.org/10.1152/ajpre​gu.00448.2012
still remains to be clarified about how the different melanin types Alfredsson, L., Armstrong, B. K., Butterfield, D. A., Chowdhury, R., de
and, possibly also, the different sizes and physical arrangements of Gruijl, F. R., Feelisch, M., Garland, C. F., Hart, P. H., Hoel, D. G.,
melanosomes impart subtly different colors to skin. Lastly, study of Jacobsen, R., Lindqvist, P. G., Llewellyn, D. J., Tiemeier, H., Weller, R.
JABLONSKI | 723

B., & Young, A. R. (2020). Insufficient sun exposure has become a real status in women of childbearing age. Journal of Photochemistry and
public health problem. International Journal of Environmental Research Photobiology B: Biology, 131(5), 90–­ 95. https://doi.org/10.1016/j.
and Public Health, 17(14), 5014. https://doi.org/10.3390/ijerp​h1714​ jphot​obiol.2014.01.002
5014 Borradale, D. C., & Kimlin, M. G. (2012). Folate degradation due to
Allentoft, M. E., Sikora, M., Sjögren, K.-­G ., Rasmussen, S., Rasmussen, M., ultraviolet radiation: Possible implications for human health
Stenderup, J., Damgaard, P. B., Schroeder, H., Ahlström, T., Vinner, L., and nutrition. Nutrition Reviews, 70(7), 414–­ 422. https://doi.
Malaspinas, A.-­S., Margaryan, A., Higham, T., Chivall, D., Lynnerup, N., org/10.1111/j.1753-­4887.2012.00485.x
Harvig, L., Baron, J., Casa, P. D., Dąbrowski, P., … Willerslev, E. (2015). Bower, C., & Stanley, F. J. (1989). Dietary folate as a risk factor for
Population genomics of Bronze Age Eurasia. Nature, 522(7555), 167–­ neural-­tube defects: Evidence from a case-­control study in Western
172. https://doi.org/10.1038/natur​e14507 Australia. The Medical Journal of Australia, 150(11), 613–­619. https://
Ambrose, S. H. (1998). Late Pleistocene human population bottlenecks, doi.org/10.5694/j.1326-­5377.1989.tb136​723.x
volcanic winter, and differentiation of modern humans. Journal Brace, S., Diekmann, Y., Booth, T. J., van Dorp, L., Faltyskova, Z., Rohland,
of Human Evolution, 34(6), 623–­ 651. https://doi.org/10.1006/ N., Mallick, S., Olalde, I., Ferry, M., Michel, M., Oppenheimer, J.,
jhev.1998.0219 Broomandkhoshbacht, N., Stewardson, K., Martiniano, R., Walsh,
Andersen, S., Jakobsen, A., & Laurberg, P. (2012). Vitamin D status in S., Kayser, M., Charlton, S., Hellenthal, G., Armit, I., … Barnes, I.
North Greenland is influenced by diet and season: Indicators of (2019). Ancient genomes indicate population replacement in Early
dermal 25-­hydroxy vitamin D production north of the Arctic Circle. Neolithic Britain. Nature Ecology & Evolution, 3(5), 765–­771. https://
British Journal of Nutrition, FirstView, 110(1), 50–­ 57. https://doi. doi.org/10.1038/s4155​9-­019-­0 871-­9
org/10.1017/S0007​11451​2004709 Branda, R. F., & Blickensderfer, D. B. (1993). Folate deficiency increases
Andersen, S., Laurberg, P., Hvingel, B., Kleinschmidt, K., Heickendorff, L., genetic damage caused by alkylating agents and y-­ irradiation in
& Mosekilde, L. (2013). Vitamin D status in Greenland is influenced Chinese hamster ovary cells. Cancer Research, 53(22), 5401–­5408.
by diet and ethnicity: A population-­based survey in an Arctic society Branda, R. F., & Eaton, J. W. (1978). Skin color and nutrient photolysis: An
in transition. The British Journal of Nutrition, 109(5), 928–­935. https:// evolutionary hypothesis. Science, 201(4356), 625–­626. https://doi.
doi.org/10.1017/s0007​11451​2002097 org/10.1126/scien​ce.675247
Aoki, K. (2002). Sexual selection as a cause of human skin colour varia- Bruls, W. A. G., Slaper, H., Van Der Leun, J. C., & Berrens, L. (1984).
tion: Darwin's hypothesis revisited. Annals of Human Biology, 29(6), Transmission of human epidermis and stratum corneum as a
589–­608. https://doi.org/10.1080/03014​46021​0 0001​9144 function of thickness in the ultraviolet and visible wavelengths.
Basu Mallick, C., Iliescu, F. M., Möls, M., Hill, S., Tamang, R., Chaubey, G., Photochemistry and Photobiology, 40(4), 485–­494. https://doi.
Goto, R., Ho, S. Y. W., Gallego Romero, I., Crivellaro, F., Hudjashov, G., org/10.1111/j.1751-­1097.1984.tb046​22.x
Rai, N., Metspalu, M., Mascie-­Taylor, C. G. N., Pitchappan, R., Singh, Buccimazza, S. S., Molteno, C. D., Dunne, T. T., & Viljoen, D. L. (1994).
L., Mirazon-­L ahr, M., Thangaraj, K., Villems, R., & Kivisild, T. (2013). Prevalence of neural tube defects in Cape Town. South Africa.
The light skin allele of SLC24A5 in South Asians and Europeans shares Teratology, 50(3), 194–­ 199. https://doi.org/10.1002/tera.14205​
identity by descent. PLoS Genetics, 9(11), e1003912. https://doi. 00304
org/10.1371/journ​al.pgen.1003912 Bustamante, M., Hernandez-­Ferrer, C., Tewari, A., Sarria, Y., Harrison, G.
Battie, C., Jitsukawa, S., Bernerd, F., Del Bino, S., Marionnet, C., & I., Puigdecanet, E., Nonell, L., Kang, W., Friedländer, M. R., Estivill,
Verschoore, M. (2014). New insights in photoaging, UVA induced X., González, J. R., Nieuwenhuijsen, M., & Young, A. R. (2020). Dose
damage and skin types. Experimental Dermatology, 23(S1), 7–­12. and time effects of solar simulated ultraviolet radiation on the in
https://doi.org/10.1111/exd.12388 vivo human skin transcriptome. British Journal of Dermatology, 182(6),
Bech-­Thomsen, N., & Wulf, H. C. (1995). Photoprotection due to pigmen- 1458–­1468. https://doi.org/10.1111/bjd.18527
tation and epidermal thickness after repeated exposure to ultravio- Byard, P. J. (1981). Quantitative genetics of human skin color. Yearbook
let light and psoralen plus ultraviolet A therapy. Photodermatology, of Physical Anthropology, 24(S2), 123–­137. https://doi.org/10.1002/
Photoimmunology & Photomedicine, 11(5–­6), 213–­218. https://doi. ajpa.13302​4 0506
org/10.1111/j.1600-­0781.1995.tb001​72.x Caccamo, D., Ricca, S., Currò, M., & Ientile, R. (2018). Health risks of hy-
Beleza, S., Johnson, N. A., Candille, S. I., Absher, D. M., Coram, M. A., povitaminosis D: A review of new molecular insights. International
Lopes, J., Campos, J., Araújo, I. I., Anderson, T. M., Vilhjálmsson, Journal of Molecular Sciences, 19(3), 892. https://doi.org/10.3390/
B. J., Nordborg, M., Silva, A. C., Shriver, M. D., Rocha, J., Barsh, G. ijms1​9030892
S., & Tang, H. (2013). Genetic architecture of skin and eye color in Carlberg, C. (2019). Vitamin D: A micronutrient regulating genes.
and African-­European admixed population. PLoS Genetics, 9(3), Current Pharmaceutical Design, 25(15), 1740–­ 1746. https://doi.
e1003372. https://doi.org/10.1371/journ​al.pgen.1003372 org/10.2174/13816​12825​66619​07051​93227
Beleza, S., Santos, A. M., McEvoy, B., Alves, I., Martinho, C., Cameron, E., Chaplin, G. (2004). Geographic distribution of environmental fac-
Shriver, M. D., Parra, E. J., & Rocha, J. (2013). The timing of pigmen- tors influencing human skin coloration. American Journal of
tation lightening in Europeans. Molecular Biology and Evolution, 30(1), Physical Anthropology, 125(3), 292–­3 02. https://doi.org/10.1002/
24–­35. https://doi.org/10.1093/molbe​v/mss207 ajpa.10263
Blum, H. F. (1961). Does the melanin pigment of human skin have adap- Chaplin, G., & Jablonski, N. G. (2002). Environmental correlates of
tive value? Quarterly Review of Biology, 36(1), 50–­63. human skin color, revisited. American Journal of Physical Anthropology,
Bonilla, C., Gutierrez, G., Parra, E. J., Kline, C., & Shriver, M. D. (2005). 117(S34), 53. https://doi.org/10.1002/ajpa.20010
Admixture analysis of a rural population of the state of Guerrero, Chaplin, G., & Jablonski, N. G. (2009). Vitamin D and the evolution of
Mexico. American Journal of Physical Anthropology, 128(4), 861–­869. human depigmentation. American Journal of Physical Anthropology,
https://doi.org/10.1002/ajpa.20227 139(4), 451–­461. https://doi.org/10.1002/ajpa.21079
Bora, S., & Cantorna, M. T. (2017). The role of UVR and vitamin D Chaplin, G., & Jablonski, N. G. (2013). The human environment and the
on T cells and inflammatory bowel disease. Photochemical & vitamin D compromise: Scotland as a case study in human biocultural
Photobiological Sciences, 16(3), 347–­353. https://doi.org/10.1039/ adaptation and disease susceptibility. Human Biology, 85(4), 529–­552.
C6PP0​0266H https://doi.org/10.3378/027.085.0402
Borradale, D. C., Isenring, E., Hacker, E., & Kimlin, M. G. (2014). Exposure Clemens, T. L., Henderson, S. L., Adams, J. S., & Holick, M. F. (1982).
to solar ultraviolet radiation is associated with a decreased folate Increased skin pigment reduces the capacity of skin to synthesise
724 | JABLONSKI

vitamin D3. The Lancet, 319(8263), 74–­76. https://doi.org/10.1016/ pigmentation in humans. Pigment Cell & Melanoma Research, 22(4),
S0140​-­6736(82)90214​-­8 420–­434. https://doi.org/10.1111/j.1755-­148X.2009.00588.x
Coussens, A. K., Naude, C. E., Goliath, R., Chaplin, G., Wilkinson, R. J., Elias, P. M., Menon, G., Wetzel, B. K., & Williams, J. W. (2010). Barrier re-
& Jablonski, N. G. (2015). High-­dose vitamin D3 reduces deficiency quirements as the evolutionary “driver” of epidermal pigmentation in
caused by low UVB exposure and limits HIV-­1 replication in urban humans. American Journal of Human Biology, 22(4), 526–­537. https://
Southern Africans. Proceedings of the National Academy of Sciences, doi.org/10.1002/ajhb.21043
112(26), 8052–­8 057. https://doi.org/10.1073/pnas.15009​09112 Elias, P. M., & Williams, M. L. (2013). Re-­appraisal of current theories for
Crawford, N. G., Kelly, D. E., Hansen, M. E. B., Beltrame, M. H., Fan, the development and loss of epidermal pigmentation in hominins and
S., Bowman, S. L., Jewett, E., Ranciaro, A., Thompson, S., Lo, Y., modern humans. Journal of Human Evolution, 64(6), 687–­692. https://
Pfeifer, S. P., Jensen, J. D., Campbell, M. C., Beggs, W., Hormozdiari, doi.org/10.1016/j.jhevol.2013.02.003
F., Mpoloka, S. W., Mokone, G. G., Nyambo, T., Meskel, D. W., … Elias, P. M., & Williams, M. L. (2016). Basis for the gain and subsequent di-
Tishkoff, S. A. (2017). Loci associated with skin pigmentation iden- lution of epidermal pigmentation during human evolution: The barrier
tified in African populations. Science, 358(6365), eaan8433. https:// and metabolic conservation hypotheses revisited. American Journal
doi.org/10.1126/scien​ce.aan8433 of Physical Anthropology, 161(2), 189–­207. https://doi.org/10.1002/
Damgaard, P. D. B., Marchi, N., Rasmussen, S., Peyrot, M., Renaud, G., ajpa.23030
Korneliussen, T., Moreno-­Mayar, J. V., Pedersen, M. W., Goldberg, A., Elwood, J. M. (1983). Can vitamins prevent neural tube defects? Canadian
Usmanova, E., Baimukhanov, N., Loman, V., Hedeager, L., Pedersen, Medical Association Journal, 129(10), 1088–­1092.
A. G., Nielsen, K., Afanasiev, G., Akmatov, K., Aldashev, A., Alpaslan, Fajuyigbe, D., Lwin, S. M., Diffey, B. L., Baker, R., Tobin, D. J., Sarkany, R.
A., … Willerslev, E. (2018). 137 ancient human genomes from across P. E., & Young, A. R. (2018). Melanin distribution in human epidermis
the Eurasian steppes. Nature, 557(7705), 369–­ 374. https://doi. affords localized protection against DNA photodamage and concurs
org/10.1038/s4158​6-­018-­0 094-­2 with skin cancer incidence difference in extreme phototypes. The
Darwin, C. (1871). The descent of man, and selection in relation to sex. FASEB Journal, 32(7), 3700–­3706. https://doi.org/10.1096/fj.20170​
London: John Murray. 1472R
de Barros Damgaard, P., Martiniano, R., Kamm, J., Moreno-­Mayar, J. V., Fajuyigbe, D., & Young, A. R. (2016). The impact of skin colour on human
Kroonen, G., Peyrot, M., Barjamovic, G., Rasmussen, S., Zacho, C., photobiological responses. Pigment Cell & Melanoma Research, 29(6),
Baimukhanov, N., Zaibert, V., Merz, V., Biddanda, A., Merz, I., Loman, 607–­618. https://doi.org/10.1111/pcmr.12511
V., Evdokimov, V., Usmanova, E., Hemphill, B., Seguin-­Orlando, A., Feng, Y., McQuillan, M. A., & Tishkoff, S. A. (2021). Evolutionary genet-
… Willerslev, E. (2018). The first horse herders and the impact of ics of skin pigmentation in African populations. Human Molecular
early Bronze Age steppe expansions into Asia. Science, 360(6396), Genetics, dab007. https://doi.org/10.1093/hmg/ddab007
eaar7711. https://doi.org/10.1126/scien​ce.aar7711 Fleming, A., & Copp, A. J. (1998). Embryonic folate metabolism and
Del Bino, S., Duval, C., & Bernerd, F. (2018). Clinical and biological char- mouse neural tube defects. Science, 280, 2107–­ 2109. https://doi.
acterization of skin pigmentation diversity and its consequences on org/10.1126/scien​ce.280.5372.2107
UV impact. International Journal of Molecular Sciences, 19(9), 2668. Fukuwatari, T., Fujita, M., & Shibata, K. (2009). Effects of UVA irradi-
https://doi.org/10.3390/ijms1​9092668 ation on the concentration of folate in human blood. Bioscience,
Dennell, R., & Petraglia, M. D. (2012). The dispersal of Homo sapi- Biotechnology, and Biochemistry, 73(2), 322–­ 327. https://doi.
ens across southern Asia: How early, how often, how complex? org/10.1271/bbb.80530
Quaternary Science Reviews, 47, 15–­ 22. https://doi.org/10.1016/j. Galway-­Witham, J., & Stringer, C. (2018). How did Homo sapiens evolve?
quasc​irev.2012.05.002 Science, 360(6395), 1296–­1298. https://doi.org/10.1126/scien​
Diamond, J. M. (1991). The rise and fall of the third chimpanzee. London: ce.aat6659
Radius. Gambichler, T., Sauermann, K., Bader, A., Altmeyer, P., & Hoffmann, K.
Diamond, J. M. (2005). Geography and skin color. Nature, 435, 283–­284. (2001). Serum folate levels after UVA exposure: A two-­group paral-
https://doi.org/10.1038/435283a lel randomised controlled trial. BMC Dermatology, 1(1), 8. https://doi.
Diffey, B. L. (1991). Solar ultraviolet radiation effects on biological sys- org/10.1186/1471-­5945-­1-­8
tems. Physics in Medicine and Biology, 36(3), 299–­328. https://doi.org Giovannucci, E., Rimm, E. B., Ascherio, A., Stampfer, M. J., Colditz, G. A.,
/10.1088/0031-­9155/36/3/001 & Willett, W. C. (1995). Alcohol, low-­methionine, low folate diets, and
Diffey, B. L. (1999). Human exposure to ultraviolet radiation. In J. L. M. risk of colon cancer in men. Journal of the National Cancer Institute,
Hawk (Ed.), Photodermatology (pp. 5–­24). Arnold. 87(4), 265–­273. https://doi.org/10.1093/jnci/87.4.265
Eaaswarkhanth, M., Xu, D., Flanagan, C., Rzhetskaya, M., Hayes, M. G., Gittelman, R. M., Schraiber, J. G., Vernot, B., Mikacenic, C., Wurfel, M.
Blekhman, R., Blekhman, R., & Gokcumen, O. (2016). Atopic derma- M., & Akey, J. M. (2016). Archaic hominin admixture facilitated adap-
titis susceptibility variants in filaggrin hitchhike hornerin selective tation to Out-­of-­Africa environments. Current Biology, 26(24), 3375–­
sweep. Genome Biology and Evolution, 8(10), 3240–­3255. https://doi. 3382. https://doi.org/10.1016/j.cub.2016.10.041
org/10.1093/gbe/evw242 Goebel, T., Waters, M. R., & O'Rourke, D. H. (2008). The Late Pleistocene
El Hayek, J., Egeland, G., & Weiler, H. (2010). Vitamin D status of dispersal of modern humans in the Americas. Science, 319(5869),
Inuit preschoolers reflects season and vitamin D intake. The 1497–­1502. https://doi.org/10.1126/scien​ce.1153569
Journal of Nutrition, 140(10), 1839–­1845. https://doi.org/10.3945/ Goodwin, Z. A., & de Guzman Strong, C. (2017). Recent positive se-
jn.110.124644 lection in genes of the mammalian epidermal differentiation com-
Elias, P. M. (2005). Stratum corneum defensive functions: An integrated plex locus. Frontiers in Genetics, 7, 227. https://doi.org/10.3389/
view. Journal of Investigative Dermatology, 125(2), 183–­200. https:// fgene.2016.00227
doi.org/10.1111/j.0022-­202X.2005.23668.x Gozdzik, A., Barta, J. L., Wu, H., Wagner, D., Cole, D. E., Vieth, R.,
Elias, P. M., & Friend, D. S. (1975). The permeability barrier in mamma- Whiting, S., & Parra, E. J. (2008). Low wintertime vitamin D lev-
lian epidermis. The Journal of Cell Biology, 65(1), 180–­191. https://doi. els in a sample of healthy young adults of diverse ancestry liv-
org/10.1083/jcb.65.1.180 ing in the Toronto area: Associations with vitamin D intake and
Elias, P. M., Menon, G., Wetzel, B. K., & Williams, J. W. (2009). Evidence skin pigmentation. BMC Public Health, 8(1), 336–­3 44. https://doi.
that stress to the epidermal barrier influenced the development of org/10.1186/1471-­2458-­8-­336
JABLONSKI | 725

Greaves, M. (2014). Was skin cancer a selective force for black pigmen- not an essential regulator. Science, 211(4482), 590–­593. https://doi.
tation in early hominin evolution? Proceedings of the Royal Society org/10.1126/scien​ce.6256855
B: Biological Sciences, 281(1781), 1–­10. https://doi.org/10.1098/ Hsieh, P., Veeramah, K. R., Lachance, J., Tishkoff, S. A., Wall, J. D.,
rspb.2013.2955 Hammer, M. F., & Gutenkunst, R. N. (2016). Whole-­genome sequence
Groucutt, H. S., Petraglia, M. D., Bailey, G., Scerri, E. M. L., Parton, A., analyses of Western Central African Pygmy hunter-­gatherers reveal
Clark-­Balzan, L., Jennings, R. P., Lewis, L., Blinkhorn, J., Drake, N. a complex demographic history and identify candidate genes under
A., Breeze, P. S., Inglis, R. H., Devès, M. H., Meredith-­Williams, M., positive natural selection. Genome Research, 26(3), 279–­290. https://
Boivin, N., … Scally, A. (2015). Rethinking the dispersal of Homo sapi- doi.org/10.1101/gr.192971.115
ens out of Africa. Evolutionary Anthropology, 24(4), 149–­164. https:// Hublin, J.-­J., Ben-­Ncer, A., Bailey, S. E., Freidline, S. E., Neubauer, S.,
doi.org/10.1002/evan.21455 Skinner, M. M., Bergmann, I., Le Cabec, A., Benazzi, S., Harvati, K., &
Hanel, A., & Carlberg, C. (2020a). Skin colour and vitamin D: An update. Gunz, P. (2017). New fossils from Jebel Irhoud, Morocco and the pan-­
Experimental Dermatology, 29(9), 864–­875. https://doi.org/10.1111/ African origin of Homo sapiens. Nature, 546(7657), 289–­292. https://
exd.14142 doi.org/10.1038/natur​e22336
Hanel, A., & Carlberg, C. (2020b). Vitamin D and evolution: Pharmacologic Iliescu, F. M., Chaplin, G., Rai, N., Jacobs, G. S., Chaplin, G., Basu Mallick,
implications. Biochemical Pharmacology, 173, 113595. https://doi. C., Mishra, A., Thangaraj, K., & Jablonski, N. G. (2018). The influence
org/10.1016/j.bcp.2019.07.024 of genes, the environment, and social factors on the evolution of
Harding, R. M., Healy, E., Ray, A. J., Ellis, N. S., Flanagan, N., Todd, C., skin color diversity in India. American Journal of Human Biology, 30(5),
Dixon, C., Sajantila, A., Jackson, I. J., Birch-­Machin, M. A., & Rees, e23170. https://doi.org/10.1002/ajhb.23170
J. L. (2000). Evidence for variable selective pressures at MC1R. Izagirre, N., Garcia, I., Junquera, C., de la Rua, C., & Alonso, S. (2006).
American Journal of Human Genetics, 66, 1351–­ 1361. https://doi. A scan for signatures of positive selection in candidate loci for skin
org/10.1086/302863 pigmentation in humans. Molecular Biology and Evolution, 23(9), 1697–­
Hasoun, L. Z., Bailey, S. W., Outlaw, K. K., & Ayling, J. E. (2015). 1706. https://doi.org/10.1093/molbe​v/msl030
Rearrangement and depletion of folate in human skin by ultraviolet Jablonski, N. G. (1992). Sun, skin colour and spina bifida: An exploration
radiation. British Journal of Dermatology, 173(4), 1087–­1090. https:// of the relationship between solar ultraviolet radiation, skin colour
doi.org/10.1111/bjd.13885 and neural tube defects. In N. W. Bruce (Ed.), Proceedings of the fifth
Headland, T. N. (1989). Population decline in a Philippine Negrito hunter-­ annual conference of the Australasian society for human biology (pp.
gatherer society. American Journal of Human Biology, 1(1), 59–­72. 455-­462). Centre for Human Biology.
https://doi.org/10.1002/ajhb.13100​10111 Jablonski, N. G. (1999). A possible link between neural tube defect
Henn, B. M., Botigué, L. R., Peischl, S., Dupanloup, I., Lipatov, M., Maples, and ultraviolet light exposure. Medical Hypotheses, 52(6), 581–­582.
B. K., Martin, A. R., Musharoff, S., Cann, H., Snyder, M. P., Excoffier, https://doi.org/10.1054/mehy.1997.0697
L., Kidd, J. M., & Bustamante, C. D. (2016). Distance from sub-­ Jablonski, N. G. (2004). The evolution of human skin and skin color. Annual
Saharan Africa predicts mutational load in diverse human genomes. Review of Anthropology, 33, 585–­623. https://doi.org/10.1146/annur​
Proceedings of the National Academy of Sciences, 113(4), E440–­E449. ev.anthro.33.070203.143955
https://doi.org/10.1073/pnas.15108​05112 Jablonski, N. G. (2012). Living color: The biological and social meaning of
Henn, B. M., Cavalli-­Sforza, L. L., & Feldman, M. W. (2012). The great skin color. Berkeley: University of California Press.
human expansion. Proceedings of the National Academy of Sciences, Jablonski, N. G., & Chaplin, G. (2000). The evolution of human skin
109(44), 17758–­17764. https://doi.org/10.1073/pnas.12123​8 0109 coloration. Journal of Human Evolution, 39(1), 57–­106. https://doi.
Henn, B. M., Steele, T. E., & Weaver, T. D. (2018). Clarifying distinct org/10.1006/jhev.2000.0403
models of modern human origins in Africa. Current Opinion in Jablonski, N. G., & Chaplin, G. (2010). Human skin pigmentation as an
Genetics & Development, 53, 148–­156. https://doi.org/10.1016/j. adaptation to UV radiation. Proceedings of the National Academy
gde.2018.10.003 of Sciences, 107(Suppl. 2), 8962–­8968. https://doi.org/10.1073/
Hennessy, A., Oh, C., Diffey, B. L., Wakamatsu, K., Ito, S., & Rees, J. L. pnas.09146​28107
(2005). Eumelanin and pheomelanin concentrations in human epi- Jablonski, N. G., & Chaplin, G. (2013). Epidermal pigmentation in the
dermis before and after UVB irradiation. Pigment Cell Research, 18(3), human lineage is an adaptation to ultraviolet radiation. Journal
220–­223. https://doi.org/10.1111/j.1600-­0749.2005.00233.x of Human Evolution, 65(5), 671–­ 675. https://doi.org/10.1016/j.
Hennessy, A., Oh, C., Rees, J., & Diffey, B. L. (2005). The photoadaptive jhevol.2013.06.004
response to ultraviolet exposure in human skin using ultraviolet spec- Jablonski, N. G., & Chaplin, G. (2014a). The evolution of skin pigmen-
trophotometry. Photodermatology, Photoimmunology & Photomedicine, tation and hair texture in people of African ancestry. Dermatologic
21(5), 229–­233. https://doi.org/10.1111/j.1600-­0781.2005.00170.x Clinics, 32(2), 113–­121. https://doi.org/10.1016/j.det.2013.11.003
Hippocrates (1849). The Genuine Works of Hippocrates: The Genuine Works Jablonski, N. G., & Chaplin, G. (2014b). Skin cancer was not a potent
of Hippocrates: Translated from the Greek with a preliminary discourse selective force in the evolution of protective pigmentation in early
and annotations. (F. Adams, Trans.). William Wood & Co. hominins. Proceedings of the Royal Society B: Biological Sciences,
Hlusko, L. J., Carlson, J. P., Chaplin, G., Elias, S. A., Hoffecker, J. F., 281(1789), https://doi.org/10.1098/rspb.2014.0517
Huffman, M., Jablonski, N. G., Monson, T. A., O’Rourke, D. H., James, W. P. T., Johnson, R. J., Speakman, J. R., Wallace, D. C., Frühbeck,
Pilloud, M. A., & Scott, G. R. (2018). Environmental selection G., Iversen, P. O., & Stover, P. J. (2019). Nutrition and its role in human
during the last ice age on the mother-­to-­infant transmission of vi- evolution. Journal of Internal Medicine, 285(5), 533–­549. https://doi.
tamin D and fatty acids through breast milk. Proceedings of the org/10.1111/joim.12878
National Academy of Sciences, 115(19), E4426–­E4432. https://doi. Jinam, T. A., Phipps, M. E., Aghakhanian, F., Majumder, P. P., Datar, F.,
org/10.1073/pnas.17117​8 8115 Stoneking, M., Sawai, H., Nishida, N., Tokunaga, K., Kawamura,
Holick, M., MacLaughlin, J., Clark, M., Holick, S., Potts, J., Anderson, S., Omoto, K., & Saitou, N. (2017). Discerning the origins of the
R., Blank, I., Parrish, J., & Elias, P. (1980). Photosynthesis of previ- Negritos, first Sundaland people: Deep divergence and archaic ad-
tamin D3 in human skin and the physiologic consequences. Science, mixture. Genome Biology and Evolution, 9(8), 2013–­2022. https://doi.
210(4466), 203–­205. https://doi.org/10.1126/scien​ce.6251551 org/10.1093/gbe/evx118
Holick, M. F., MacLaughlin, J. A., & Doppelt, S. H. (1981). Regulation of John, P. R., Makova, K., Li, W.-­H., Jenkins, T., & Ramsay, M. (2003). DNA
cutaneous previtamin D3 photosynthesis in man: Skin pigment is polymorphism and selection at the melanocortin-­1 receptor gene in
726 | JABLONSKI

normally pigmented southern African individuals. Annals of the New Lindqvist, P. G., Epstein, E., Landin-­Olsson, M., Åkerlund, M., & Olsson,
York Academy of Sciences, 994, 299–­ 3 06. https://doi.org/10.1111/ H. (2020). Women with fair phenotypes seem to confer a sur-
j.1749-­6632.2003.tb031​93.x vival advantage in a low UV milieu. A nested matched case control
Jones, P., Lucock, M. D., Veysey, M., Jablonski, N. G., Chaplin, G., & study. PLoS One, 15(1), e0228582. https://doi.org/10.1371/journ​
Beckett, E. (2018). Frequency of folate-­related polymorphisms var- al.pone.0228582
ies by skin pigmentation. American Journal of Human Biology, 30(2), Lips, P. (2006). Vitamin D physiology. Progress in Biophysics and
e23079. https://doi.org/10.1002/ajhb.23079 Molecular Biology, 92(1), 4–­8. https://doi.org/10.1016/j.pbiom​
Ju, D., & Mathieson, I. (2020). The evolution of skin pigmentation associ- olbio.2006.02.016
ated variation in West Eurasia. Proceedings of the National Academy of Lips, P. (2007). Vitamin D status and nutrition in Europe and Asia. Journal
Sciences. 118(1), e2009227118. https://doi.org/10.1073/pnas.20092​ of Steroid Biochemistry and Molecular Biology, 103(3–­5), 620–­625.
27118 https://doi.org/10.1016/j.jsbmb.2006.12.076
Juzeniene, A., Tam, T. T. T., Iani, V., & Moan, J. (2009). Lips, P., van Schoor, N. M., & de Jongh, R. T. (2014). Diet, sun, and
5-­Methyltetrahydrofolate can be photodegraded by endogenous lifestyle as determinants of vitamin D status. Annals of the New
photosensitizers. Free Radical Biology and Medicine, 47(8), 1199–­ York Academy of Sciences, 1317(1), 92–­98. https://doi.org/10.1111/
1204. https://doi.org/10.1016/j.freer​adbio​med.2009.07.030 nyas.12443
Kaidbey, K. H., Agin, P. P., Sayre, R. M., & Kligman, A. M. (1979). Liu, D., Fernandez, B. O., Hamilton, A., Lang, N. N., Gallagher, J. M.
Photoprotection by melanin -­ a comparison of black and Caucasian C., Newby, D. E., Feelisch, M., & Weller, R. B. (2014). UVA irra-
skin. Journal of the American Academy of Dermatology, 1(3), 249–­260. diation of human skin vasodilates arterial vasculature and lowers
https://doi.org/10.1016/S0190​-­9622(79)70018​-­1 blood pressure independently of nitric oxide synthase. Journal of
Kenney, W. L. (2017). Edward F. Adolph Distinguished Lecture: Skin-­ Investigative Dermatology Advance Online, https://doi.org/10.1038/
deep insights into vascular aging. Journal of Applied Physiology, jid.2014.27
123(5), 1024–­1038. https://doi.org/10.1152/jappl​p hysi​ Lona-­Durazo, F., Hernandez-­Pacheco, N., Fan, S., Zhang, T., Choi, J.,
ol.00589.2017 Kovacs, M. A., Loftus, S. K., Le, P., Edwards, M., Fortes-­Lima, C. A.,
Klaus, S. N. (1973). Some early scientific studies of skin color. Yale Journal Eng, C., Huntsman, S., Hu, D., Gómez-­C abezas, E. J., Marín-­Padrón,
of Biology and Medicine, 46(5), 334–­336. L. C., Grauholm, J., Mors, O., Burchard, E. G., Norton, H. L., … Parra,
Kuan, V., Martineau, A., Griffiths, C., Hypponen, E., & Walton, R. (2013). E. J. (2019). Meta-­analysis of GWA studies provides new insights on
DHCR7 mutations linked to higher vitamin D status allowed early the genetic architecture of skin pigmentation in recently admixed
human migration to Northern latitudes. BMC Evolutionary Biology, populations. BMC Genetics, 20(1), 59. https://doi.org/10.1186/s1286​
13(1), 144. https://doi.org/10.1186/1471-­2148-­13-­144 3-­019-­0765-­5
Lamason, R. L., Mohideen, M.-­A .-­P.-­K., Mest, J. R., Wong, A. C., Norton, Loomis, W. F. (1967). Skin-­pigment regulation of vitamin-­D biosynthesis
H. L., Aros, M. C., Jurynec, M. J., Mao, X., Humphreville, V. R., in man. Science, 157(3788), 501–­506. https://doi.org/10.1126/scien​
Humbert, J. E., Sinha, S., Moore, J. L., Jagadeeswaran, P., Zhao, W., ce.157.3788.501
Ning, G., Makalowska, I., McKeigue, P. M., O'donnell, D., Kittles, R., Lucock, M. D., Beckett, E., Martin, C., Jones, P., Furst, J., Yates, Z.,
… Cheng, K. C. (2005). SLC24A5, a putative cation exchanger, affects Jablonski, N. G., Chaplin, G., & Veysey, M. (2017). UV-­associated de-
pigmentation in zebrafish and humans. Science, 310(5755), 1782–­ cline in systemic folate: Implications for human nutrigenetics, health,
1786. https://doi.org/10.1126/scien​ce.1116238 and evolutionary processes. American Journal of Human Biology,
Langston, M., Dennis, L., Lynch, C., Roe, D., & Brown, H. (2017). Temporal 29(2), e22929. https://doi.org/10.1002/ajhb.22929
trends in satellite-­derived Erythemal UVB and implications for ambi- Lucock, M. D., & Daskalakis, I. (2000). New perspectives on folate status:
ent sun exposure assessment. International Journal of Environmental A differential role for the vitamin in cardiovascular disease, birth de-
Research and Public Health, 14(2), 176. https://doi.org/10.3390/ijerp​ fects and other conditions. British Journal of Biomedical Science, 57(3),
h1402​0176 254–­260.
Lao, O., de Gruijter, J. M., van Duijn, K., Navarro, A., & Kayser, M. (2007). Lucock, M. D., Daskalakis, I., Hinkins, M., & Yates, Z. (2001). An examina-
Signatures of positive selection in genes associated with human skin tion of polymorphic genes and folate metabolsim in mothers affected
pigmentation as revealed from analyses of single nucleotide poly- by spina bifida pregnancy. Molecular Genetics and Metabolism, 73(4),
morphisms. Annals of Human Genetics, 71(3), 354–­369. https://doi. 322–­332. https://doi.org/10.1006/mgme.2001.3205
org/10.1111/j.1469-­1809.2006.00341.x Lucock, M. D., Glanville, T., Ovadia, L., Yates, Z., Walker, J., & Simpson,
Latreille, J., Ezzedine, K., Elfakir, A., Ambroisine, L., Gardinier, S., Galan, N. (2010). Photoperiod at conception predicts C677T-­M THFR
P., Hercberg, S., Gruber, F., Rees, J., Tschachler, E., & Guinot, C. genotype: A novel gene-­e nvironment interaction. American Journal
(2009). MC1R gene polymorphism affects skin color and phenotypic of Human Biology, 22(4), 484–­4 89. https://doi.org/10.1002/
features related to sun sensitivity in a population of French adult ajhb.21022
women. Photochemistry and Photobiology, 85(6), 1451–­1458. https:// Mackintosh, J. A. (2001). The antimicrobial properties of melanocytes,
doi.org/10.1111/j.1751-­1097.2009.00594.x melanosomes and melanin and the evolution of black skin. Journal
Li, S., Schlebusch, C., & Jakobsson, M. (2014). Genetic variation reveals of Theoretical Biology, 211(2), 101–­ 113. https://doi.org/10.1006/
large-­scale population expansion and migration during the expan- jtbi.2001.2331
sion of Bantu-­speaking peoples. Proceedings of the Royal Society B: Madrigal, L., & Kelly, W. (2007). Human skin-­color sexual dimorphism:
Biological Sciences, 281(1793), 20141448. https://doi.org/10.1098/ A test of the sexual selection hypothesis. American Journal of
rspb.2014.1448 Physical Anthropology, 132(3), 470–­4 82. https://doi.org/10.1002/
Lieberman, D. E. (2011). Human locomotion and heat loss: An evolution- ajpa.20453
ary perspective. Comprehensive Physiology, 5(1), 99–­117. https://doi. Majumder, P. P., & Basu, A. (2014). A genomic view of the peopling and
org/10.1002/cphy.c140011 population structure of India. Cold Spring Harbor Perspectives in Biology,
Lin, M., Siford, R. L., Martin, A. R., Nakagome, S., Möller, M., Hoal, E. 7(4), a008540. https://doi.org/10.1101/cshpe​rspect.a008540
G., Bustamante, C. D., Gignoux, C. R., & Henn, B. M. (2018). Rapid Makova, K., & Norton, H. L. (2005). Worldwide polymorphism at
evolution of a skin-­lightening allele in southern African KhoeSan. the MC1R locus and normal pigmentation variation in humans.
Proceedings of the National Academy of Sciences, 115(52), 13324–­ Peptides, 26(10), 1901–­1908. https://doi.org/10.1016/j.pepti​
13329. https://doi.org/10.1073/pnas.18019​48115 des.2004.12.032
JABLONSKI | 727

Malaspinas, A.-­S., Westaway, M. C., Muller, C., Sousa, V. C., Lao, O., Nielsen, R., Akey, J. M., Jakobsson, M., Pritchard, J. K., Tishkoff, S., &
Alves, I., Bergström, A., Athanasiadis, G., Cheng, J. Y., Crawford, J. Willerslev, E. (2017). Tracing the peopling of the world through ge-
E., Heupink, T. H., Macholdt, E., Peischl, S., Rasmussen, S., Schiffels, nomics. Nature, 541(7637), 302–­310. https://doi.org/10.1038/natur​
S., Subramanian, S., Wright, J. L., Albrechtsen, A., Barbieri, C., … e21347
Willerslev, E. (2016). A genomic history of Aboriginal Australia. Norton, H. L. The evolution of the human pigmentation phenotype. In D.
Nature, 538(7624), 207–­214. https://doi.org/10.1038/natur​e18299 H. O'Rourke (Ed.), A companion to anthropological genetics (pp. 251-­
Martin, A. R., Lin, M., Granka, J. M., Myrick, J. W., Liu, X., Sockell, A., 265). Hoboken, New Jersey: Wiley-­Blackwell.
Atkinson, E. G., Werely, C. J., Möller, M., Sandhu, M. S., Kingsley, D. Norton, H. L., Friedlaender, J. S., Merriwether, D. A., Koki, G., Mgone, C.
M., Hoal, E. G., Liu, X., Daly, M. J., Feldman, M. W., Gignoux, C. R., S., & Shriver, M. D. (2006). Skin and hair pigmentation variation in
Bustamante, C. D., & Henn, B. M. (2017). An unexpectedly complex Island Melanesia. American Journal of Physical Anthropology, 130(2),
architecture for skin pigmentation in Africans. Cell, 171(6), 1340–­ 254–­268. https://doi.org/10.1002/ajpa.20343
1353.e1314. https://doi.org/10.1016/j.cell.2017.11.015 Norton, H. L., Kittles, R. A., Parra, E., McKeigue, P., Mao, X., Cheng, K.,
Martínez-­C adenas, C., López, S., Ribas, G., Flores, C., García, O., Sevilla, Canfield, V. A., Bradley, D. G., McEvoy, B., & Shriver, M. D. (2007).
A., Smith-­Zubiaga, I., Ibarrola-­V illaba, M., Pino-­Yanes, M. D. M., Genetic evidence for the convergent evolution of light skin in
Gardeazabal, J., Boyano, D., García de Galdeano, A., Izagirre, N., de Europeans and East Asians. Molecular Biology and Evolution, 24(3),
la Rúa, C., & Alonso, S. (2013). Simultaneous purifying selection on 710–­722. https://doi.org/10.1093/molbe​v/msl203
the ancestral MC1R allele and positive selection on the melanoma-­ Norton, H. L., Werren, E., & Friedlaender, J. (2015). MC1R diversity in
risk allele V60L in South Europeans. Molecular Biology and Evolution, Northern Island Melanesia has not been constrained by strong puri-
30(12), 2654–­2665. https://doi.org/10.1093/molbe​v/mst158 fying selection and cannot explain pigmentation phenotype variation
Mastropaolo, W., & Wilson, M. A. (1993). Effect of light on serum in the region. BMC Genetics, 16(1), 1–­15. https://doi.org/10.1186/
B12 and folate stability. Clinical Chemistry, 39(5), 913. https://doi. s1286​3-­015-­0277-­x
org/10.1093/clinc​hem/39.5.913 Oh, C., Hennessy, A., Ha, T., Bisset, Y., Diffey, B. L., & Rees, J. L. (2004).
Mathieson, I., Lazaridis, I., Rohland, N., Mallick, S., Patterson, N., The time course of photoadaptation and pigmentation studied
Roodenberg, S. A., Harney, E., Stewardson, K., Fernandes, D., Novak, using a novel method to distinguish pigmentation from erythema.
M., Sirak, K., Gamba, C., Jones, E. R., Llamas, B., Dryomov, S., Pickrell, Journal of Investigative Dermatology, 123(5), 965–­ 972. https://doi.
J., Arsuaga, J. L., de Castro, J. M. B., Carbonell, E., … Reich, D. (2015). org/10.1111/j.0022-­202X.2004.23464.x
Genome-­wide patterns of selection in 230 ancient Eurasians. Nature, Olivarius, F. F., Wulf, H. C., Therkildsen, P., Poulsen, T., Crosby, J., &
528(7583), 499–­503. https://doi.org/10.1038/natur​e16152 Norval, M. (1997). Urocanic acid isomers: Relation to body site,
McEvoy, B., Beleza, S., & Shriver, M. D. (2006). The genetic architecture pigmentation, stratum corneum thickness and photosensitivity.
of normal variation in human pigmentation: An evolutionary per- Archives of Dermatological Research, 289(9), 501–­ 505. https://doi.
spective and model. Human Molecular Genetics, 15(suppl_2), R176–­ org/10.1007/s0040​3 0050230
181. https://doi.org/10.1093/hmg/ddl217 Osborne, D. L., & Hames, R. (2014). A life history perspective on skin
McPeters, R. D., Bhartia, P. K., Krueger, A. J., Herman, J. R., Schlesinger, cancer and the evolution of skin pigmentation. American Journal
B. M., Wellemeyer, C. G., Seftor, C. J., Jaross, G., Taylor, S. L., Swissler, of Physical Anthropology, 153(1), 1–­8. https://doi.org/10.1002/
T., Torres, O., Labow, G., Byerly, W., & Cebula, R. P. (1996). Nimbus-­7 ajpa.22408
Total Ozone Mapping Spectrometer (TOMS) Data Products User's Guide. Pagani, L., Lawson, D. J., Jagoda, E., Mörseburg, A., Eriksson, A., Mitt,
NASA -­Scientific and Technical Information Branch. M., Clemente, F., Hudjashov, G., DeGiorgio, M., Saag, L., Wall, J.
Meinhardt, M., Krebs, R., Anders, A., Heinrich, U., & Tronnier, H. (2008). D., Cardona, A., Mägi, R., Sayres, M. A. W., Kaewert, S., Inchley, C.,
Wavelength-­dependent penetration depths of ultraviolet radiation Scheib, C. L., Järve, M., Karmin, M., … Metspalu, M. (2016). Genomic
in human skin. Journal of Biomedical Optics, 13(4), 044030. https:// analyses inform on migration events during the peopling of Eurasia.
doi.org/10.1117/1.2957970 Nature, 538(7624), 238–­242. https://doi.org/10.1038/natur​e19792
Moorjani, P., Thangaraj, K., Patterson, N., Lipson, M., Loh, P.-­R ., Patin, E., Lopez, M., Grollemund, R., Verdu, P., Harmant, C., Quach,
Govindaraj, P., Berger, B., Reich, D., & Singh, L. (2013). Genetic ev- H., Laval, G., Perry, G. H., Barreiro, L. B., Froment, A., Heyer, E.,
idence for recent population mixture in India. The American Journal Massougbodji, A., Fortes-­Lima, C., Migot-­Nabias, F., Bellis, G.,
of Human Genetics, 93(3), 422–­438. https://doi.org/10.1016/j. Dugoujon, J.-­M., Pereira, J. B., Fernandes, V., Pereira, L., … Quintana-­
ajhg.2013.07.006 Murci, L. (2017). Dispersals and genetic adaptation of Bantu-­speaking
Moreno-­Mayar, J. V., Potter, B. A., Vinner, L., Steinrücken, M., populations in Africa and North America. Science, 356(6337), 543–­
Rasmussen, S., Terhorst, J., Kamm, J. A., Albrechtsen, A., Malaspinas, 546. https://doi.org/10.1126/scien​ce.aal1988
A.-­S., Sikora, M., Reuther, J. D., Irish, J. D., Malhi, R. S., Orlando, L., Pavan, W. J., & Sturm, R. A. (2019). The genetics of human skin and hair
Song, Y. S., Nielsen, R., Meltzer, D. J., & Willerslev, E. (2018). Terminal pigmentation. Annual Review of Genomics and Human Genetics, 20(1),
Pleistocene Alaskan genome reveals first founding population of 41–­72. https://doi.org/10.1146/annur​ev-­genom​- ­0 8311​8-­015230
Native Americans. Nature, 553, 203–­207. https://doi.org/10.1038/ Perry, C. A., Renna, S. A., Khitun, E., Ortiz, M., Moriarty, D. J., & Caudill,
natur​e25173 M. A. (2004). Ethnicity and race influence the folate status response
MRC Vitamin Study Research Group. (1991). Prevention of neu- to controlled folate intakes in young women. Journal of Nutrition,
ral tube defects: Results of the Medical Research Council 134(7), 1786–­1792. https://doi.org/10.1093/jn/134.7.1786
Vitamin Study. The Lancet, 338(8760), 131–­134. https://doi. Posth, C., Renaud, G., Mittnik, A., Drucker, D. G., Rougier, H., Cupillard,
org/10.1016/0140-­6736(91)90133​-­A C., Valentin, F., Thevenet, C., Furtwängler, A., Wißing, C., Francken,
Murray, F. G. (1934). Pigmentation, sunlight, and nutritional disease. M., Malina, M., Bolus, M., Lari, M., Gigli, E., Capecchi, G., Crevecoeur,
American Anthropologist, 36(3), 438–­4 45. https://doi.org/10.1525/ I., Beauval, C., Flas, D., … Krause, J. (2016). Pleistocene mitochondrial
aa.1934.36.3.02a00100 genomes suggest a single major dispersal of non-­Africans and a Late
Ng, X., Boyd, L., Dufficy, L., Naumovski, N., Blades, B., Travers, C., Lewis, Glacial population turnover in Europe. Current Biology, 26(6), 827–­
P., Sturm, J., Yates, Z., Townley-­Jones, M., Roach, P., Veysey, M., & 833. https://doi.org/10.1016/j.cub.2016.01.037
Lucock, M. (2009). Folate nutritional genetics and risk for hyper- Powell, A., Shennan, S., & Thomas, M. G. (2009). Late Pleistocene de-
tension in an elderly population sample. Journal of Nutrigenetics and mography and the appearance of modern human behavior. Science,
Nutrigenomics, 2(1), 1–­8. https://doi.org/10.1159/00016​0 079 324(5932), 1298–­1301. https://doi.org/10.1126/scien​ce.1170165
728 | JABLONSKI

Quillen, E. E. (2015). The evolution of tanning needs its day in the sun. Scerri, E. M. L., Thomas, M. G., Manica, A., Gunz, P., Stock, J. T., Stringer,
Human Biology, 87(4), 352–­360. https://doi.org/10.13110/​human​ C., Grove, M., Groucutt, H. S., Timmermann, A., Rightmire, G. P.,
biolo​g y.87.4.0352 d’Errico, F., Tryon, C. A., Drake, N. A., Brooks, A. S., Dennell, R. W.,
Quillen, E. E., Bauchet, M., Bigham, A. W., Delgado-­Burbano, M. E., Faust, Durbin, R., Henn, B. M., Lee-­Thorp, J., deMenocal, P., … Chikhi, L.
F. X., Klimentidis, Y. C., Mao, X., Stoneking, M., & Shriver, M. D. (2012). (2018). Did our species evolve in subdivided populations across
OPRM1 and EGFR contribute to skin pigmentation differences be- Africa, and why does it matter? Trends in Ecology & Evolution, 33(8),
tween Indigenous Americans and Europeans. Human Genetics, 131(7), 582–­594. https://doi.org/10.1016/j.tree.2018.05.005
1073–­1080. https://doi.org/10.1007/s0043​9-­011-­1135-­1 Sheehan, J. M., Potten, C. S., & Young, A. R. (1998). Tanning in human
Quillen, E. E., Norton, H. L., Parra, E. J., Lona-­Durazo, F., Ang, K. C., skin types II and III offers modest photoprotection against ery-
Illiescu, F. M., Pearson, L. N., Shriver, M. D., Lasisi, T., Gokcumen, O., thema. Photochemistry and Photobiology, 68(4), 588–­592. https://doi.
Starr, I., Lin, Y.-­L ., Martin, A. R., & Jablonski, N. G. (2019). Shades org/10.1111/j.1751-­1097.1998.tb025​18.x
of complexity: New perspectives on the evolution and genetic ar- Skoglund, P., & Mathieson, I. (2018). Ancient genomics of modern hu-
chitecture of human skin. American Journal of Physical Anthropology, mans: The first decade. Annual Review of Genomics and Human
168(S67), 4–­26. https://doi.org/10.1002/ajpa.23737 Genetics, 19(1), 381–­4 04. https://doi.org/10.1146/annur​ev-­genom​
Rana, B. K., Hewett-­Emmett, D., Jin, L., Chang, B.-­H.-­J., Sambuughin, N., -­0 8311​7-­021749
Lin, M., Watkins, S., Bamshad, M., Jorde, L. B., Ramsay, M., Jenkins, Stanhewicz, A. E., Alexander, L. M., & Kenney, W. L. (2015). Folic acid
T., & Li, W. H. (1999). High polymorphism at the human melanocortin supplementation improves microvascular function in older adults
1 receptor locus. Genetics, 151(4), 1547–­1557. through nitric oxide-­dependent mechanisms. Clinical Science, 129(2),
Rees, J. L. (2000). The melanocortin 1 receptor (MC1R): More than 159–­167. https://doi.org/10.1042/cs201​4 0821
just red hair. Pigment Cell Research, 13(3), 135–­ 140. https://doi. Stanhewicz, A. E., Bruning, R. S., Smith, C. J., Kenney, W. L., & Holowatz,
org/10.1034/j.1600-­0749.2000.130303.x L. A. (2012). Local tetrahydrobiopterin administration augments re-
Rees, J. L. (2003). Genetics of hair and skin color. Annual Review flex cutaneous vasodilation through nitric oxide-­dependent mecha-
of Genetics, 37, 67–­90. https://doi.org/10.1146/annur​ nisms in aged human skin. Journal of Applied Physiology, 112(5), 791–­
ev.genet.37.110801.143233 797. https://doi.org/10.1152/jappl​physi​ol.01257.2011
Reyes-­Centeno, H. (2016). Out of Africa and into Asia: Fossil and Stanhewicz, A. E., & Kenney, W. L. (2017). Role of folic acid in nitric oxide
genetic evidence on modern human origins and dispersals. bioavailability and vascular endothelial function. Nutrition Reviews,
Quaternary International, 416, 249–­ 262. https://doi.org/10.1016/j. 75(1), 61–­70. https://doi.org/10.1093/nutri​t /nuw053
quaint.2015.11.063 Stokowski, R. P., Pant, P. V. K., Dadd, T., Fereday, A., Hinds, D. A., Jarman,
Reyes-­Centeno, H., Ghirotto, S., Détroit, F., Grimaud-­Hervé, D., C., Filsell, W., Ginger, R. S., Green, M. R., van der Ouderaa, F. J., &
Barbujani, G., & Harvati, K. (2014). Genomic and cranial phenotype Cox, D. R. (2007). A genomewide association study of skin pigmenta-
data support multiple modern human dispersals from Africa and tion in a South Asian population. American Journal of Human Genetics,
a southern route into Asia. Proceedings of the National Academy of 81(6), 1119–­1132. https://doi.org/10.1086/522235
Sciences, 111(20), 7248–­7253. https://doi.org/10.1073/pnas.13236​ Stover, P. J. (2009). One-­ c arbon metabolism-­ genome interactions in
66111 folate-­associated pathologies. Journal of Nutrition, 139(12), 2402–­
Roberts, D. F., & Kahlon, D. P. S. (1972). Skin pigmentation and assorta- 2405. https://doi.org/10.3945/jn.109.113670
tive mating in Sikhs. Journal of Biosocial Science, 4(1), 91–­100. https:// Stringer, C. (2016). The origin and evolution of Homo sapiens. Philosophical
doi.org/10.1017/S0021​93200​0 00835X Transactions of the Royal Society B: Biological Sciences, 371, 1698.
Roberts, D. F., & Kahlon, D. P. S. (1976). Environmental correlations https://doi.org/10.1098/rstb.2015.0237
of skin colour. Annals of Human Biology, 3(1), 11–­ 22. https://doi. Sturm, R. A., Duffy, D. L., Box, N. F., Chen, W., Smit, D. J., Brown, D.
org/10.1080/03014​46760​0 001101 L., Stow, J. L., Leonard, J. H., & Martin, N. G. (2003). The role of
Robins, A. H. (1991). Biological perspectives on human pigmentation. melanocortin-­ 1 receptor polymorphism in skin cancer risk phe-
Cambridge Series in Biological and Evolutionary Anthropology. (Vol. 7). notypes. Pigment Cell Research, 16(3), 266–­ 272. https://doi.
Cambridge: Cambridge University Press. org/10.1034/j.1600-­0749.2003.00041.x
Robins, A. H. (2009). The evolution of light skin color: Role of vitamin D Sulem, P., Gudbjartsson, D. F., Stacey, S. N., Helgason, A., Rafnar,
disputed. American Journal of Physical Anthropology, 139(4), 447–­450. T., Magnusson, K. P., Manolescu, A., Karason, A., Palsson, A.,
https://doi.org/10.1002/ajpa.21077 Thorleifsson, G., Jakobsdottir, M., Steinberg, S., Pálsson, S.,
Rocha, J. (2020). The evolutionary history of human skin pigmentation. Jonasson, F., Sigurgeirsson, B., Thorisdottir, K., Ragnarsson, R.,
Journal of Molecular Evolution, 88, 77–­87. https://doi.org/10.1007/ Benediktsdottir, K. R., Aben, K. K., … Stefansson, K. (2007). Genetic
s0023​9-­019-­09902​-­7 determinants of hair, eye and skin pigmentation in Europeans.
Rogers, A. R., Iltis, D., & Wooding, S. (2004). Genetic variation at the Nature Genetics, 39(12), 1443–­1452. https://doi.org/10.1038/
MC1R locus and the time since loss of human body hair. Current ng.2007.13
Anthropology, 45(1), 105–­124. https://doi.org/10.1086/381006 Swope, V. B., & Abdel-­Malek, Z. A. (2018). MC1R: Front and center in the
Ross, A. B., Johansson, A., Ingman, M., & Gyllensten, U. (2006). Lifestyle, bright side of dark eumelanin and DNA repair. International Journal
genetics, and disease in Sami. Croatian Medical Journal, 47(4), of Molecular Sciences, 19(9), 2667. https://doi.org/10.3390/ijms1​
553–­565. 9092667
Røyrvik, E. C., Yuldasheva, N., Tonks, S., Winney, B., Ruzibakiev, R., Wells, Tadokoro, T., Kobayashi, N., Zmudzka, B. Z., Ito, S., Wakamatsu, K.,
R. S., & Bodmer, W. F. (2018). Genetic patterning in Central Eurasia: Yamaguchi, Y., Korossy, K. S., Miller, S. A., Beer, J. Z., & Hearing, V. J.
Population history and pigmentation. bioRxiv. 255117. https://doi. (2003). UV-­induced DNA damage and melanin content in human skin
org/10.1101/255117 differing in racial/ethnic origin. The FASEB Journal, 17, 1177–­1179.
Sahani, R. (2010). Nutritional and Health Status of Foragers of Andaman https://doi.org/10.1096/fj.02-­0 865fje
Islands Gauhati University, Guwahati, India. Retrieved from http://hdl. Van den Berghe, P. L., & Frost, P. (1986). Skin color preference, sexual
handle.net/10603/​67536 dimorphism and sexual selection: A case of gene culture evolution?
Sassi, F., Tamone, C., & D’Amelio, P. (2018). Vitamin D: Nutrient, hor- Ethnic and Racial Studies, 9, 87–­113.
mone, and immunomodulator. Nutrients, 10(11), 1656. https://doi. Villmoare, B., Kimbel, W. H., Seyoum, C., Campisano, C. J., DiMaggio, E.
org/10.3390/nu101​11656 N., Rowan, J., Braun, D. R., Arrowsmith, J. R., & Reed, K. E. (2015).
JABLONSKI | 729

Early Homo at 2.8 Ma from Ledi-­Geraru, Afar, Ethiopia. Science, Y., Li, M., Cao, X., … Su, B. (2016). A genetic mechanism for conver-
347(6228), 1352–­1355. https://doi.org/10.1126/scien​ce.aaa1343 gent skin lightening during recent human evolution. Molecular Biology
Walter, H. V. (1958). Der zusammenhang von hautfarbenverteilung und and Evolution, 33(5), 1177–­1187. https://doi.org/10.1093/molbe​v/
intensitat der ultravioletten strahlung. HOMO, 9(1), 1–­13. msw003
Wassermann, H. P. (1965). Human pigmentation and environmental ad- Young, A. R., Claveau, J., & Rossi, A. B. (2017). Ultraviolet radiation and
aptation. Archives of Environmental Health, 11(5), 691–­694. https:// the skin: Photobiology and sunscreen photoprotection. Journal of the
doi.org/10.1080/00039​896.1965.10664280 American Academy of Dermatology, 76(3, Supplement 1), S100-­S109.
White, T. D., Asfaw, B., DeGusta, D., Gilbert, H., Richards, G. D., Suwa, G., https://doi.org/10.1016/j.jaad.2016.09.038
& Howell, F. C. (2003). Pleistocene Homo sapiens from Middle Awash, Young, A. R., Morgan, K. A., Ho, T.-­W., Ojimba, N., Harrison, G. I.,
Ethiopia. Nature, 423, 742–­747. https://doi.org/10.1038/natur​e01669 Lawrence, K. P., Jakharia-­Shah, N., Wulf, H. C., Cruickshank, J. K.,
Wolf, S. T., Jablonski, N. G., Ferguson, S. B., Alexander, L. M., & Kenney, & Philipsen, P. A. (2020). Melanin has a small inhibitory effect on
W. L. (2020). Four weeks of vitamin D supplementation improves ni- cutaneous vitamin D synthesis: A comparison of extreme pheno-
tric oxide-­mediated microvascular function in college-­aged African types. Journal of Investigative Dermatology, 140(7), 1418–­1426e.1411.
Americans. American Journal of Physiology-­ Heart and Circulatory https://doi.org/10.1016/j.jid.2019.11.019
Physiology, 319(4), H906–­H914. https://doi.org/10.1152/ajphe​ Zihlman, A. L., & Cohn, B. A. (1988). The adaptive response of human
art.00631.2020 skin to the savanna. Human Evolution, 3(5), 397–­4 09. https://doi.
Wolf, S. T., & Kenney, W. L. (2019). The vitamin D-­folate hypothesis in org/10.1007/BF024​47222
human vascular health. American Journal of Physiology-­Regulatory,
Integrative and Comparative Physiology, 317(3), R491–­R501. https://
doi.org/10.1152/ajpre​gu.00136.2019
How to cite this article: Jablonski NG. The evolution of human
Yang, Z., Shi, H., Ma, P., Zhao, S., Kong, Q., Bian, T., Gong, C., Zhao, Q.
I., Liu, Y., Qi, X., Zhang, X., Han, Y., Liu, J., Li, Q., Chen, H., & Su, B.
skin pigmentation involved the interactions of genetic,
(2018). Darwinian positive selection on the pleiotropic effects of environmental, and cultural variables. Pigment Cell Melanoma
KITLG explain skin pigmentation and winter temperature adapta- Res. 2021;34:707–­729. https://doi.org/10.1111/pcmr.12976
tion in Eurasians. Molecular Biology and Evolution, 35(9), 2272–­2283.
https://doi.org/10.1093/molbe​v/msy136
Yang, Z., Zhong, H., Chen, J., Zhang, X., Zhang, H., Luo, X., Xu, S., Chen,
H., Lu, D., Han, Y., Li, J., Fu, L., Qi, X., Peng, Y. I., Xiang, K., Lin, Q., Guo,
The evolution of skin pigmentation-associated
variation in West Eurasia
Dan Jua,1 and Iain Mathiesona,1
a
Department of Genetics, Perelman School of Medicine, University of Pennsylvania, Philadelphia, PA 19104

Edited by Anne C. Stone, Arizona State University, Tempe, AZ, and approved November 4, 2020 (received for review May 11, 2020)

Skin pigmentation is a classic example of a polygenic trait that has may have been selected independently or in parallel in one or
experienced directional selection in humans. Genome-wide associ- more of these source populations or may instead only have been
ation studies have identified well over a hundred pigmentation- selected after admixture. The impact of ancient shifts in ancestry
associated loci, and genomic scans in present-day and ancient pop- is difficult to resolve using present-day data, but using ancient
ulations have identified selective sweeps for a small number of DNA, we can separate the effects of ancestry and selection and
light pigmentation-associated alleles in Europeans. It is unclear identify which loci were selected in which populations.
whether selection has operated on all of the genetic variation as- Here, we use ancient DNA to track the evolution of loci that
sociated with skin pigmentation as opposed to just a small number are associated with skin pigmentation in present-day Europeans.
of large-effect variants. Here, we address this question using an- Although we cannot make predictions about the phenotypes of
cient DNA from 1,158 individuals from West Eurasia covering a ancient individuals or populations, we can assess the extent to
period of 40,000 y combined with genome-wide association sum- which they carried the same light pigmentation alleles that are
mary statistics from the UK Biobank. We find a robust signal of present today. This allows us to identify which pigmentation-
directional selection in ancient West Eurasians on 170 skin associated variants have changed in frequency due to positive
pigmentation-associated variants ascertained in the UK Biobank. selection and the timing of these selective events. We present a
However, we also show that this signal is driven by a limited num-
systematic survey of the evolution of European skin pigmentation-

EVOLUTION
ber of large-effect variants. Consistent with this observation, we
associated variation, tracking over a hundred loci over 40,000 y of
find that a polygenic selection test in present-day populations fails
human history—almost the entire range of modern human occu-
to detect selection with the full set of variants. Our data allow us
pation of Europe.
to disentangle the effects of admixture and selection. Most nota-
bly, a large-effect variant at SLC24A5 was introduced to Western Results
Europe by migrations of Neolithic farming populations but contin-
Skin Pigmentation-Associated SNPs Show a Signal of Positive Selection
ued to be under selection post-admixture. This study shows that
in Ancient West Eurasians. We obtained skin pigmentation-associated
Downloaded from https://www.pnas.org by 110.235.181.121 on April 5, 2024 from IP address 110.235.181.121.

the response to selection for light skin pigmentation in West Eur-


asia was driven by a relatively small proportion of the variants
SNPs from the UK Biobank GWAS for skin color released by the
that are associated with present-day phenotypic variation. Neale Lab (14). We analyzed the evolution of these variants using
two datasets of publicly available ancient DNA (Fig. 1 A–C and SI
skin pigmentation | polygenic selection | complex traits | evolution | Appendix, Fig. S1). The first (“capture-shotgun”) consists of data
from 1,158 individuals dating from 45,000–715 y BP genotyped at
ancient DNA
∼1.2 million variants (7, 16, 18–52). The second (“shotgun”) is a

S kin pigmentation exhibits a gradient of variation across hu-


man populations that tracks with latitude (1). This gradient is
thought to reflect selection for lighter skin pigmentation at higher
Significance

latitudes, as lower UVB exposure reduces vitamin D biosynthesis, Some of the genes responsible for the evolution of light skin
which affects calcium homeostasis and immunity (2–4). Studies of pigmentation in Europeans show signals of positive selection
present-day and ancient populations have revealed signatures of in present-day populations. Recently, genome-wide association
selection at skin pigmentation loci (5–9), and single-nucleotide studies have highlighted the highly polygenic nature of skin
polymorphisms (SNPs) associated with light skin pigmentation at pigmentation. It is unclear whether selection has operated on
some of these genes exhibit a signal of polygenic selection in all of these genetic variants or just a subset. By studying vari-
Western Eurasians (10). However, this observation, the only docu- ation in over a thousand ancient genomes from West Eurasia
mented signal of polygenic selection for skin pigmentation, is based covering 40,000 y, we are able to study both the aggregate
on just four loci (SLC24A5, SLC45A2, TYR, and APBA2/OCA2). behavior of pigmentation-associated variants and the evolu-
Therefore, while the existence of selective sweeps at a handful tionary history of individual variants. We find that the evolu-
of skin pigmentation loci is well established, the evidence for tion of light skin pigmentation in Europeans was driven by
polygenic selection—a coordinated shift in allele frequencies frequency changes in a relatively small fraction of the genetic
across many trait-associated variants (11)—is less clear. Re- variants that are associated with variation in the trait today.
cently, genome-wide association studies (GWAS) of larger sam-
Author contributions: D.J. and I.M. designed research; D.J. and I.M. performed research;
ples and more diverse populations (12–15) have emphasized the I.M. contributed new reagents/analytic tools; D.J. analyzed data; and D.J. and I.M. wrote
polygenic architecture of skin pigmentation. This raises the the paper.
question of whether selection on skin pigmentation acted on all of The authors declare no competing interest.
this variation or was indeed driven by selective sweeps at a rela- This article is a PNAS Direct Submission.
tively small number of loci (11). This open access article is distributed under Creative Commons Attribution License 4.0
The impact of demographic transitions on the evolution of (CC BY).
skin pigmentation also remains an open question. The Holocene 1
To whom correspondence may be addressed. Email: danju@pennmedicine.upenn.edu or
history (∼12,000 y before present [BP]) of Europe was marked by mathi@pennmedicine.upenn.edu.
waves of migration and admixture between three highly diverged This article contains supporting information online at https://www.pnas.org/lookup/suppl/
populations: hunter-gatherers, Early Farmers, and Steppe an- doi:10.1073/pnas.2009227118/-/DCSupplemental.
cestry populations (16, 17). Skin pigmentation-associated loci Published December 21, 2020.

PNAS 2021 Vol. 118 No. 1 e2009227118 https://doi.org/10.1073/pnas.2009227118 | 1 of 8


Shotgun (40-1 kyBP) Capture-shotgun (40-1 kyBP) Capture-shotgun (15-1 kyBP)

A B C
1.0 1.0 1.0 Population
EF
HG
AFR SP
Ust Ishim Anzick Ust Ishim Mota
MA−1
EAS AFR
Tianyuan Mota Anzick
SAS Tianyuan MA−1 EAS
UK Biobank

EUR SAS

Score
Score

Score
0.5 0.5 EUR 0.5

242 SNPs 170 SNPs 170 SNPs


n = 248 n = 1158 n = 1079
p=2.67x10-3 p<1x10-4 p<1x10-4
0.0 0.0 0.0

40000 30000 20000 10000 0 40000 30000 20000 10000 0 10000 5000
Date (years BP) Date (years BP) Date (years BP)

D E F
MA−1
1.0 1.0 1.0 Population
EF
HG
SP
Ust Ishim
AFR
Manually curated

Tianyuan
Mota
Anzick
Tianyuan MA−1 Anzick AFR
Mota EAS
Score

Score

Score
Ust Ishim EAS
SAS
0.5 SAS 0.5 0.5

18 SNPs EUR 10 SNPs EUR


10 SNPs
n = 225 n = 1093 n = 1018
p=0.17 p<1x10-4 p<1x10-4
Downloaded from https://www.pnas.org by 110.235.181.121 on April 5, 2024 from IP address 110.235.181.121.

0.0 0.0 0.0


40000 30000 20000 10000 0 40000 30000 20000 10000 0 10000 5000
Date (years BP) Date (years BP) Date (years BP)

Fig. 1. Genetic scores for skin pigmentation over time. The solid lines indicate fitted mean scores with gray 95% confidence intervals. Weighted scores based
on UK Biobank SNPs and unweighted scores based on manually curated SNPs for samples in the (A and D) shotgun dataset, (B and E) capture-shotgun dataset,
and (C and F) capture-shotgun dataset dated within the past 15,000 y. Scores of labeled samples are plotted but not included in regressions. Averages for 1000
Genomes superpopulations are plotted at time 0. Area of points scale with number of SNPs genotyped in the individual.

subset of 248 individuals with genome-wide shotgun sequence data groups (Fig. 1C). Mesolithic hunter-gatherers carried fewer light
(16, 18–25, 28, 32, 35, 38, 40–44, 47, 49, 51, 53–60). This smaller pigmentation alleles than Early Farmer or Steppe ancestry pop-
dataset allows us to capture variants that may not be well tagged by ulations. This difference (0.091 score units) was similar to the
the genotyped SNPs in the capture-shotgun dataset. Weighting difference between Mesolithic hunter-gatherers and Early Upper
variants by their GWAS-estimated effect sizes, we show that the Paleolithic populations (0.097 score units). We tested for evidence
polygenic score—the weighted proportion of dark pigmentation of change in score over time within groups while controlling for
alleles—decreased significantly over the past 40,000 y (Fig. 1 A and changes in ancestry, finding no significant change within hunter-
B, P < 1 × 10−4 based on a genomic null distribution and accounting gatherer or Early Farmer populations (P = 0.32; P = 0.29). Steppe
for changes in ancestry; SI Appendix, Fig. S2). We recapitulate this ancestry populations, however, exhibited a significant decrease
significant decrease using a different set of UK Biobank GWAS (P = 0.007), with a steeper slope than that of Early Farmer or
summary statistics calculated using linear mixed models (61) (SI
hunter-gatherer (P = 0.004 and P = 0.08, respectively).
Appendix, Fig. S3) and identifying independent signals based on While the UK Biobank is well powered to detect variation that
predefined approximately independent linkage disequilibrium (LD)
is segregating in present-day Europeans, it may not detect vari-
blocks (62) (SI Appendix, Fig. S4), rather than LD clumping. The
ants that are no longer segregating. We therefore manually cu-
genetic scores of the oldest individuals in the dataset fall within the
range of present-day West African populations, showing that Early rated a separate list of 18 SNPs identified as associated with skin
Upper Paleolithic (∼50–20,000 y BP) individuals, such as Ust’Ishim, pigmentation from seven studies of populations representing
carried few of the light skin pigmentation alleles that are common in diverse ancestries (Dataset S1E). Using this set of SNPs to
present-day Europe. generate unweighted scores, we observed similar results as with
Over the past 15,000 y, which covers most of the data, the the UK Biobank ascertained SNPs (Fig. 1 D–F). We find a sig-
polygenic score decreased (P < 1 × 10−4) at a similar rate to the nificant decrease in genetic score over both 40,000 and 15,000 y
estimated rate over the 40,000-y period (5.27 × 10−5 per year vs. in the capture-shotgun dataset (P < 1 × 10−4 for both), although
3.47 × 10−5 per year; P = 0.04). Stratifying individuals by ancestry the decrease in the shotgun dataset was not significant (P =
(hunter-gatherer, Early Farmer, and Steppe) using ADMIX- 0.17), probably reflecting a lack of power from the small number
TURE revealed baseline differences in genetic score across of ancient samples and SNPs.

2 of 8 | PNAS Ju and Mathieson


https://doi.org/10.1073/pnas.2009227118 The evolution of skin pigmentation-associated variation in West Eurasia
Ancestry and Selection Have Different Effects on the Histories of Each allele at rs6120849 (EDEM2) is also significantly associated with
Variant. Next, we investigated the evolution of each variant in- several other traits including increased sitting and standing height
dependently. We separated the effects of ancestry and selection in the UK Biobank, increased blood creatinine, and increased
in the data by regressing presence or absence of the alternate heel bone mineral density (64). The dark allele of rs7870409
allele for each pigmentation-associated SNP on both date and (RALGPS1) is associated with increased arm and trunk fat per-
ancestry, inferred using principal component analysis (PCA) centage (65). These observations raise the possibility of pleiotropic
(63). Significant effects of ancestry on frequency can be inter- effects constraining the evolution of these loci or of spurious
preted as changes in allele frequency that can be explained by pleiotropy due to uncorrected population stratification in the
changes in ancestry. Significant changes in frequency over time, GWAS.
after accounting for ancestry, can be interpreted as evidence for On the other hand, the changes in frequencies of several vari-
selection occurring during the analyzed time period. ants appear to have been driven largely by changes in ancestry. For
The SNP rs16891982 at the SLC45A2 locus shows the stron- example, rs4778123 near OCA2, rs2153271 near BNC2, and
gest evidence for selection across analyses (Fig. 2), but we note rs3758833 near CTSC show evidence that their changes in fre-
this does not necessarily imply positive selection over the entire quency were consistent with changes in genome-wide ancestry
time transect. We observed little evidence for demographic tran- (Fig. 2 B and C). At SLC24A5, rs2675345 shows evidence of
sitions in driving the increase in light allele frequency over time for change with both ancestry and time, suggesting that even after the
this SNP. In addition, we found robust signals of selection for the spread of the light allele from Anatolia into Western Europe in
light allele of rs1126809 near TYR, rs12913832 and rs1635168 near the Neolithic (7) selection continued to occur post-admixture.
HERC2, and rs7109255 near GRM5 (Fig. 2 B and C). In some Again, not all of these cases involve an increase in the frequency of
cases—for example at rs6120849 near EDEM2 and rs7870409 the light pigmentation allele over time. The light allele of
near RALGPS1—light alleles decreased in frequency more than rs4778123 (OCA2) was at high frequency in hunter-gatherers but
can be explained by changes in ancestry (Fig. 2 B and C). The dark lower in later populations (SI Appendix, Fig. S6B). From the

EVOLUTION
A B C
Downloaded from https://www.pnas.org by 110.235.181.121 on April 5, 2024 from IP address 110.235.181.121.

D E F

Fig. 2. Plot of P values for ancestry and time for individual skin pigmentation SNPs. The dashed lines indicate fifth percentile of P values. UK Biobank SNPs
using the shotgun dataset are plotted in A and capture-shotgun dataset in B and C. Manually curated SNPs using the shotgun dataset are plotted in D and
capture-shotgun dataset in E and F. SNPs are colored according to whether the light allele is increasing over time (red) or decreasing (blue), with saturation
determined by the magnitude of change.

Ju and Mathieson PNAS | 3 of 8


The evolution of skin pigmentation-associated variation in West Eurasia https://doi.org/10.1073/pnas.2009227118
manually curated set of SNPs (Fig. 2 D–F), rs12203592 near IRF4 over time than those with small effect sizes (P = 2.8 × 10−18;
also displays a marked effect of ancestry with higher light allele adjusted R2 = 0.36) (SI Appendix, Fig. S13A). To understand the
frequency in hunter-gatherers (SI Appendix, Fig. S6E). While contribution of these large GWAS effect size SNPs, we itera-
rs12203592 was not present in the UK Biobank summary statistics, tively removed SNPs from the polygenic score calculations (Fig.
another SNP at the IRF4 locus, rs3778607, was present but with a 4A). Removing the SNP rs16891982 at SLC45A2, we still found a
smaller ancestry effect (SI Appendix, Fig. S7). significant decrease of score over time (P < 1 × 10−4), but the
estimated rate of change (βtime) decreased by 39%. Removing
Lack of PBS Selection Signal across Pigmentation SNPs in Source the top two SNPs at the SLC45A2 and SLC24A5 locus further
Populations of Europeans. As we described in the previous sec- attenuated the signal with a decrease in βtime of 58% (P = 3.96 ×
tion, the frequencies of several skin pigmentation SNPs appear 10−4). We removed the five SNPs with the largest GWAS effect
to have been driven by changes in ancestry. This observation sizes and found the effect of time attenuated (P = 0.026) with a
suggests that their frequencies have diverged between the Eu- decrease in βtime of 78%. Removing the top 10 SNPs attenuated
ropean source populations and then subsequently been driven by the signal even more (P = 0.050) with an 81% decrease and
admixture. Frequency differences across the source populations removal of 15 or 20 SNPs abolished the signal (P = 0.147,
might reflect the action of either genetic drift or selection. To P = 0.52).
examine this question, we looked for evidence of selection at skin Using present-day populations from the 1000 Genomes Proj-
pigmentation SNPs in each ancestral population using the pop- ect (67), we tested for polygenic selection with the Qx statistic
ulation branch statistic (PBS) test (66). We ran ADMIXTURE (10), which measures overdispersion of trait-associated SNPs
with K = 3 on the capture-shotgun dataset (SI Appendix, Fig. S8) relative to the genome-wide expectation. We performed the test
and used the inferred source population allele frequencies to with different numbers of SNPs ordered by largest to smallest
calculate PBS for each source in 20-SNP nonoverlapping win- absolute GWAS effect size (Fig. 4B). Using all 170 SNPs in the
dows, with the 1000 Genomes CHB (Han Chinese) and YRI test, we failed to observe a signal of polygenic selection (P =
(Yoruba) populations as outgroups. 0.27). However, when we restricted to the top five largest GWAS
Few of the skin pigmentation loci show extreme PBS values effect size SNPs, we found a significant signal (P < 2.5 × 10−7).
(Fig. 3), and even fewer show evidence of divergent frequencies We observed significant Qx statistics at P < 0.05 up to the top 30
across ancient groups (SI Appendix, Fig. S10). On the Early SNPs, but it appears this signal is driven by the top SNPs of
Farmer and Steppe ancestry branches, but not the hunter- largest effect. After removing the top 5 SNPs and testing the
gatherer branch, the SLC24A5 locus exhibited the strongest remaining 25 top SNPs, the signal of polygenic selection dis-
signal (Fig. 3 B and C), indicating selection at the locus in both appeared (P = 0.59). Examining the predicted variance explained
Early Farmer and Steppe source populations. The RABGAP1 per SNP (Fig. 4C and SI Appendix, Fig. S14), we predict that, of
locus also exhibited an elevated signal of selection in Early the UK Biobank SNPs, two SNPs at SLC45A2 and SLC24A5
Farmer and Steppe, which remained when using a 40-SNP win- make the largest contribution to the difference between West
Downloaded from https://www.pnas.org by 110.235.181.121 on April 5, 2024 from IP address 110.235.181.121.

dow size (SI Appendix, Fig. S11 B and C). In all three ancestry African and European populations. Most UK Biobank skin
groups, we observed an elevated PBS at the OCA2 locus, with pigmentation-associated SNPs are predicted to contribute rela-
the most extreme value on the hunter-gatherer branch. However, tively little to the between-population variance.
unlike the SLC24A5 variant, neither of these SNPs (rs644490
and rs9920172) showed a substantial effect of ancestry from the Discussion
individual SNP regression analysis. As a whole, the PBS values of Large whole-genome ancient DNA datasets have, in the past few
the 170 skin pigmentation SNPs do not significantly deviate from years, allowed us to track the evolution of variation associated
the genome-wide distribution of PBS for any of the ancient with both simple and complex traits (68, 69). The majority of
source populations (SI Appendix, Fig. S12). samples are from Western Eurasia, and we focus on that region,
noting that a parallel process of selection for light skin pig-
Signals of Selection Are Restricted to the SNPs with the Largest Effect mentation has acted in East Asian populations (9, 70). In West
Sizes. Since our results suggest that pigmentation-associated Eurasia, selection for skin pigmentation appears to have been
variants exhibit different changes in allele frequency, we fur- dominated by a small number of selective sweeps at large-effect
ther examined the signal of selection found in Fig. 1. In general, variants. There are a number of possible explanations for this.
SNPs with larger GWAS effect sizes changed more in frequency Many of the variants detected by the UK Biobank GWAS may

Fig. 3. Joint PBS distributions for 20-SNP windows across the genome for GBR-CHB-YRI on the x axis and X-CHB-YRI on the y axis, with X being (A) hunter-
gatherer, (B) Early Farmer, and (C) Steppe. The boxes represent windows centered around UK Biobank skin pigmentation SNPs and are colored redder according
to how much higher the light allele frequency is in X compared to YRI. Nearest genes are labeled. The orange and red lines represent the top 1 and 0.1 percentiles.

4 of 8 | PNAS Ju and Mathieson


https://doi.org/10.1073/pnas.2009227118 The evolution of skin pigmentation-associated variation in West Eurasia
A B C

Fig. 4. (A) Regressions of genetic score based on UK Biobank SNPs using the capture-shotgun dataset over date, with scores using all SNPs and iteratively
removing top GWAS effect size SNPs. (B) Qx empirical −log10(P values) using UK Biobank skin pigmentation-associated variants with all 1000 Genomes
populations. Different numbers of SNPs were used to calculate Qx, which were ordered by GWAS-estimated effect size. (C) Variance explained by UK Biobank
SNPs with nearest gene labeled between UK and YRI populations (y axis) and within the UK population (x axis).

have such small effects that they were effectively neutral or ex- MC1R explains a relatively large amount of variation within the
perienced such weak selection that we cannot detect it. Other UK but is predicted to explain relatively little of the variation
variants may not have responded to selection because of pleio- between Europe and West Africa (Fig. 4C). The TYRP1 locus

EVOLUTION
tropic constraint. The five SNPs with the largest effect sizes in has been previously identified as a target of selection in Euro-
UK Biobank do have some have small but significant associa- peans (5, 8, 9, 74, 75), although some studies (76, 77) have
tions with anthropometric phenotypes (for example, rs1805007 questioned this finding. Our analysis shows some support for
at MC1R with standing height), but by far their most significant selection at this locus, with a 20-SNP window centered around
associations are with hair color, facial aging, tanning, melanoma rs1325132 being in the top 1.1% of genome-wide PBS windows
risk, and other phenotypes that are likely related to pigmentation on the hunter-gatherer lineage. Some studies (76, 78) detect a
(64). Finally, GWAS effect size estimates in European pop- signal of recent selection in Europeans at the KITLG SNP
Downloaded from https://www.pnas.org by 110.235.181.121 on April 5, 2024 from IP address 110.235.181.121.

ulations today may not reflect the impact these variants had in rs12821256 that is functionally associated with blond hair color
the past due to epistatic or gene–environment interactions. This (79). We find no evidence of selection on the West Eurasian
would not affect the individual SNP time series and PBS analyses lineage at this SNP in our time series analysis, but this may reflect
since they do not rely on effect sizes as weights but would reduce a lack of power to detect a relatively small change in frequency.
power for the weighted polygenic score and Qx analyses. Finally, at the OCA2 locus, we recapitulate an observation of in-
Our analyses centered on 170 skin pigmentation-associated dependent selection in Europeans and East Asians (70, 80) (SI
SNPs that are present on the 1240K capture array. To check Appendix, Fig. S9). In the ancient populations, we found an ele-
for possible bias introduced by the capture array, we checked vated PBS signal around rs9920172 that was most prominent in
how many of the 242 SNPs used in the shotgun analysis were well hunter-gatherers but also elevated in Early Farmer and Steppe
tagged by the capture array. Of the 242 SNPs, 33 are shared, a ancestry populations (Fig. 3), suggesting a relatively early episode
further 67 are tagged at r2 ≥ 0.8, and a further 19 SNPs are of selection in West Eurasians.
tagged at r2 ≥ 0.5. However, while the 142 SNPs that are not Relatively dark skin pigmentation in Early Upper Paleolithic
tagged at r2 ≥ 0.8 by the capture array do contain some moderate Europe would be consistent with those populations being relatively
effect size SNPs, they do not contribute to the selection signal poorly adapted to high-latitude conditions as a result of having
(P = 0.31) (SI Appendix, Fig. S15D). Therefore, although the recently migrated from lower latitudes. On the other hand, although
capture array does not capture all of the variation contributing to we have shown that these populations carried few of the light pig-
present-day skin pigmentation variation, it does capture the vast mentation alleles that are segregating in present-day Europe, they
majority of the variation that contributed to the evolution of the may have carried different alleles that we cannot now detect. As an
phenotype, justifying the use of the capture-shotgun dataset for extreme example, Neanderthals and the Altai Denisovan individual
detailed investigation of the evolutionary trends. show genetic scores that are in a similar range to Early Upper
We find little evidence of parallel selection on independent Paleolithic individuals (SI Appendix, Table S1), but it is highly
haplotypes at skin pigmentation loci, suggesting that that dif- plausible that these populations, who lived at high latitudes for
ferences in allele frequency across ancestry groups were mostly hundreds of thousands of years, would have adapted independently
due to genetic drift. One exception is that the light allele at to low UV levels. For this reason, we cannot confidently make
SLC24A5 was nearly fixed in both Early Farmer and Steppe statements about the skin pigmentation of ancient populations.
ancestry populations due to selection. However, even for this Our study focused, for reasons of data availability, on the
variant we observe a signal of ongoing selection in our data even history of skin pigmentation evolution in West Eurasia. How-
after admixture with hunter-gatherers, indicating continued se- ever, there is strong evidence that a parallel trend of adaptation
lection after admixture. This is analogous to the rapid selection to low UVB conditions occurred in East Asia (15, 70, 81, 82).
at the same locus for the light allele introduced via admixture Less is known about the loci that have been under selection in
into the KhoeSan, who now occupy southern Africa (12, 71). East Asia, aside from some variants at OCA2 (80–82). Similarly,
We are also able to test previous claims about selection on multiple studies have documented selection for both lighter and
particular pigmentation genes. We find no evidence of positive darker skin pigmentation in parts of Africa (12, 13, 83). Future
selection in Europeans at the MC1R locus in contrast to previous work should test whether the process of adaptation in other parts
reports (72, 73). Among UK Biobank SNPs, rs1805007 near of the world was similar to that in Europe. The lack of known

Ju and Mathieson PNAS | 5 of 8


The evolution of skin pigmentation-associated variation in West Eurasia https://doi.org/10.1073/pnas.2009227118
skin pigmentation loci in scans of positive selection in East Asian from a GWAS in Europeans (95) with evidence of replication (P < 5 × 10−8) in
populations (84, 85) raises the possibility that selection may have the UK Biobank GWAS. For a GWAS in African populations (12), we selected
been more polygenic in East Asians than in Europeans. Finally, a SNP for each LD block, which was defined by the study authors, with the
greatest support based on multiple lines of evidence (e.g., P value and
the evidence for polygenic directional selection on other complex
functional annotations). For a GWAS in South Asians (94), we picked SNPs
traits in humans is inconclusive (86, 87). We suggest that detailed that were 200 kb away from each other and based the selection of SNPs that
studies of other phenotypes using ancient DNA can be helpful at were in physical proximity by considering P value. The other studies pre-
more generally identifying the types of processes that are im- sented a set of independent SNPs that we considered for inclusion. Lastly, we
portant in human evolution. manually clumped the SNPs curated across studies, selecting SNPs 200 kb
apart with more statistical and/or functional evidence than nearby SNPs.
Methods Specific details about the choices made on which SNPs to exclude can be
Capture-Shotgun Ancient DNA Dataset. We downloaded version 37.2 of the found in the “Notes” column for highlighted SNPs in Dataset S1C. We ended
publicly available datasets of 2,107 ancient and 5,637 present-day (Human with a final list of 18 SNPs for the manually curated set (Dataset S1E), 10 of
Origins dataset) samples from the Reich Lab website: https://reich.hms.harvard. which are present on the 1240K capture array. We also made pseudohaploid
edu/downloadable-genotypes-present-day-and-ancient-dna-data-compiled- calls for each of the 18 SNPs in the shotgun dataset, picking a random allele
published-papers. Ancient individuals were treated as pseudohaploid (i.e., from the reads at each site as for the overall dataset.
carry only one allele) because most samples are low coverage. For many of
these samples, 1,233,013 sites were genotyped using an in-solution capture ADMIXTURE Analysis on Capture-Shotgun Data. We performed unsupervised
method (88), while those with whole-genome shotgun sequence data were ADMIXTURE (98) on the dataset of ancient individuals with K = 3, which we
genotyped at the same set of sites. For our regression models, we included found to produce the lowest cross-validation error for 2 < K < 9. The three
ancient individuals from West Eurasia, defined here as the region west of the identified clusters could easily be identified as corresponding to hunter-
Ural Mountains (longitude, <60° E) and north of 35° N (SI Appendix, Fig. S1). gatherer, Early Farmer, and Steppe (also referred to as Yamnaya) ancestry
We excluded samples with less than 0.1× coverage. We removed duplicate and (SI Appendix, Fig. S8).
closely related samples (e.g., first-degree relatives), by calculating pairwise
identity by state (IBS) between ancient individuals. For each individual, we Time Series Analysis of Genetic Scores. We calculated genetic scores for
identified a corresponding individual with the highest IBS. Based on the dis- each individual from present-day 1000 Genomes populations (67) and an-
tribution of highest IBS values, we identified the pairs with IBS Z score >7 as cient individuals. We computed the scores in two ways. First, we weighted
duplicates and retained the individual with fewest missing sites. We also by the GWAS-estimated effect sizes. Because of variable coverage
manually removed related samples that were explicitly annotated, selecting across ancient samples, not all SNPs were present in a given sample. To
the highest coverage representative. Finally, we incorporated additional an- account for missing information in the creation of weighted scores, we
cient samples from the Iberian Peninsula from two recent papers (36, 45), devised a weighted proportion in which we divided the realized score over
applying the same criteria for coverage and relatedness. Ultimately, we com- the maximum possible score given the SNPs present in the sample:
piled a list of 1,158 ancient pseudohaploid individuals in our analyses covering
Scoreweighted = ∑m m
i=1 di βi =∑i=1 βi , where m is the number of skin pigmentation
a time span of the last 40,000 y. Only 11 individuals were dated to >15,000 y
SNPs genotyped in the individual, di is the presence of the dark allele at the
BP, and 1,147 individuals lived during the last 15,000 y.
ith SNP, and βi is the GWAS-estimated effect size (Fig. 1 A–C). For the
Downloaded from https://www.pnas.org by 110.235.181.121 on April 5, 2024 from IP address 110.235.181.121.

manually curated list of SNPs where we did not have comparable effect size
Shotgun Ancient DNA Dataset. We organized a dataset of samples that were
estimates, we computed an unweighted score, effectively assuming that all
shotgun sequenced without enrichment using the capture array (23, 25–32,
variants had the same effect size. This score is the proportion of dark alleles
35, 39, 42, 45, 47–51, 54, 56, 58, 60–67). To generate the first 10 PCs from
an individual carries out of the SNPs used in the construction of the score:
sites on the 1240K capture array, we combined shotgun samples with
Scoreunweighted = (1=m)∑m i=1 di , where m is the number of SNPs genotyped in
available genotype data from v37.2 of the Reich laboratory dataset and 15
the individual and di is the presence of the dark allele at the ith SNP
individuals that we ourselves pulled down from BAM files. We manually
removed duplicate samples in the shotgun dataset, preserving the sample (Fig. 1 D–F).
with higher coverage. We checked for first-degree relatives in the dataset by We used logistic regression to examine the association between weighted
considering familial annotations but found none. The final shotgun dataset and unweighted genetic score and time for all ancient samples. We fitted
included 249 individuals. separate models for ancient samples in the shotgun and capture-shotgun
datasets. We included ancestry as a covariate in the model by including the
Skin Pigmentation SNP Curation. We obtained summary statistics for UK first 10 PCs. PCA was performed using smartpca v16000 (63) to generate PCs
Biobank GWAS for skin color (data field 1717) from the publicly available from present day populations and we projected ancient individuals onto
release by the Neale Lab (version 3, Manifest Release 20180731) (14). The these axes of variation. For the shotgun dataset and the capture-shotgun
GWAS measured self-reported skin color as a categorical variable (very fair, dataset that included all samples (40,000 y BP), we used 1000 Genomes sam-
fair, light olive, dark olive, brown, black). To identify genome-wide signifi- ples. For the capture-shotgun dataset that restricted to samples later than
cant and independent SNPs, we performed clumping using PLINK v1.90b6.6 15,000 BP, we used West Eurasians from the Human Origins dataset (16).
(89) with 1000 Genomes GBR as an LD reference panel (--clump-p1 5 × 10−8 We model the score of each individual as the proportion of successes in a
--clump-r2 0.05 --clump-kb 250), and followed up with clumping based on binomial sample of size mi. That is, mi Si ∼ Binomial(mi , pi ), where, for indi-
physical distance to exclude SNPs within 100 kb of each other. We made two vidual i, mi is the number of SNPs genotyped, Si is the (weighted or un-
separate lists of UK Biobank SNPs for the shotgun and capture-shotgun weighted) score, the probability of success pi is given by the following:
datasets because the capture-shotgun dataset was restricted to the 1240K
array sites. For the capture-shotgun dataset, we intersected all UK Biobank pi
log( ) = α + βdate datei + βlat lati + βlon loni + βPC1 PC1i + . . . + βPC10 PC10i ,
SNPs with 1240K array SNPs before identifying 170 independent and 1 − pi
genome-wide significant SNPs (Dataset S1A), which were used in Figs. 1 B
and C, 2 B and C, 3, and 4. For the shotgun dataset, we identified 242 in- and datei, lati, and loni are the dates, latitude, and longitude of each indi-
dependent and genome-wide significant SNPs (Dataset S1B), which were vidual. For the weighted score, miSi is noninteger, but the likelihood can be
used in Figs. 1A and 2A. For SI Appendix, Fig. S4, we identified 93 SNPs from computed in the same way.
the Neale Lab GWAS (Dataset S1F) by picking the top SNPs that are genome- For the stratified analysis, we divided the capture-shotgun dataset into
wide significant from nearly independent LD blocks in the human genome mutually exclusive ancestry groups. We categorized individuals for this
specified beforehand (62). For SI Appendix, Fig. S3, we identified 176 SNPs analysis as hunter-gatherer if they carried over 60% of the ADMIXTURE-
(Dataset S1G) using the same clumping parameters with UK Biobank GWAS estimated hunter-gatherer component. For placement into the Early
summary statistics calculated with fastGWA (61). Farmer group, we required that an individual carry over 60% of that com-
We also manually curated a list of skin pigmentation-associated SNPs from ponent. Individuals we categorized as Steppe had over 30% of the AD-
the literature (Dataset S1C). We identified 12 suitable studies (12, 13, 15, 80, MIXTURE component and were dated to less than 5,000 y BP. Individuals
90–97), 9 of which were GWAS conducted in diverse populations (Dataset that are more recent than 5,000 y BP and have at least 30% of the Steppe
S1D). We did not include SNPs from all of the considered papers for our ADMIXTURE component were classified as Steppe even if they had more
analysis because some studies reported associations that were not genome- than 60% of the Early Farmer component. Based on these cutoffs, there
wide significant (P < 5 × 10−8). However, we included subsignificant SNPs were 102 hunter-gatherer, 499 Early Farmer, and 478 Steppe ancestry

6 of 8 | PNAS Ju and Mathieson


https://doi.org/10.1073/pnas.2009227118 The evolution of skin pigmentation-associated variation in West Eurasia
assigned individuals. We used the same logistic regression model as above (99). To estimate the variance explained between GBR and YRI, we calcu-
for the stratified analysis to control for ancestry. lated the between population variance as follows:
Because the fitted model parameters are overdispersed (SI Appendix, Fig.
S2), we computed P values from a genome-wide empirical null distribution VEbetween = β2 (favg (1 − favg ) − 0.5[f1 (1 − f1 ) + f2 (1 − f2 )]),
for βdate. We made this null distribution by running the regression model
described above on scores from sets of random, frequency-matched (±1%) where f1 and f2 are the allele frequencies of GBR and YRI, β is the UK Bio-
SNPs across the genome, maintaining the same effect sizes for the weighted bank GWAS-estimated effect size, and favg = 0.5(f1 + f2).
score. We matched the derived allele frequency based on the EUR super-
population frequencies from 1000 Genomes. We reported P values based on Population Branch Statistic. We calculated the PBS (66) for different sets of
10,000 random samples. ancient groups and present-day groups. For the ancient populations, we
used inferred source population allele frequencies from ADMIXTURE (the P
Time Series Analysis of Individual SNP Allele Frequencies. We performed lo- matrix), whereas for present-day groups, we used the observed allele fre-
gistic regression for each individual SNP using date of the sample and ancestry quencies estimated from 1000 Genomes YRI, CHB, and GBR. Using these
as covariates. That is, the probability that the haplotype sampled from in- frequencies, we estimated pairwise Hudson’s FST between groups for the PBS
dividual carries the derived allele is pi, where test. The estimator is calculated as follows:

pi (p1 − p2 )2 − p1n(1−p 1)
− p2n(1−p 2)
log( ) = α + βdate datei + βPC1 PC1i + . . . + βPC10 PC10i . F̂
ST =
1 −1 2 −1
,
1 − pi p1 (1 − p2 ) + p2 (1 − p1 )
We compared the full model above to a nested model with no principal where for population i, ni is the sample size and pi is the allele frequency of
components by performing a likelihood ratio test [in R we use anova(nested the sample. We excluded SNPs that had a missing rate of >90% in the
model, full model, test = ‘Chisq’)] to obtain a P value for the ancestry term capture-shotgun dataset. Because of variable coverage in ancient genomes,
which encompasses βPC1. . .βPC10. To obtain a P value for βdate, we compare a not all sites were used in the calculation. FST was calculated using 20- and
nested model without βdate to the full model. 40-SNP nonoverlapping windows throughout the genome. FST values were
transformed to calculate genetic divergence between populations as
Qx Polygenic Selection Test. We used the Qx test for directional, polygenic T = −log(1 – FST). We calculated PBS for ancient group X (i.e., hunter-
selection on skin pigmentation (10). For this test, we used 170 skin pig- gatherers, Early Farmers, or Steppe) using 1000 Genomes Han Chinese
mentation SNPs obtained from the UK Biobank in all 1000 Genomes pop- (CHB) and Yoruba (YRI) as follows:
ulations. We restricted sites to those on the 1240K and constructed a

EVOLUTION
covariance matrix from a total of one million SNPs. To calculate empirical P T X,YRI + T X,CHB − T YRI,CHB
values, we sampled a total of 500,000 null genetic values matching skin PBSX = .
2
pigmentation SNPs based on a ±2% frequency on the minor allele. However,
we report one million runs for the top seven SNPs and two million runs for We identified nearby genes by examining the National Center for Biotech-
the top five and six to better distinguish the deviation of the tested SNP sets nology Information RefSeq track for assembly GRCh37/hg19 (downloaded
from the expectation under drift. from https://genome.ucsc.edu/cgi-bin/hgTables) for genes that overlap with
windows with high PBS values. We note that the gene names are used simply
Downloaded from https://www.pnas.org by 110.235.181.121 on April 5, 2024 from IP address 110.235.181.121.

Variance in Skin Pigmentation Explained by Individual SNPs. We estimated the as labels for each locus and do not necessarily represent the causal genes.
variance explained in the phenotype by a SNP within the UK population using
the following: Code Availability. Scripts used to generate the main results and figures of this
paper are available at GitHub, https://github.com/mathilab/SkinPigmentationCode.
VEwithin = 2β2 f (1 − f ),
ACKNOWLEDGMENTS. We thank Sarah Tishkoff, Alexander Platt, Bárbara
where β is the UK Biobank GWAS-estimated effect size and f is the allele Bitarello, Samantha Cox, Arslan Zaidi, and Andrew Chen for helpful com-
frequency in the GBR population from 1000 Genomes. We calculated the ments on earlier versions of this manuscript. This research was supported by
proportion of variance explained as follows: a Research Fellowship from the Alfred P. Sloan Foundation (FG-2018-10647),
a New Investigator Research Grant from the Charles E. Kaufman Foundation
VEwithin (KA2018-98559), National Institute of General Medical Sciences Award
PVEwithin = ,
VEwithin + 2Nσ 2 f (1 − f ) R35GM133708, and National Research Service Award T32GM008216. The
content is solely the responsibility of the authors and does not necessarily
where N is the GWAS sample size (356,530) and σ is the SE of the estimated β represent the official views of the NIH.

1. N. G. Jablonski, G. Chaplin, The evolution of human skin coloration. J. Hum. Evol. 39, 13. A. R. Martin et al., An unexpectedly complex architecture for skin pigmentation in
57–106 (2000). Africans. Cell 171, 1340–1353.e14 (2017).
2. N. G. Jablonski, G. Chaplin, Colloquium paper: Human skin pigmentation as an ad- 14. Neale Lab, UK Biobank GWAS. www.nealelab.is/uk-biobank. Accessed 25 October
aptation to UV radiation. Proc. Natl. Acad. Sci. U.S.A. 107 (suppl. 2), 8962–8968 (2010). 2018.
3. A. A. Zaidi et al., Investigating the case of human nose shape and climate adaptation. 15. K. Adhikari et al., A GWAS in Latin Americans highlights the convergent evolution of
PLoS Genet. 13, e1006616 (2017). lighter skin pigmentation in Eurasia. Nat. Commun. 10, 358 (2019).
4. F. Sassi, C. Tamone, P. D’Amelio, D. Vitamin, Nutrient, hormone, and immunomod- 16. I. Lazaridis et al., Ancient human genomes suggest three ancestral populations for
ulator. Nutrients 10, 1656 (2018). present-day Europeans. Nature 513, 409–413 (2014).
5. B. F. Voight, S. Kudaravalli, X. Wen, J. K. Pritchard, A map of recent positive selection 17. I. Lazaridis, The evolutionary history of human populations in Europe. Curr. Opin.
in the human genome. PLoS Biol. 4, e72 (2006). Genet. Dev. 53, 21–27 (2018).
6. P. C. Sabeti et al.; International HapMap Consortium, Genome-wide detection and 18. C. Gamba et al., Genome flux and stasis in a five millennium transect of European
characterization of positive selection in human populations. Nature 449, 913–918 (2007). prehistory. Nat. Commun. 5, 5257 (2014).
7. I. Mathieson et al., Genome-wide patterns of selection in 230 ancient Eurasians. Na- 19. G. González-Fortes et al., Paleogenomic evidence for multi-generational mixing be-
ture 528, 499–503 (2015). tween Neolithic farmers and Mesolithic hunter-gatherers in the lower Danube basin.
8. J. K. Pickrell et al., Signals of recent positive selection in a worldwide sample of human Curr. Biol. 27, 1801–1810.e10 (2017).
populations. Genome Res. 19, 826–837 (2009). 20. T. Günther et al., Ancient genomes link early farmers from Atapuerca in Spain to
9. O. Lao, J. M. de Gruijter, K. van Duijn, A. Navarro, M. Kayser, Signatures of positive modern-day Basques. Proc. Natl. Acad. Sci. U.S.A. 112, 11917–11922 (2015).
selection in genes associated with human skin pigmentation as revealed from anal- 21. Z. Hofmanová et al., Early farmers from across Europe directly descended from
yses of single nucleotide polymorphisms. Ann. Hum. Genet. 71, 354–369 (2007). Neolithic Aegeans. Proc. Natl. Acad. Sci. U.S.A. 113, 6886–6891 (2016).
10. J. J. Berg, G. Coop, A population genetic signal of polygenic adaptation. PLoS Genet. 22. E. R. Jones et al., Upper Palaeolithic genomes reveal deep roots of modern Eurasians.
10, e1004412 (2014). Nat. Commun. 6, 8912 (2015).
11. J. K. Pritchard, J. K. Pickrell, G. Coop, The genetics of human adaptation: Hard sweeps, 23. E. R. Jones et al., The Neolithic transition in the Baltic was not driven by admixture
soft sweeps, and polygenic adaptation. Curr. Biol. 20, R208–R215 (2010). with early European farmers. Curr. Biol. 27, 576–582 (2017).
12. N. G. Crawford et al., Loci associated with skin pigmentation identified in African 24. G. M. Kılınç et al., The demographic development of the first farmers in Anatolia.
populations. Science 358, eaan8433 (2017). Curr. Biol. 26, 2659–2666 (2016).

Ju and Mathieson PNAS | 7 of 8


The evolution of skin pigmentation-associated variation in West Eurasia https://doi.org/10.1073/pnas.2009227118
25. A. Keller et al., New insights into the Tyrolean Iceman’s origin and phenotype as 64. G. McInnes et al., Global Biobank Engine: Enabling genotype-phenotype browsing for
inferred by whole-genome sequencing. Nat. Commun. 3, 698 (2012). Biobank summary statistics. Bioinformatics 35, 2495–2497 (2019).
26. M. Krzewin  ska et al., Ancient genomes suggest the eastern Pontic-Caspian steppe as 65. O. Canela-Xandri, K. Rawlik, A. Tenesa, An atlas of genetic associations in UK Bio-
the source of western Iron Age nomads. Sci. Adv. 4, eaat4457 (2018). bank. Nat. Genet. 50, 1593–1599 (2018).
27. M. Krzewin  ska et al., Genomic and strontium isotope variation reveal immigration 66. X. Yi et al., Sequencing of fifty human exomes reveals adaptation to high altitude.
patterns in a Viking Age town. Curr. Biol. 28, 2730–2738.e10 (2018). Science 329, 75–78 (2010).
28. M. E. Allentoft et al., Population genomics of Bronze Age Eurasia. Nature 522, 67. The 1000 Genomes Project Consortium, A global reference for human genetic vari-
167–172 (2015). ation. Nature 526, 68–74 (2015).
29. I. Lazaridis et al., Genetic origins of the Minoans and Mycenaeans. Nature 548, 68. S. L. Cox, C. B. Ruff, R. M. Maier, I. Mathieson, Genetic contributions to variation in human
214–218 (2017). stature in prehistoric Europe. Proc. Natl. Acad. Sci. U.S.A. 116, 21484–21492 (2019).
30. I. Lazaridis et al., Genomic insights into the origin of farming in the ancient Near East. 69. S. Mathieson, I. Mathieson, FADS1 and the timing of human adaptation to agricul-
Nature 536, 419–424 (2016). ture. Mol. Biol. Evol. 35, 2957–2970 (2018).
31. M. Lipson et al., Parallel palaeogenomic transects reveal complex genetic history of 70. H. L. Norton et al., Genetic evidence for the convergent evolution of light skin in
early European farmers. Nature 551, 368–372 (2017). Europeans and East Asians. Mol. Biol. Evol. 24, 710–722 (2007).
32. R. Martiniano et al., Genomic signals of migration and continuity in Britain before the 71. M. Lin et al., Rapid evolution of a skin-lightening allele in southern African KhoeSan.
Anglo-Saxons. Nat. Commun. 7, 10326 (2016). Proc. Natl. Acad. Sci. U.S.A. 115, 13324–13329 (2018).
33. R. Martiniano et al., The population genomics of archaeological transition in west 72. S. A. Savage et al., Nucleotide diversity and population differentiation of the mela-
Iberia: Investigation of ancient substructure using imputation and haplotype-based nocortin 1 receptor gene, MC1R. BMC Genet. 9, 31 (2008).
methods. PLoS Genet. 13, e1006852 (2017). 73. C. Martínez-Cadenas et al., Simultaneous purifying selection on the ancestral MC1R
34. I. Mathieson et al., The genomic history of southeastern Europe. Nature 555, 197–203 allele and positive selection on the melanoma-risk allele V60L in south Europeans.
(2018). Mol. Biol. Evol. 30, 2654–2665 (2013).
35. I. Olalde et al., Derived immune and ancestral pigmentation alleles in a 7,000-year-old 74. A. Urnikyte et al., Patterns of genetic structure and adaptive positive selection in the
Mesolithic European. Nature 507, 225–228 (2014). Lithuanian population from high-density SNP data. Sci. Rep. 9, 9163 (2019).
36. I. Olalde et al., The genomic history of the Iberian Peninsula over the past 8000 years. 75. S. R. Grossman et al., A composite of multiple signals distinguishes causal variants in
Science 363, 1230–1234 (2019). regions of positive selection. Science 327, 883–886 (2010).
37. I. Olalde et al., The Beaker phenomenon and the genomic transformation of north- 76. A. J. Stern, P. R. Wilton, R. Nielsen, An approximate full-likelihood method for in-
west Europe. Nature 555, 190–196 (2018). ferring selection and allele frequency trajectories from DNA sequence data. PLoS
38. I. Olalde et al., A common genetic origin for early farmers from Mediterranean Genet. 15, e1008384 (2019).
Cardial and Central European LBK cultures. Mol. Biol. Evol. 32, 3132–3142 (2015). 77. J. M. de Gruijter et al., Contrasting signals of positive selection in genes involved in
39. C. E. G. Amorim et al., Understanding 6th-century barbarian social organization and human skin-color variation from tests based on SNP scans and resequencing. Investig.
migration through paleogenomics. Nat. Commun. 9, 3547 (2018). Genet. 2, 24 (2011).
40. A. Omrak et al., Genomic evidence establishes Anatolia as the source of the European 78. Y. Field et al., Detection of human adaptation during the past 2000 years. Science
Neolithic gene pool. Curr. Biol. 26, 270–275 (2016). 354, 760–764 (2016).
41. L. Saag et al., Extensive farming in Estonia started through a sex-biased migration 79. C. A. Guenther, B. Tasic, L. Luo, M. A. Bedell, D. M. Kingsley, A molecular basis for
from the steppe. Curr. Biol. 27, 2185–2193.e6 (2017). classic blond hair color in Europeans. Nat. Genet. 46, 748–752 (2014).
42. S. Schiffels et al., Iron age and Anglo-Saxon genomes from East England reveal British 80. M. Edwards et al., Association of the OCA2 polymorphism His615Arg with melanin
migration history. Nat. Commun. 7, 10408 (2016). content in East Asian populations: Further evidence of convergent evolution of skin
43. M. Sikora et al., Ancient genomes show social and reproductive behavior of early pigmentation. PLoS Genet. 6, e1000867 (2010).
Downloaded from https://www.pnas.org by 110.235.181.121 on April 5, 2024 from IP address 110.235.181.121.

Upper Paleolithic foragers. Science 358, 659–662 (2017). 81. K. Eaton et al., Association study confirms the role of two OCA2 polymorphisms in
44. P. Skoglund et al., Genomic diversity and admixture differs for Stone-Age Scandina- normal skin pigmentation variation in East Asian populations. Am. J. Hum. Biol. 27,
vian foragers and farmers. Science 344, 747–750 (2014). 520–525 (2015).
45. V. Villalba-Mouco et al., Survival of late Pleistocene hunter-gatherer ancestry in the 82. Z. Yang et al., A genetic mechanism for convergent skin lightening during recent
Iberian Peninsula. Curr. Biol. 29, 1169–1177.e7 (2019). human evolution. Mol. Biol. Evol. 33, 1177–1187 (2016).
46. M. Unterländer et al., Ancestry and demography and descendants of Iron Age no- 83. F. Tekola-Ayele et al., Novel genomic signals of recent selection in an Ethiopian
mads of the Eurasian steppe. Nat. Commun. 8, 14615 (2017). population. Eur. J. Hum. Genet. 23, 1085–1092 (2015).
47. L. M. Cassidy et al., Neolithic and Bronze Age migration to Ireland and establishment 84. Y. Okada et al., Deep whole-genome sequencing reveals recent selection signatures
of the insular Atlantic genome. Proc. Natl. Acad. Sci. U.S.A. 113, 368–373 (2016). linked to evolution and disease risk of Japanese. Nat. Commun. 9, 1631 (2018).
48. P. B. Damgaard et al., 137 ancient human genomes from across the Eurasian steppes. 85. Y. Yasumizu et al., Genome-wide natural selection signatures are linked to genetic
Nature 557, 369–374 (2018). risk of modern phenotypes in the Japanese population. Mol. Biol. Evol. 37, 1306–1316
49. P. de Barros Damgaard et al., The first horse herders and the impact of early Bronze (2020).
Age steppe expansions into Asia. Science 360, eaar7711 (2018). 86. M. Sohail et al., Polygenic adaptation on height is overestimated due to uncorrected
50. D. M. Fernandes et al., A genomic Neolithic time transect of hunter-farmer admixture stratification in genome-wide association studies. eLife 8, 1–17 (2019).
in central Poland. Sci. Rep. 8, 14879 (2018). 87. J. J. Berg et al., Reduced signal for polygenic adaptation of height in UK Biobank.
51. Q. Fu et al., The genetic history of Ice Age Europe. Nature 534, 200–205 (2016). eLife 8, e39725 (2019).
52. Q. Fu et al., An early modern human from Romania with a recent Neanderthal an- 88. W. Haak et al., Massive migration from the steppe was a source for Indo-European
cestor. Nature 524, 216–219 (2015). languages in Europe. Nature 522, 207–211 (2015).
53. Q. Fu et al., Genome sequence of a 45,000-year-old modern human from western 89. C. C. Chang et al., Second-generation PLINK: Rising to the challenge of larger and
Siberia. Nature 514, 445–449 (2014). richer datasets. Gigascience 4, 7 (2015).
54. T. Günther et al., Population genomics of Mesolithic Scandinavia: Investigating early 90. C. T. Miller et al., cis-Regulatory changes in Kit ligand expression and parallel evo-
postglacial migration routes and high-latitude adaptation. PLoS Biol. 16, e2003703 (2018). lution of pigmentation in sticklebacks and humans. Cell 131, 1179–1189 (2007).
55. S. Brace et al., Ancient genomes indicate population replacement in Early Neolithic 91. J. Han et al., A genome-wide association study identifies novel alleles associated with
Britain. Nat. Ecol. Evol. 3, 765–771 (2019). hair color and skin pigmentation. PLoS Genet. 4, e1000074 (2008).
56. A. Seguin-Orlando et al., Genomic structure in Europeans dating back at least 92. S. I. Candille et al., Genome-wide association studies of quantitatively measured skin,
36,200 years. Science 346, 1113–1118 (2014). hair, and eye pigmentation in four European populations. PLoS One 7, e48294 (2012).
57. K. R. Veeramah et al., Population genomic analysis of elongated skulls reveals ex- 93. N. Hernandez-Pacheco et al., Identification of a novel locus associated with skin
tensive female-biased immigration in Early Medieval Bavaria. Proc. Natl. Acad. Sci. colour in African-admixed populations. Sci. Rep. 7, 44548 (2017).
U.S.A. 115, 3494–3499 (2018). 94. R. P. Stokowski et al., A genomewide association study of skin pigmentation in a
58. M. A. Yang et al., 40,000-year-old individual from Asia provides insight into early South Asian population. Am. J. Hum. Genet. 81, 1119–1132 (2007).
population structure in Eurasia. Curr. Biol. 27, 3202–3208.e9 (2017). 95. F. Liu et al., Genetics of skin color variation in Europeans: Genome-wide association
59. M. Rasmussen et al., The genome of a Late Pleistocene human from a Clovis burial site studies with functional follow-up. Hum. Genet. 134, 823–835 (2015).
in western Montana. Nature 506, 225–229 (2014). 96. H. Nan, P. Kraft, D. J. Hunter, J. Han, Genetic variants in pigmentation genes, pigmen-
60. M. Gallego Llorente et al., Ancient Ethiopian genome reveals extensive Eurasian tary phenotypes, and risk of skin cancer in Caucasians. Int. J. Cancer 125, 909–917 (2009).
admixture throughout the African continent. Science 350, 820–822 (2015). 97. S. Beleza et al., Genetic architecture of skin and eye color in an African-European
61. L. Jiang et al., A resource-efficient tool for mixed model association analysis of large- admixed population. PLoS Genet. 9, e1003372 (2013).
scale data. Nat. Genet. 51, 1749–1755 (2019). 98. D. H. Alexander, J. Novembre, K. Lange, Fast model-based estimation of ancestry in
62. T. Berisa, J. K. Pickrell, Approximately independent linkage disequilibrium blocks in unrelated individuals. Genome Res. 19, 1655–1664 (2009).
human populations. Bioinformatics 32, 283–285 (2016). 99. H. Shim et al., A multivariate genome-wide association analysis of 10 LDL sub-
63. N. Patterson, A. L. Price, D. Reich, Population structure and eigenanalysis. PLoS Genet. fractions, and their response to statin treatment, in 1868 Caucasians. PLoS One 10,
2, e190 (2006). e0120758 (2015).

8 of 8 | PNAS Ju and Mathieson


https://doi.org/10.1073/pnas.2009227118 The evolution of skin pigmentation-associated variation in West Eurasia
Diversity and evolution of the primate
rspb.royalsocietypublishing.org skin microbiome
Sarah E. Council1,3, Amy M. Savage4, Julie M. Urban3, Megan E. Ehlers3,
J. H. Pate Skene5, Michael L. Platt5,6,7,8, Robert R. Dunn9,10
Research and Julie E. Horvath2,3,11
1
Center for Science, Math and Technology Education, and 2Department of Biological and Biomedical Sciences,
Cite this article: Council SE, Savage AM,
North Carolina Central University, Durham, NC 27707, USA
Urban JM, Ehlers ME, Skene JHP, Platt ML, 3
North Carolina Museum of Natural Sciences, Raleigh, NC 27601, USA
4
Dunn RR, Horvath JE. 2016 Diversity and Department of Biology, Center for Computational & Integrative Biology, Rutgers University, Camden,
evolution of the primate skin microbiome. NJ 08103, USA
5
Department of Neurobiology, Duke University, Research Drive, Durham, NC 27710, USA
Proc. R. Soc. B 283: 20152586. 6
Department of Neuroscience, Perelman School of Medicine, 7Department of Psychology, School of Arts and
http://dx.doi.org/10.1098/rspb.2015.2586 Sciences, and 8Department of Marketing, the Wharton School, University of Pennsylvania, Philadelphia,
PA 19104, USA
9
Department of Applied Ecology and Keck Center for Behavioral Biology, North Carolina State University,
Raleigh, NC 27607, USA
10
Received: 27 October 2015 Center for Macroecology, Evolution and Climate, Natural History Museum of Denmark, University of
Copenhagen, Copenhagen Ø 2100, Denmark
Accepted: 1 December 2015 11
Department of Evolutionary Anthropology, Duke University, Durham, NC 27708, USA

Skin microbes play a role in human body odour, health and disease. Compared
with gut microbes, we know little about the changes in the composition of skin
Subject Areas: microbes in response to evolutionary changes in hosts, or more recent behav-
ioural and cultural changes in humans. No studies have used sequence-based
evolution, microbiology
approaches to consider the skin microbe communities of gorillas and chimpan-
zees, for example. Comparison of the microbial associates of non-human
Keywords:
primates with those of humans offers unique insights into both the ancient
microbiota, microbe, microbiome, and modern features of our skin-associated microbes. Here we describe the
primate, skin, axilla microbes found on the skin of humans, chimpanzees, gorillas, rhesus maca-
ques and baboons. We focus on the bacterial and archaeal residents in the
axilla using high-throughput sequencing of the 16S rRNA gene. We find
that human skin microbial communities are unique relative to those of other
Author for correspondence:
primates, in terms of both their diversity and their composition. These differ-
Julie E. Horvath ences appear to reflect both ancient shifts during millions of years of primate
e-mail: julie.horvath@naturalsciences.org evolution and more recent changes due to modern hygiene.

1. Introduction
The skin is the largest mammalian organ [1,2], and the primary physical barrier
between mammals and the outside world. The outermost layer of the skin com-
prises host cells along with billions of microbial symbionts. Skin microbes
inhibit the colonization of opportunistic or pathogenic microbiota [3,4], regulate
immune activation [5–7], and produce compounds that function as both phero-
mones [8,9] and allomones [10]. Skin microbes also influence the attractiveness
of hosts to blood-feeding insects, including mosquitoes, known to transmit
malaria, dengue and chikungunya [11 –14]. The significance of skin microbes
to human health and disease suggests that the features of the body that influ-
ence its composition might coevolve with host traits. On human skin, the
abundance and composition of microbes varies relatively predictably among
habitats on the body, as a function of many factors including the distribution
of glands and moisture [15 –17]. For example, the apocrine glands in the axillary
Electronic supplementary material is available organ in the armpit produce secretions that provide food for the microbes living
therein [18 –20]. Microbial constituents of the skin microbe communities play an
at http://dx.doi.org/10.1098/rspb.2015.2586 or
important role in odour creation due to their production of volatile organic
via http://rspb.royalsocietypublishing.org. compounds [18,19]. Body odour plays a central role in primate society in the

& 2016 The Author(s) Published by the Royal Society. All rights reserved.
context of mating, child rearing, predatory protection and ter- 30 CGBR HBR HR HC HC 2
ritorial marking [21,22]. While we are beginning to learn which

rspb.royalsocietypublishing.org
host and environmental factors influence skin microbes, we

no. OTUs (transformed)


know little about how the microbial composition of the skin
varies among closely related primates. 20
The composition of microbes on human skin might be
expected to differ significantly from that of our closest relatives,
the non-human primates, for at least three reasons. First, even 10
closely related primates differ in the distribution and abun-
dance (and likely chemical composition) of different skin
glands. Skin glands provide both food sources and habitat
for microbes [23–27]. For example, our closest evolutionary 0

Proc. R. Soc. B 283: 20152586


an

ee

la

on

e
relatives, chimpanzees and bonobos, share the majority of

u
ril
nz
m

aq
bo
go
hu

pa

ac
ba
our nuclear genome [28], yet differ in their abundance and

im

sm
ch
location of eccrine and apocrine glands [26,29] (see the elec-

u
es
rh
tronic supplementary material, text), the form of sebaceous
gland exudate [25] and the expression of immune-related Figure 1. Number of OTUs per primate species after rarefying to 1000
genes compared with humans [30]. The non-human apes, in sequence reads per individual and square-root transforming. In order to fit
turn, differ from the Old World monkeys (e.g. rhesus macaques the assumptions of an ANOVA, square-root transformations were conducted
and baboons, approx. 30 Myr divergent from the apes) [31] in because of unequal variances. Levene’s Test, one-way ANOVA and Tukey’s
other attributes of similar features. For example, while humans, multiple comparison test were completed on the mean of each host group
chimpanzees and gorillas have apocrine glands clustered in the (F4,62 ¼ 51.3, p , 0.05). Letters above columns correspond to the first
axillae (armpits), in Old World monkeys apocrine glands are letter of the primate host that is significantly different ( p , 0.05).
not clustered and can be found (typically at lower densities)
covering the entire body and in equal number to eccrine
so, we reveal a clear signature of host evolution and biology
glands [26,29] (electronic supplementary material, text).
on the microbiome of the axillae. These results emphasize the
Second, over the last 100 years, human hygienic behaviour has
role of evolutionary relationships in defining the human skin
dramatically changed such that the skin of most humans is
ecosystem, and have implications for the role of skin microbes
now exposed daily to soaps, detergents and underarm products,
in health and disease.
some of which affect skin microbes [16,32,33]. Third, microbes
might differ among primate hosts as a function of their evol-
utionary dissimilarity [34,35], whether as a result of drift or
selection on host traits that influence skin microbes. Collectively, 2. Results and discussion
then, we might predict differences in the skin microbiotas among
primate hosts, including humans, that reflect both the recent (a) Axillary microbial richness
shifts in human hygiene and more ancient divergences in the We obtained ribosomal RNA gene sequences from 63 samples
biology of the skin in addition to host evolutionary history. of primate axillae using high-throughput pyrosequencing
Specifically, we might predict large differences between the (Roche 454), and processed them using the Quantitative
monkey and the ape hosts, and, within the apes, more subtle Insights into Microbial Ecology (QIIME) pipeline [38]. To evalu-
differences between human and non-human apes. ate microbial diversity, QIIME categorizes and groups similar
The skin habitats with high microbial abundance and sequences that share a threshold level of sequence identity
functional consequences (e.g. on mating or social behaviour) into a defined operational taxonomic unit (OTU); it then ident-
are perhaps the most likely to be strongly influenced by host ifies a representative sequence from each OTU group to assign
evolution (akin to the situation in the gut [34]). The axillae taxonomic lineage. Using a 97% nucleotide-sequence identity
have one of the highest bacterial biomasses on the skin sur- threshold to define OTUs, we narrowed our samples to a
face [36,37] and are directly sustained by food resources single axilla sample per individual and rarefied to 1000
primarily provided by apocrine glands [18]. If the influence sequences per individual (see Methods; electronic supplemen-
of evolution is to be apparent on any part of the skin micro- tary material, figure S1). We identified 5309 unique OTUs
biome, it should be in the axillae, especially given the known from our primate axillae samples; 30 OTUs were unclassified
differences in apocrine gland distribution and density among (probably sequencing artefacts), 15 were archaeal in origin
primates (electronic supplementary material, text). and 5264 were bacterial OTUs.
Here we assess the microbial (bacterial and archaeal) com- Using the rarefied dataset, we found that humans had the
munities in the axillae of five primate species. Samples were lowest average richness of microbial OTUs per individual
collected from humans, zoo populations of chimpanzees, gor- (one-way ANOVA and subsequent pairwise testing using
illas and baboons, and semi-free-ranging rhesus macaques. We Tukey’s multiple comparison test: p , 0.05; figure 1). Chim-
used high-throughput pyrosequencing of the 16S rRNA gene panzees and gorillas hosted a similar richness of skin
to identify microbes present in primate axillae. We hypo- microbes and each had a significantly lower richness of
thesized that the composition and diversity of axillary skin microbial OTUs than did rhesus macaques (one-way
microbes would be different for each host species and pre- ANOVA and Tukey’s tests: p , 0.05; figure 1). In addition,
dicted that their composition would track the phylogenetic chimpanzees had significantly higher OTU richness than did
relatedness of primate hosts. Exclusive of studies focused on baboons (one-way ANOVA and Tukey’s tests: p , 0.05;
humans, our study is the first to consider the microbes on pri- figure 1). Furthermore, the axillae of rhesus macaques and
mate skin with high-throughput sequencing methods. In doing baboons had a significantly higher OTU richness than did
those of humans and chimpanzees (one-way ANOVA and microbe composition by host species 3
Tukey’s tests: p , 0.05). In general, the higher OTU richness of 1.50 2D stress = 0.14 PerMANOVA: p = 0.0001

rspb.royalsocietypublishing.org
skin microbes in monkeys relative to apes is in line with expec-
tations if the increased number of axillary apocrine glands in
0.75
apes [29] tends to favour a subset of bacterial taxa at the expense
of diversity (electronic supplementary material, text).
The four most common human skin bacterial phyla—

axis 2
0
Firmicutes, Actinobacteria, Proteobacteria and Bacteroidetes
[8]—comprised over 97% of the sequence reads in the human
axillae samples (electronic supplementary material, figure
–0.75
S2), similar to that observed in studies focused on human
skin microbes [15]. These same phyla made up 85% of the

Proc. R. Soc. B 283: 20152586


sequence reads in chimpanzees and gorillas, 82% in baboons –1.50
and 88% in rhesus macaques (electronic supplementary –1.50 –0.75 0 0.75 1.50
axis 1
material, figure S2).
rhesus macaques
baboons
(b) Composition of axillary microbes among hosts gorillas
To understand the microbial composition among host chimpanzees
humans
primates, we first assessed the underlying compositional data
matrix (sequence count of each microbial genus by individual Figure 2. NMDS plot of the composition of skin microbial communities across all
host) to confirm that there was structure in the dataset using the five primate species. Each point represents an individual host; triangles represent
KR-means with a Simprof significance test in PRIMER-E v. 7.0.8. monkeys (baboons and rhesus macaques), while circles represent apes (gorillas,
We found that there was significant structure in microbial com- chimpanzees and humans). Species are coded with different colours. Larger sym-
position among hosts (R ¼ 0.91, p ¼ 0.01). We then applied a bols represent the centroid for each species + 1 se. The composition of skin
hierarchical approach to assess whether the composition of microbiota on monkeys was significantly different from that of apes (PerMANOVA:
skin microbes differed predictably among primate groups p , 0.01). Within monkeys, baboons and rhesus macaques had significantly
(monkeys and apes), within each group (monkeys: baboons, different skin microbiota (PerMANOVA: p , 0.01). Within apes, gorillas and
rhesus macaques; apes: humans, chimpanzees, gorillas). We chimpanzees had significantly different skin microbiota than humans (PerMA-
then examined the individual variation in skin microbial com- NOVA: p , 0.01 for both comparisons); however, the composition of gorilla
position within host species. Finally, we compared the skin and chimpanzee skin microbiota did not differ significantly from each other (Per-
microbes of human zoo workers with other humans and MANOVA: p ¼ 0.12). (Online version in colour.)
to apes living in zoos to explore proximity transfer of microbes.
We visualized relative composition per host using non-
metric multi-dimensional scaling (NMDS) and employed
PerMANOVA (a permutational ANOVA/MANOVA) and has been shown for gut and saliva microbiomes [35,43]
PermDISP ( permuted dispersion, which tests for homogeneity (electronic supplementary material, figure S3a,b).
of dispersions) to statistically evaluate differences among hosts When we investigated differences among individuals within
(see Methods). primate groups, we found that within monkeys, there were
After demonstrating overall structure by microbial compo- significant differences between the skin microbiota in the axillae
sition of host axillae, we tested primate groups and found of baboons and in those of rhesus macaques (PerMANOVA, p ¼
that the composition of axillary microbes differed significantly 0.01; figure 2). Prevotella contributed to 10% of the total differ-
between monkeys and apes (PerMANOVA, p , 0.01; figure 2). ences between the two monkey hosts, while Staphylococcaceae
These differences were primarily driven by Corynebacterium was responsible for 8% of the total differences between baboons
and a genus of Staphylococcaceae (not further differentiated and rhesus macaques (electronic supplementary material,
using the 16S rRNA segment we sequenced, so hereafter table S1). Prevotella, an anaerobic group of microbes found in
referred to as Staphylococcaceae), both of which were much the mouth, vagina and gut of humans, represented the most
more abundant in apes than in monkeys; both Corynebacteria common taxon (22% of all reads; figure 3, green block) in the axil-
and Staphylococcaceae/Staphylococcus have also been found lae of rhesus macaques. The abundance of Prevotella was greater
in high abundance in studies of human skin microbes in the axillae of rhesus macaques compared with baboons
[15,20,36,39–42]. Together, Corynebacteria and Staphylococca- (one-way ANOVA: p , 0.05; electronic supplementary material,
ceae contributed to more than 37% of differences among apes figure S4). The composition of microbes in baboon axillae
and monkeys (SIMPER analysis; table 1). Interestingly, a was diverse. The most common taxon was Staphylococcaceae,
small number of individual gorillas and chimpanzees had which accounted for 12% of reads (figure 3, red block). In con-
skin microbes that were compositionally more similar to mon- trast to differences in composition within hosts, there were no
keys than was the case for the average gorilla or chimpanzee significant differences in beta diversity (a measure of variability
(figure 2; see circles near triangles). These individual hosts among groups) between baboons (zoo animals) and rhesus
had lower abundances of Corynebacterium than other apes macaques (semi-free-ranging; PermDISP: p ¼ 0.89; figure 4).
(unpaired t-test: p , 0.01), but no significant difference in This is somewhat surprising given that differences in microbial
Staphylococcaceae (unpaired t-test: p . 0.05). Despite such composition have been previously noted between captive and
differences in microbe composition among individuals wild host species [43,44].
within host species, the skin microbes on different host species Within apes (gorillas, chimpanzees and humans), the axil-
generally tracked the evolutionary history of those hosts, just as lary microbial communities of humans were significantly
Table 1. Comparison of the average abundances (number of sequence reads) and percentage contributions of the 10 microbial taxa that contributed the most 4
to differences between apes and monkeys (SIMPER analysis, PRIMER-E v. 7.0.8). Undetermined microbial taxa (each of which is a single OTU) are those that did

rspb.royalsocietypublishing.org
not match sequences for known, named taxa at genus level. In some cases, with longer sequence reads, these taxa (e.g. Staphylococcaceae) might correspond
with known phyla, classes, genera and species, respectively.

average abundance
% contribution to % cumulative
comparison family genus apes monkeys differences contribution

apes versus Staphylococcaceae undetermined 0.37 0.08 20.79 20.79


monkeys Corynebacteriaceae Corynebacterium 0.29 0.04 16.56 37.35

Proc. R. Soc. B 283: 20152586


Prevotellaceae Prevotella 0.02 0.09 4.77 42.11
Clostridiaceae Anaerococcus 0.06 0.01 3.86 45.97
undetermined phylum 0.01 0.07 3.61 49.58
Pasteurellaceae undetermined 0.00 0.03 1.89 51.47
Streptococcaceae Streptococcus 0.00 0.03 1.83 53.30
Ruminococcaceae Ruminococcus 0.01 0.02 1.58 54.88
undetermined Actinomycetales 0.00 0.03 1.41 56.29
undetermined Streptophyta 0.01 0.02 1.33 57.62

100 50
A
AB
(among-host dissimilarity)

80 B
percentage of total reads

40
beta diversity

60
30
40
20
20
10
0
human chimpanzee gorilla baboon rhesus macaque
host primate species 0
human chimpanzee gorilla baboon rhesus macaque
Corynebacterium
apes monkeys
Staphylococcaceae
Figure 4. Beta diversity bar charts showing dissimilarity in microbial content
Prevotella
(based on number of sequences and taxonomic classification) between
Anaerococcus primate species within apes and monkeys. Within apes, humans and gorillas
other microbial species were significantly different (PermDISP: p , 0.05). There was no significant
difference between humans and chimpanzees or chimpanzees and gorillas.
Figure 3. Percentage of total reads of the Staphylococcaceae family, and Within monkeys, there were no significant differences between baboons
genera of Corynebacterium, Anaerococcus and Prevotella in each of the five and rhesus macaques.
host primate species. Columns represent average sequence reads per individ-
ual within host group.
electronic supplementary material, figure S4). Corynebacterium
abundance was significantly higher in humans and chim-
different from those of chimpanzees and gorillas (PerMA- panzees when compared with baboons (one-way ANOVA:
NOVA: p , 0.01 for both). However, there were no significant p , 0.05; electronic supplementary material, figure S4).
differences in the composition of skin microbiota between Humans had the highest level of compositional variability
chimpanzees and gorillas (PerMANOVA: p ¼ 0.12; figure 2; from one individual to the next, in line with the expectation
electronic supplementary material, table S1). Greater abun- if human underarm products have a strong influence on the axil-
dances of Corynebacterium and Staphylococcaceae in humans lary microbiome. Recent studies investigating the effects
than in other apes underlie most of this divergence in axillary of underarm product use found shifts from Corynebacterium to
microbiota between humans and other apes. These two Staphylococcaceae in people who use underarm products such
genera contributed 52% of the total compositional differences as antiperspirant and deodorant [32,33]. Additionally, people
between humans and chimpanzees, and 41% of the total com- who routinely wore no product hosted higher abundances
positional differences between humans and gorillas of Corynebacterium than of Staphylococcaceae [32]. Here, our
(electronic supplementary material, table S1). The abundance results recontextualize those findings and suggest that in the
of Staphylococcaceae was significantly higher in humans com- absence of deodorant and antiperspirant, humans have axillary
pared with all other primate hosts (one-way ANOVA: p , 0.05; microbes more similar to apes (more Corynebacterium, less
Staphylococcaceae), but that by altering the relative abundance of primates, particularly monkeys. Second, humans have a less 5
these taxa, underarm product use makes us less similar to diverse skin microbial community, more dominated by Sta-

rspb.royalsocietypublishing.org
the other apes (electronic supplementary material, table S1). phylococcaceae than is the case for any other primate
Within host species, the composition of microbes tended to sampled. The latter is interesting in as much as Staphylococcus
vary more from one individual to the next within ape species has long been viewed as the ‘normal’ skin symbiont. It is also
versus within monkey species (PermDISP, p , 0.01; electronic one that has been noted to attract mosquito species, including
supplementary material, figure S5). at least one vector of malaria [11]. Our work suggests the
Given that the skin is constantly in contact with the sur- modern composition of our axillary microbiome, including
rounding environment, and in the light of recent evidence of the abundance of Staphylococcaceae, is new at least relative
transfer of microbes during contact sports [45], we investigated to our divergence with other primates. Whether it is new rela-
whether people who came into contact with non-human pri- tive to changes in the last few hundred years remains to be
mates acquired microbes from the animals they worked with. seen. Nonetheless, our findings advance our understanding

Proc. R. Soc. B 283: 20152586


Individuals who worked with non-human primates did not of the evolutionary context of the diversity of human skin
differ from other humans in terms of their axillary microbiome, microbiota, and suggest that phylogenetic relatedness among
nor did they differ from non-human apes (electronic sup- hosts may strongly influence the composition of skin micro-
plementary material, text and figure S6a,b). Thus, if transfer biomes across all primates.
occurs between humans and non-human animals it does not
eclipse the differences intrinsic to host species [43], but it
may be associated with more subtle shifts (hence our inability 4. Methods
to distinguish those with close contact with other primates
compared with none). (a) Sample collection
Samples from both the left and right human axillae of participants
were collected between August 2012 and May 2013. Human volun-
(c) A core primate axillary microbiome teers from the North Carolina Museum of Natural Sciences
Despite the differences among individual hosts and among (NCMNS) and North Carolina State University (NCSU) were
host species, we identified a core primate axillary microbiome asked to refrain from using deodorant or antiperspirant for 2
comprising taxa shared by at least 95% of individuals [46]. days prior to sampling, but were allowed to shower normally. Par-
OTUs of the genus Corynebacterium were found in all primate ticipants were told to swab each axilla with a sterile dual-tipped
individuals sampled. Corynebacterium, Prevotella, Anaerococcus rayon swab (BD BBL Cat no B4320135) for at least 45 s while rotat-
and a genus of Staphylococcaceae were also among the most ing the swab so that all sides of each swab came in contact with the
abundant taxa based on read number (figure 3; electronic axillary skin. In this study, we engaged 20 human participants
from a concurrent study, in addition to 17 collected specifically
supplementary material, figure S7). These taxa are consist-
for this study, for a total of 37 human participants. We collected
ently frequent (found on many individuals) and abundant
samples from non-human primates during routine physicals of
( present as many reads). Meanwhile, the diversity of OTUs baboons, chimpanzees and gorillas at the North Carolina Zoo,
outside of the top four genera increased with evolutionary and rhesus macaques in Cayo Santiago in Puerto Rico. All zoo pri-
distance from humans (figure 3; electronic supplementary mates were captive born except for one chimpanzee born in the
material, figure S7), an effect likely to be due to a complex wild. None of the gorillas were taking antibiotics at the time of
mix of the influence of human hygiene, the evolutionary his- swabbing, but all of the gorillas had been treated within the past
tory of the axillary organ and other host differences that have six months with a dewormer (mebendazole or pyrantel pamoate).
accrued over evolutionary time. Among the most common None of the other non-human primates were being given any
taxa in the primates most evolutionarily distant from medication. All zoo primates had outdoor and soil access, and
humans (in our sampling), the monkeys, were OTUs of ate vegetarian diets. The rhesus macaques lived entirely outdoors
and were rationed with monkey chow. In total, we sampled
Prevotella and Actinomycetales. In addition, the monkeys
11 baboons, seven chimpanzees, five gorillas and two rhesus
have a large number of microbial genera (more than 30
macaques. Swabs were stored at 48C until processed. Human
genera) in their core axillary microbiome that are generally samples were processed within weeks of initial sampling; primate
associated with soil, gut and oral microbiomes such as Acine- samples were processed within months after initial sampling.
tobacter, Ruminococcaceae and Porphyromonas, respectively. It
is unclear whether these taxa live on the skin persistently, or
represent frequent contamination of the skin with soil and (b) DNA extraction and pyrosequencing
faeces (or both). Each armpit was swabbed with a dual-tipped rayon swab. One
of each of these tips was used for qualitative plating for outreach
events at NCMNS; the second was used for DNA extraction for
pyrosequencing (454 Roche Genome Sequencer FLX system).
3. Conclusion DNA was extracted with the PowerSoil DNA Isolation Kit (MO
A large body of recent research has suggested that human BIO Laboratories, Carlsbad, CA, USA) following manufacturer’s
exposure to microbes has changed dramatically over the last protocol with a modification of a two-step elution process into 50
100 or 200 years (and prior to that, with the origin of agriculture total microlitres of elution buffer C6. DNA was stored at – 208 C
until amplified and sequenced. Isolated DNA was PCR amplified
more than 10 000 years ago), with consequent shifts in gut [35]
with 16S rRNA primers (515F and 806R(49)) with 454 adapters
and oral microbes [43], in extreme cases leading researchers
and index sequences using Premix Ex Taq (Takara Bio, Siga,
to the ‘hygiene hypothesis’ [47,48]. Here, we find that the Japan) [38]. PCR amplification was done as follows: initial dena-
shift is variable among individuals, and rather than being turation at 948C for 2 min, five cycles of 948C for 20 s, 538C 30 s,
associated with sweeping changes in microbial fauna, it is 728C for 1 min, then 30 cycles of 948C for 20 s, 558C for 30 s then
associated with two key changes. First, humans seem to be 728C for 1 min followed by a final 728C extension for 7 min and a
less covered with faecal and soil microbes than are other 108C hold. Each host individual had a unique index and PCRs
were run with 3 – 6 ml of isolated microbial DNA in triplicate in (e) Clustering via KR-clustering 6
25 ml reactions with 1.25 mM each primer with a 1X PCR Before conducting statistical analyses of the composition data, we
master mix. PCRs were visualized using gel electrophoresis

rspb.royalsocietypublishing.org
used KR-clustering in PRIMER v. 7.0.8 [52] to identify whether
and then triplicate reactions from each individual were pooled structure existed with regard to the differences in microbe species
and purified using UltraClean-htp 96-well PCR Clean-up kit composition among individual hosts and host species at the
(MO BIO) per manufacturer’s protocol. An equal mass of genus level. We constructed a dissimilarity matrix, in which
purified pooled product, as measured by Qubit dsDNA HS microbial abundances were compared between samples. With
Assay Kit (Invitrogen, Eugene, OR, USA), was added to a this matrix, we conducted the KR-cluster analysis using 100
single tube and then ethanol precipitated to concentrate the mix- restarts and a Simprof significance test with two to five groups
ture. The ethanol-precipitated 16S rRNA pooled mixture was and a ¼ 0.001. Individual hosts clustered more than would be
measured by Qubit and was sent for Roche 454 next-generation expected based on chance. On the basis of this result, we next
pyrosequencing (Selah Genomics, Greenville, SC, USA). analysed the relationship between skin microbial composition
and attributes of hosts, which might explain this clustering.

Proc. R. Soc. B 283: 20152586


(c) Sequence analysis
In addition to the 17 newly collected samples taken from non-
human primates and humans for this study, we included data
from 20 additional human participants collected for a different (f ) Analysis of skin microbe compositional diversity
study [32]. Combining these two datasets yielded 798 818 sequence We compared the composition of skin microbes among the axil-
reads with 698 799 passing quality and length filters. Amplicons of lae of individuals of all five primate hosts; we tested whether the
the V4–V5 region [49] of the 16S rRNA gene were processed and differences among host species in the composition of microbial
analysed using the QIIME pipeline (MacQIIME 1.7.0-20130523; communities were greater than the differences among host indi-
electronic supplementary material, text) [38]. DNA sequences viduals within species using a PerMANOVA. All analyses of
were processed, filtered and assigned to individuals according to composition, including Bray – Curtis dissimilarity, non-metric
a 13 bp barcode, derived from the standard 12 bp barcode [49] multi-dimensional scaling plot, SIMPER analysis and PermDISP,
with one bp added in silico to all samples in the mapping file were conducted using PRIMER-E v. 7.0.8 with the PerMANOVA
and the original fna files in order to combine the Urban et al. ext. v. 1.0.3 [52]. We first constructed a resemblance matrix
samples that used the same 12 bp barcodes. Denoising of samples based on the number of sequence reads of each microbe taxon
was not done due to flow pattern B processing of 454 sequence on each host individual using Bray– Curtis dissimilarities.
data, which enabled longer reads but removed the denoizing capa- Bray – Curtis dissimilarity is a standardized metric of compo-
bility. OTUs were clustered by UCLUST v. 1. 2.22q [40] using a 97% sitional dissimilarity among ecological samples, such as swabs
nucleotide similarity threshold such that reads 3% different (or of axillae or quadrats in old fields. Values range between 0 and
less) from each other were assigned to the same taxon. Taxonomic 1, with 0 indicating no difference among samples and 1 indicat-
identities were aligned using PYNAST v. 1.2 and assigned based ing that two samples do not share any microbial species [53].
off GREENGENES v. 12-10) [50,51]. OTUs of the family Staphylococca- Bray – Curtis dissimilarity has been recommended as a robust
ceae were common in our dataset (more than 50% sequences), but measure of ecological distance for complex communities [54].
could only be defined at the family-level classification. The OTUs Next, we created a NMDS with a Type I Kruskal fit scheme
of the family Staphylococcaceae taxon fell within a single 97% and 100 restarts; NMDS is a preferred approach when differences
de novo OTU, thus we reference them in comparison to other among samples may be due to categorical variables (e.g. host
genus-level OTUs. All samples were rarefied to 1000 sequence species) rather than continuous gradients (as in detrended
reads per sample prior to downstream analyses. Five samples correspondence analysis) [55]. To test our a priori hypothesis
were removed from analysis due to limited read depth. Analysing regarding the causes of differences in microbe composition
53 individuals with both left and right samples, there was no sig- among hosts, we used PerMANOVA with the independent
nificant difference (two-tailed unpaired t-test: p ¼ 0.81) in genus- factor of apes (humans, gorillas, chimpanzees) versus monkeys
level OTU counts in left versus right axilla samples. Because we (baboons, macaques; 9999 iterations, and Type III sums of
found no significant difference between left and right axilla squares). When differences between apes and monkeys were
samples, and some host individuals provided only one axillary significant, we assessed pairwise differences among host species
sample, we analysed a single axillary sample for each individual within apes and within monkeys using PerMANOVA with
in this study (electronic supplementary material, figure S1). For species as a factor, 9999 iterations and Type III sums of squares.
individuals providing both axillary samples, one side was To determine the difference between zoo workers, other humans
chosen based on the highest genus-level OTU value by sample and zoo animals (chimpanzees, gorillas and baboons), we per-
per individual. Data from both axillae are deposited in NCBI as formed PerMANOVA with the independent factor HumanZoo
Bioproject PRJNA281417 (http://www.ncbi.nlm.nih.gov/biopro- (human zoo workers þ humans working with wild primates,
ject/; electronic supplementary material, text). other humans and non-human primates housed in zoos).
We performed SIMPER analysis to determine which OTUs
contributed the most to differences among host groups.
(d) Operational taxonomic unit data processing We assessed the variation in skin microbial composition
All further data analysis used only one axillae sample per partici- among host species using PermDISP. Specifically, we trans-
pant. The homogeneity of variances of host OTUs was assessed formed the data matrix into a presence/absence matrix, and
using Levene’s test (IBM SPSS Statistics for Windows, v. 23.0; then constructed a resemblance matrix using Bray– Curtis dis-
Armonk, NY) to determine whether the data fit the assumption similarities using these transformed values. Using this matrix,
of an ANOVA. After square-root transformation, a one-way we performed a PermDISP using host cluster as the independent
ANOVA and Tukey’s multiple comparison test were used to com- factor, 9999 iterations and centroids as the measure of central ten-
pare the mean of the square root transformed OTUs in each host dency. Here, we were testing whether taxa differed in terms of
group using GraphPad PRISM v. 5.01 (La Jolla, CA, USA). Further how variable composition was among individuals within the
downstream analysis was calculated using QIIME output at the host (e.g. among chimpanzees versus among gorillas). Again,
phylum level (L2) or the genus level (L6). Sequence reads and when differences between apes and monkeys were significant,
abundance datasets are available on FigShare under the manu- we assessed pairwise differences for each species pair within
script title (Figshare.com). apes and monkeys as described above.
of the study and drafted the manuscript; A.M.S. carried out the stat-
(g) Core microbiome istical analyses and drafted the manuscript; M.E.E. collected field
7
Primate core microbe taxa were determined based on the QIIME L6
data; J.M.U. helped analyse data; J.H.P.S. and M.L.P. aided in data

rspb.royalsocietypublishing.org
genus-level output. Microbes found in more than 95% of all indi- collection and manuscript revisions; J.E.H. and R.R.D. conceived of
viduals sampled across all hosts were determined to be in the the study, designed the study, coordinated the study, participated
primate core microbiome [46]. Microbes found in all non-human in data analysis and drafted the manuscript. All authors provided
primate species were termed non-human primate core microbes. valuable insights on the manuscript and gave final approval
Any microbe found in all monkey individuals (baboons and for publication.
rhesus macaques) was considered to be a monkey core microbe. Competing interests. We have no competing interests.
Funding. This project was funded through PI R.R.D. (W911NF-
Ethics. All human participants were provided a written informed con- 14-1-0556 from the Army Research Office) and is part of a
sent form approved by North Carolina State University Institutional larger collaborative grant to PI M.L.P. (NIMH R01-1MH096875-
Review Board for the Protection of Human Subjects in Research (IRB) 01A1).
(approval no. 1987) or North Carolina Central Institutional Review Acknowledgements. We thank Dr Lauren Brent, Dr Richard Bergl, Chris

Proc. R. Soc. B 283: 20152586


Board (approval no. 1201121). Goldston and Jennifer Ireland for sample collection of the non-
Data accessibility. DNA sequences: NCBI as Bioproject PRJNA281417. human primates. We thank Dr Dan Fergus, Dr Holly Menninger
Also on FigShare under the manuscript title (Figshare.com). and Dr Greg Pahel for thoughtful discussions. We thank all of
Authors’ contributions. S.E.C. carried out the molecular laboratory work, the dedicated citizens who participated in our study and who
participated in data and statistical analysis, participated in the design contributed samples and ideas for this project.

References
1. Leider M, Buncke CM. 1954 Physical dimensions of mosquitoes. PLoS ONE 6, e28991. (doi:10.1371/ 21. Archie EA, Theis KR. 2011 Animal behaviour meets
the skin: determination of the specific gravity of journal.pone.0028991) microbial ecology. Anim. Behav. 82, 425–436.
skin, hair, and nail. AMA Arch. Dermatol. Syphilol. 12. McBride CS, Baier F, Omondi AB, Spitzer SA, (doi:10.1016/j.anbehav.2011.05.029)
69, 563–569. (doi:10.1001/archderm.1954. Lutomiah J, Sang R, Ignell R, Vosshall LB. 2014 22. Cernoch JM, Porter RH. 1985 Recognition of
01540170033005) Evolution of mosquito preference for humans linked maternal axillary odors by infants. Child Dev. 56,
2. Goldsmith LA. 1990 My organ is bigger than your to an odorant receptor. Nature 515, 222 –227. 1593– 1598. (doi:10.2307/1130478)
organ. Arch. Dermatol. 126, 301– 302. (doi:10. (doi:10.1038/nature13964) 23. Montagna W, Yun J, Machida H. 1964 The skin
1001/archderm.1990.01670270033005) 13. Smallegange RC, Verhulst NO, Takken W. 2011 of primates XVIII: the skin of the rhesus monkey
3. Christensen G, Brüggemann H. 2014 Bacterial skin Sweaty skin: an invitation to bite? Trends Parasitol. (Macaca mulatta). Am. J. Phys. Anthropol. 22,
commensals and their role as host guardians. 27, 143– 148. (doi:10.1016/j.pt.2010.12.009) 307–319. (doi:10.1002/ajpa.1330220317)
Beneficial Microb. 5, 201–215. (doi:10.3920/ 14. Smallegange RC, Takken W. 2010 Host-seeking 24. Montagna W, Yun JS. 1962 The skin of primates
BM2012.0062) behaviour of mosquitoes: responses to olfactory VIII: the skin of the anubis baboon (Papio doguera).
4. Park B, Iwase T, Liu GY. 2011 Intranasal application stimuli in the laboratory. In Olfaction in vector-host Am. J. Phys. Anthropol. 20, 131 –141. (doi:10.1002/
of S. epidermidis prevents colonization by interactions, vol. 2 (eds W Takken, BGJ Kools), pp. ajpa.1330200214)
methicillin-resistant Staphylococcus aureus in mice. 143 –180. Wageningen, The Netherlands: 25. Montagna W, Yun J. 1963 The skin of primates XV:
PLoS ONE 6, e25880. (doi:10.1371/journal.pone. Wageningen Academic Publishers. the skin of the chimpanzee (Pan satyrus). Am. J.
0025880) 15. Grice EA et al. 2009 Topographical and temporal Phys. Anthropol. 21, 189– 203. (doi:10.1002/ajpa.
5. Chehoud C, Rafail S, Tyldsley AS, Seykora JT, Lambris diversity of the human skin microbiome. Science 1330210211)
JD, Grice EA. 2013 Complement modulates the 324, 1190–1192. (doi:10.1126/science.1171700) 26. Montagna W. 1985 The evolution of human skin.
cutaneous microbiome and inflammatory milieu. 16. Hulcr J, Latimer AM, Henley JB, Rountree NR, Fierer J. Hum. Evol. 14, 3–22. (doi:10.1016/S0047-
Proc. Natl Acad. Sci. USA 110, 15 061 –15 066. N, Lucky A, Lowman MD, Dunn RR. 2012 A jungle 2484(85)80090-7)
(doi:10.1073/pnas.1307855110) in there: bacteria in belly buttons are highly diverse, 27. Ellis RA, Montagna W. 1962 The skin of primates VI:
6. Belkaid Y, Segre JA. 2014 Dialogue between skin but predictable. PLoS ONE 7, e47712. (doi:10.1371/ the skin of the gorilla (Gorilla gorilla). Am. J. Phys.
microbiota and immunity. Science 346, 954–959. journal.pone.0047712) Anthropol. 20, 79–93. (doi:10.1002/ajpa.1330200210)
(doi:10.1126/science.1260144) 17. Kong HH. 2011 Skin microbiome: genomics-based 28. Prüfer K et al. 2012 The bonobo genome compared
7. Tlaskalová-Hogenová H et al. 2004 Commensal insights into the diversity and role of skin microbes. with the chimpanzee and human genomes. Nature
bacteria (normal microflora), mucosal immunity and Trends Mol. Med. 17, 320–328. (doi:10.1016/j. 486, 527–531. (doi:10.1038/nature11128)
chronic inflammatory and autoimmune diseases. molmed.2011.01.013) 29. Folk GEJr, Semken AJr. 1991 The evolution of sweat
Immunol. Lett. 93, 97– 108. (doi:10.1016/j.imlet. 18. Shelley WB, Hurley HJ, Nichols AC. 1953 Axillary glands. Int. J. Biometeorol. 35, 180 –186. (doi:10.
2004.02.005) odor: experimental study of the role of bacteria, 1007/BF01049065)
8. Grice EA, Segre JA. 2011 The skin microbiome. Nat. Rev. apocrine sweat, and deodorants. AMA Arch. 30. Nguyen DH, Hurtado-Ziola N, Gagneux P, Varki A.
Microbiol. 9, 244–253. (doi:10.1038/nrmicro2537) Dermatol. Syphilol. 68, 430–446. (doi:10.1001/ 2006 Loss of Siglec expression on T lymphocytes
9. Theis KR, Schmidt TM, Holekamp KE. 2012 Evidence archderm.1953.01540100070012) during human evolution. Proc. Natl Acad.
for a bacterial mechanism for group-specific social 19. James AG, Austin CJ, Cox DS, Taylor D, Calvert R. Sci. USA 103, 7765–7770. (doi:10.1073/pnas.
odors among hyenas. Sci. Rep. 2, 615. (doi:10.1038/ 2013 Microbiological and biochemical origins of 0510484103)
srep00615) human axillary odour. FEMS Microbiol. Ecol. 83, 31. Perelman P et al. 2011 A molecular phylogeny of
10. Verhulst NO, Beijleveld H, Knols BG, Takken W, Schraa 527 –540. (doi:10.1111/1574-6941.12054) living primates. PLoS Genet. 7, e1001342. (doi:10.
G, Bouwmeester HJ, Smallegange RC. 2009 Cultured 20. Leyden JJ, McGinley KJ, Hölzle E, Labows JN, 1371/journal.pgen.1001342)
skin microbiota attracts malaria mosquitoes. Malar. J. Kligman AM. 1981 The microbiology of the human 32. Urban J, Ehlers M, Fergus D, Menninger H, Dunn R,
8, 302. (doi:10.1186/1475-2875-8-302) axilla and its relationship to axillary odor. J. Invest. Horvath J. In press. Deodorants and antiperspirants
11. Verhulst NO et al. 2011 Composition of human skin Dermatol. 77, 413–416. (doi:10.1111/1523-1747. inhibit growth and modulate species composition in
microbiota affects attractiveness to malaria ep12494624) human axilla. Peer J.
33. Callewaert C, Hutapea P, Van de Wiele T, Boon N. 41. Callewaert C, Kerckhof F, Granitsiotis MS, Van Gele Curr. Allergy Asthma Rep. 13, 487–494. (doi:10. 8
2014 Deodorants and antiperspirants affect the M, Van de Wiele T, Boon N. 2013 Characterization of 1007/s11882-013-0382-8)

rspb.royalsocietypublishing.org
axillary bacterial community. Arch. Dermatol. Res. Staphylococcus and Corynebacterium clusters in the 49. Bates ST, Berg-Lyons D, Caporaso JG, Walters WA,
306, 701–710. (doi:10.1007/s00403-014-1487-1) human axillary region. PLoS ONE 8, e70538. (doi:10. Knight R, Fierer N. 2011 Examining the global
34. Ley RE et al. 2008 Evolution of mammals and their 1371/journal.pone.0070538) distribution of dominant archaeal populations
gut microbes. Science 320, 1647 –1651. (doi:10. 42. Bouslimani A et al. 2015 Molecular cartography of in soil. ISME J. 5, 908–917. (doi:10.1038/ismej.
1126/science.1155725) the human skin surface in 3D. Proc. Natl Acad. Sci. 2010.171)
35. Ochman H, Worobey M, Kuo C, Ndjango JN, Peeters USA 112, E2120–E2129. (doi:10.1073/pnas. 50. DeSantis TZ et al. 2006 Greengenes, a chimera-
M, Hahn BH, Hugenholtz P. 2010 Evolutionary 1424409112) checked 16S rRNA gene database and workbench
relationships of wild hominids recapitulated by gut 43. Li J et al. 2013 The saliva microbiome of Pan and compatible with ARB. Appl. Environ. Microbiol. 72,
microbial communities. PLoS Biol. 8, e1000546. Homo. BMC Microbiol. 13, 204. (doi:10.1186/1471- 5069– 5072. (doi:10.1128/AEM.03006-05)
(doi:10.1371/journal.pbio.1000546) 2180-13-204) 51. Werner JJ, Koren O, Hugenholtz P, DeSantis TZ,

Proc. R. Soc. B 283: 20152586


36. Taylor D, Daulby A, Grimshaw S, James G, Mercer J, 44. Kohl KD, Skopec MM, Dearing MD. 2014 Captivity Walters WA, Caporaso JG, Angenent LT, Knight R,
Vaziri S. 2003 Characterization of the microflora of results in disparate loss of gut microbial diversity in Ley RE. 2012 Impact of training sets on classification
the human axilla. Int. J. Cosmetic Sci. 25, 137–145. closely related hosts. Conserv. Physiol. 2, pcou009. of high-throughput bacterial 16s rRNA gene surveys.
(doi:10.1046/j.1467-2494.2003.00181.x) 45. Meadow JF, Bateman AC, Herkert KM, O’Connor TK, ISME J. 6, 94 –103. (doi:10.1038/ismej.2011.82)
37. McBride ME, Duncan WC, Knox JM. 1977 The Green JL. 2013 Significant changes in the skin 52. Clarke K, Gorley R. 2015 PRIMER v7: user manual/
environment and the microbial ecology of human microbiome mediated by the sport of roller derby. tutorial. Plymouth, UK: PRIMER-E Ltd.
skin. Appl. Environ. Microbiol. 33, 603 –608. Peer J. 1, e53. (doi:10.7717/peerj.53) 53. Somerfield PJ. 2008 Identification of the Bray –
38. Caporaso JG et al. 2010 QIIME allows analysis of 46. Huse SM, Ye Y, Zhou Y, Fodor AA. 2012 A core Curtis similarity index: comment on Yoshioka
high-throughput community sequencing data. Nat. human microbiome as viewed through 16S rRNA (2008). Mar. Ecol. Prog. Ser. 372, 303–306. (doi:10.
Methods 7, 335– 336. (doi:10.1038/nmeth.f.303) sequence clusters. PLoS ONE 7, e34242. (doi:10. 3354/meps07841)
39. Costello EK, Lauber CL, Hamady M, Fierer N, Gordon 1371/journal.pone.0034242) 54. Faith DP, Minchin PR, Belbin L. 1987 Compositional
JI, Knight R. 2009 Bacterial community variation in 47. Strachan DP. 1989 Hay fever, hygiene, and dissimilarity as a robust measure of ecological
human body habitats across space and time. Science household size. Br. Med. J. 299, 1259–1260. distance. Vegetation 69, 57 –68. (doi:10.1007/
326, 1694 –1697. (doi:10.1126/science.1177486) (doi:10.1136/bmj.299.6710.1259) BF00038687)
40. Edgar RC. 2013 UPARSE: highly accurate OTU 48. Bendiks M, Kopp MV. 2013 The relationship 55. Ramette A. 2007 Multivariate analyses in microbial
sequences from microbial amplicon reads. Nat. between advances in understanding the ecology. FEMS Microbiol. Ecol. 62, 142– 160.
Methods 10, 996– 998. (doi:10.1038/nmeth.2604) microbiome and the maturing hygiene hypothesis. (doi:10.1111/j.1574-6941.2007.00375.x)

You might also like