Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Evolving Earth 1 (2023) 100005

Contents lists available at ScienceDirect

Evolving Earth
journal homepage: www.sciencedirect.com/journal/evolving-earth

Ecological polarities of African Miocene apes


Gregory J. Retallack
Department of Earth Sciences, University of Oregon, Eugene, OR, 97403-1272, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Humans have been considered ecologically unspecialized, and our evolution a compromise path through a maze
Miocene of conflicting influences. Generalist ecological roles contrast with ecological polarities such as competitor, breeder
Kenya or tolerator. Ecological polarities can be approximated in fossil mammals by relative size of canines for com-
Proconsul
petitors, incisors for breeders, and molars for tolerators. Considered in this way, early Miocene apes of Kenya were
Ekembo
Kenyapithecus
generalists, but show a greater range of ecological polarity than modern apes, including ecological polarities
Victoriapithecus found in living apes, macaques, and vervets. As many as six primates in a single Miocene aleosol show diverse
ecological polarities, implying competitive exclusion. Middle Miocene monkeys and apes were more molarized
and marginally more tolerant than early Miocene primates, so more like humans in that respect. This adaptive
shift of 15 Ma was at a time of climatic aridity and open vegetation indicated by associated paleosols. A 20 m.yr
record of Kenyan paleoprecipitation from paleosols indicates that 15 Ma was unusually dry after exceptionally wet
paleoclimate of 16 Ma. This new Kenyan paleosol record of paleoclimate is from the same localities as the fossil
apes, and uninfluenced by whole ocean mixing, salinity and ice volume effects which compromise similar marine
isotopic proxies of global change. Molarization of apes at 15 Ma ago was only one of a series of adaptations
selected by at least 9 dry alternating with wet episodes which shifted forest to shrubland ecotones over the past
20 m.yr. Our evolutionary lineage ran a gauntlet of Neogene climatic and vegetation changes in Africa by
adopting a generalist ecological role.

1. Introduction here is “ecological polarities”, for axes of relative proportional size of


adaptive features (Retallack, 2004). The term polarity avoids anthropo-
1.1. Paleoecological question morphism of “strategy” and simple linear connotation of “dimension.”
This is not the same as evolutionary polarity in the sense of relative
Humans and other primates have long been considered ecologically ordering of primitive or derived, apomorphic or plesiomorphic charac-
unspecialized (Simpson, 1949), and this study aims to quantitatively ters (Stevens, 1980; Bittrich and Struck, 1989), although Andews (1982)
evaluate that view. Human limbs retain a pentadactyl plan, rather than has used “ecological polarity” in that sense of evolutionary polarity. Nor
elongated limbs of horses or the flippers of whales. Humans do not have is it the same as developmental or cell polarity, such as differentiation of
huge canines for threat or predation, like sabre-tooth cats and other anterior or posterior (Kropf, 1992; Campanale et al., 2017). The aim of
competitors. Nor are humans adapted to resource scarcity, like camels, ecological polarity studies is to quantify adaptation to particular niches.
and other tolerators. Nor are humans as fecund as rodents, or other Other animals may have been at a competitive, reproductive, or tolerance
breeders. Competitors, tolerators and breeders are three evolutionary disadvantage during extremes of global change, but our generalist line-
polarities, or roles toward which organisms may specialize (Retallack, age may have persisted by virtue of flexibility to cope with spatial and
2004). These observations can be quantified by simple measurements temporal disturbances at small population sizes (Potts, 1996).
and plotting within a triangular diagram for proportional representation This study quantifies and tests such ecological generalizations for
of these extremes, with specialist species out toward the apices and early and middle Miocene primates of Kenya (Fleagle and Kay, 1985;
generalist near the middle (Fig. 1). McNulty et al., 2015; Almecija et al., 2021), and a variety of modern
Competitor, tolerator and breeder ecological roles have been widely primates (Fleagle, 2013). These extinct apes are largely known from
quantified by plant ecologists as evolutionary “strategies” (Grime, 1979) fossil teeth. A previously published morphospace (Retallack, 2004)
and “dimensions” (Westoby et al., 2002), but the terminology preferred chosen to approximate mammalian ecological polarities was the relative

E-mail address: gregr@uoregon.edu.

https://doi.org/10.1016/j.eve.2023.100005
Received 15 June 2023; Received in revised form 3 August 2023; Accepted 7 August 2023
2950-1172/© 2023 The Author. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
G.J. Retallack Evolving Earth 1 (2023) 100005

incisorized (lower left hand corner), sabre-tooth cats more than 50%
caninized (upper corner) and ungulates more than 50% molarized (lower
left hand corner).
While differences in incisorization, molarization and caninization are
significant across all mammals (Retallack, 2004), does that extend also to
the lesser variation in primates? Modern and fossil apes, monkeys and
lemurs occupy the central triangle of generalists (Retallack, 2004), in a
clear visualization of common wisdom (Simpson, 1949). The teeth of
primates are ecologically unspecialized compared with other mammals,
so this study quantifies limited primate variation in tooth proportions.
Incisorization in primates has been regarded as a “sclerocarp syndrome”
for feeding on hard shelled fruits (Kinzey, 1992; McCrossin and Benefit,
1997; Ward et al., 1997; Kay, 2010), which may parallel the toughness,
small size, and high nutritional value of seeds fostering fecundity in ro-
dents (Retallack, 2004). Molarization in primates has been linked to
declining encephalization and less complex behavior in primates, but not
without exceptions (Jimenez-Arenas et al., 2014). Some primates have
molarized premolars (Selig and Silcox, 2022), but premolars were not
measured in the present work for consistency with prior work (Retallack,
2004). Caninization in primates is primarily linked with sexual dimor-
phism and agonistic behavior during mate isolation and selection
Fig. 1. Ecological polarity concept.
(Thoren et al., 2006), but female canine size also increases with male
canine size (Plavcan et al., 1995; Plavcan, 1998). This can be a problem
emphasis of functional areas of their three main tooth types, canines, for evaluating fossils depending on the proportion of male or female
incisors and molars. Simple measures of upper tooth dimensions were canines measured, although sex has been interpreted for many fossil
used to compute the functional area of molars, incisors and canines, and specimens (Pickford, 1985; Kelley, 1986, 1995; Pickford et al., 1999).
the proportion of the sum of the three tooth areas for each tooth type
(Fig. 2). Primate canine size is significantly correlated with both intra-
specific and interspecific aggression (Kay et al., 1988; Manning and 1.2. Geological background
Chamberlain, 1993; Kappeler, 1996). Mammalian incisor proportions
show significant correlation with fecundity measured as offspring per This study is limited to data from around Lake Victoria for the early-
year (Retallack, 2004). Mammalian molar occlusal area is strongly middle Miocene Nyanza Rift and Lake Victoria Basin (Fig. 3), which has
correlated with body mass, home range size, basal metabolic rate, life long yielded important Miocene primate fossils (Leakey, 1952; Pickford,
span, gestation time and age at first birth, and inversely correlated with 1986a; Tuttle, 2006). These localities are on a subsidiary rift to the main
fecundity and population density (Martin, 1990; Janis and Carrano, East African or Gregory Rift, and were characterized by terrestrial
1992). The ecological polarity approach of ratios of areas factors out carbonatite-nephelinite volcanism, with a few lacustrine horizons
body mass, so that relative molar size represents metabolic and popula- (Pickford, 1986b, 1995; Retallack, 1991; Bestland and Krull, 1999;
tion characters of tolerator polarity (Retallack, 2004). Each species can Lukens et al., 2017; Michel et al., 2014, 2020; Peppe et al., 2023).
be plotted as a point with a standard deviation envelope within a trian- Particular paleosols were named by descriptive terms in local lan-
gular diagram with apices of 100% canines, 100% molars and 100% guages, and characterized chemically and petrographically by Retallack
incisors (Fig. 1). Within this convention, rodents are more than 50% (1991, Retallack et al., 1995, 2002). Choka and Tut paleosols at Songhor

Fig. 2. Measurements and formulae for calculating ecological polarity of mammals demonstrated on skull of Ekembo heseloni.

2
G.J. Retallack Evolving Earth 1 (2023) 100005

Fig. 3. Geological map of the Nyanza Rift, southwest Kenya (adapted from Retallack, 1991).

and Kwar paleosols at Koru are all roughly equivalent in age, 20.3  0.5 2. Materials and methods
Ma by corrected K–Ar method for bracketing tuffs (Supplementary In-
formation Table S1), and represent lowland and volcanic flank paleo- 2.1. Taxonomic basis
environments respectively (Bishop et al., 1969; Retallack, 1991). On
Rusinga Island, Chido paleosols in the Kiahera Formation some 18 Ma in Taxonomy of Kenyan Miocene apes used here generally follows
age are stratigraphically lower than Tek paleosols in the Hiwegi For- Harrison (2002, 2010), Ward and Duren (2002), Pickford and Kunimatsu
mation some 17.8  0.16 Ma by K–Ar (Drake et al., 1988; Retallack et al., (2005), and Almecija et al. (2021). Only those taxa that could be linked to
1995; Bestland and Krull, 1999). Yom paleosols of a former alluvial specific paleosols and sites (Pickford, 1986a; Retallack, 1991) were
lowland (dambo) on Maboko Island are beneath a tuff dated by 40Ar/39Ar included in this analysis (Table 1), other specimens from other sites or
at 14.71  0.16 Ma (Feibel and Brown, 1991; Retallack et al., 2002). subsequently described without precise locality data are not considered.
Chogo and Onuria paleosols at Fort Ternan formed on volcanic slopes at Victoriapithecus macinnesi is retained as a valid species (Benefit, 1987,
about 13.7  0.3 Ma by 40Ar/39Ar (Bishop et al., 1969; Pickford et al., 1993; Benefit and McCrossin, 1997; Miller et al., 2009), rather than
2006). By mid-late Miocene the subsidence and sedimentation in the emended to the genus Prohylobates (Pickford and Kunimatsu, 2005).
Nyanza Rift was waning (Pickford, 1986a, 1986b), but paleoenvir- Ugandapithecus major is accepted in place of Proconsul major (Le Gros
onmental time series of radiometrically dated paleosols can be extended Clark and Leakey, 1951; Harrison, 2010), because its large postrcrania at
to the present in the Gregory Rift: in the Tugen Hills of central Kenya for Songhor and in Ugandan localities are distinct from those known in
the mid-late Miocene (Jacobs and Deino, 1996; Deino et al., 1990; Proconsul (Senut, 2016, Senut et al., 2000; Gommery et al. (2002).
Behrensmeyer et al., 2002; Pickford and Kunimatsu, 2005; Senut et al., Ekembo nyanzae and E. heseloni are used here instead of Proconsul,
2000), and the Turkana Basin of northern Kenya for the late Miocene to although that genus is retained for Proconsul africanus and P. meswae
Quaternary (Wynn, 2001, 2004a, 2004b). (McNulty et al., 2015). Equatorius africanus is used here, rather than the
genus Kenyapithecus (following Ward et al., 1999; Kelley et al., 2002,

Table 1
Tooth measurements of selected Miocene taxa and localities from Kenya.
Taxon Locality Canines Incisors Molars Source

area (mm2) error no. area (mm2) error no. area (mm2) error no.

Dendropithecus macinnesi Hiwegi, Rusinga Is. 32.72 13.47 9 31.36 8.91 12 49.92 7.6 13 Andrews (1978)
Dendropithecus macinnesi Songhor 33.61 7.84 4 32.24 12.9 5 50.4 0 1 Andrews (1978)
Equatorius africanus Maboko 53.47 12.17 4 43.54 21.61 4 111.02 19.9 4 Pickford (1985)
Kenyapithecus wickeri Fort Ternan 78.23 5.2 2 56.31 29.85 2 119.4 15.1 3 Pickford (1985)
Limnopithecus legetet Songhor 17.89 3.11 4 27.46 5.04 8 35.48 4.16 18 Andrews (1978)
Proconsul africanus Koru 57.61 6.67 2 48.65 19.3 2 80.57 14.8 9 Andrews (1978)
Ekembo heseloni Kiahera, Rusinga Is. 42.05 12.5 5 54.24 15.5 10 82.4 14 32 Andrews (1978)
Ekembo nyanzae Hiwegi, Rusinga Is. 31.71 1.62 2 84.1 24.1 14 131.16 28.1 16 Andrews (1978)
Rangwapithecus gordoni Songhor 41.25 1 0 60.9 13.04 4 83.03 18.4 12 Andrews (1978)
Ugandapithecus major Koru 159.71 34 2 92.43 39.98 3 162.26 44.5 16 Andrews (1978)
Victoriapithecus macinnesi Maboko 17.64 4.79 159 35.27 17.31 130 48.46 9.42 339 Benefit (1987)

3
G.J. Retallack Evolving Earth 1 (2023) 100005

2008; Harrison, 2010), or the Eurasian genus Griphopithecus, contrary to Ma) by Jacobs (200a, b) represent unusually humid conditions for the
Begun (2007), Benefit and McCrossin (2000) and Pickford and Kuni- region (Supplementary Information Table S1).
matsu, 2005). Less controversial are the taxonomy of Dendropithecus
(Andrews and Simons, 1977), Limnopithecus (Harrison, 2002, 2010; Cote 3. Results
et al., 2016), and Rangwapithecus (Hill et al., 2013; Cote et al., 2014).
3.1. Ecological polarities of Miocene apes and monkeys
2.2. Dental measurements of fossil apes
Ecological polarity analysis gives insight into the puzzling diversity of
Dental measurements and formulae used to approximate ecological Kenyan Miocene apes. There are only eight living species of great apes
polarity are shown in Table 1 and Fig. 2, with full data in Supplementary (Hominidae) today, with largely disjunct ranges (Fleagle, 2013), yet
Information Table S2. Only those remains from single collections and single paleosols such as those at the 20 Ma fossil site of Songhor contain
known paleosols that represent attritional assemblages were used as many as 6 species of sympatric stem-apes (Retallack, 1991). These
(Retallack, 1991). The measurements are for individual teeth, both iso- assemblages are attritional accumulations of bones broken and weath-
lated and within complete mandibles or maxillae. There has been no ered during soil formation (Pickford, 1986b), and so represent former
adjustment to relative representation of teeth for a particular dental communities, rather than regionally homogenized assemblages found in
formula: only relative sizes of the three tooth categories. Paleosol loca- lake or stream deposits (Verhaegen et al., 2002). Their paleosols also
tions of these specimens in the Kenyan National Museum were estab- show little evidence of gleization due to waterlogging (Retallack, 1991;
lished from museum records, and discussions and correspondence with Retallack et al., 2002), and no support for “aquarboreal human ances-
collectors (Pickford, 1986a, 1986b). tors” (Verhaegen et al., 2002).
Each ape species plots in different regions of the generalist central
2.3. Paleosol evidence of paleoclimate and ecosystems triangle of incisor-canine-molar morphospace (Fig. 4). Ugandapithecus
major, for example, shows greater caninization than the other species and
Ecological polarity analysis of paleosols followed methods described is most similar in that respect to living orangutans (Pongo pygmaeus). The
elsewhere (Retallack, 2004), and included measurement of soil fertility large size and limbs of Ugandapithecus major appear adapted for slow,
(100 times the molar ratio of bases, MgO, CaO, Na2O and K2O, to Al2O3), suspensory climbing (Gommery et al., 2002; Senut, 2016), again similar
of soil youth (100 minus soil development index out of 100) and of to orangutans. Other taxa (Dendropithecus macinnesi, Proconsul africanus)
hardship, such as aridity indicated by shallow depth of calcic horizon have caninization more like living macaques, and yet other taxa (Lim-
(1000 divided by depth to Bk in cm). nopithecus legetet, Rangwapithecus gordoni) are more like living vervets in
Depth to carbonate horizon (D in cm) in soils is related to mean this respect. These pairs may be further divided by their postcranial
annual precipitation (P in mm) by equation (1) below with R2 ¼ 0.52, anatomy, which indicate that Dendropithecus macinnesi was an under-
S.E. ¼  147 mm. p ¼ 0.00001 (Retallack, 2005). This depth can be branch suspensory feeder, whereas Proconsul africanus, and Rangwapi-
corrected for compaction due to overlying sediment using geological thecus gordoni were overbranch climbers and feeders (Andrews and
estimates of overburden and standard formulae (Sheldon and Retallack, Simons, 1977; Walker and Shipman, 2005; Harrison, 2010; Fleagle,
2001). Also related to mean annual precipitation (P) is nutrient base 2013). Limnopithecus may also have been suspensory, but its postcrania
content (C–– Al2O3/(Al2O3þCaO þ MgO þ Na2O) in moles) of soil Bt are not unambiguously known (Harrison, 2010). Ecological polarity
horizons by equation (2) below with R2 ¼ 0.72, S.E. ¼  182 mm, p ¼ analysis thus confirms that Miocene apes were not only similar to modern
0.00001 (Sheldon et al., 2002). great apes, but to modern monkeys as well (Andrews, 1996; Almecija
et al., 2021). The taxonomic diversity of Kenyan Miocene apes was
P ¼ 137.24 þ 6.45D – 0.013D2 equation 1 matched by their ecological diversity, exploiting a variety of forest tiers,
0.197C branch sizes, fruits, and leaves.
P ¼ 221.12.e equation 2
Modern Homo sapiens straddles the 50% molarization line like Ekembo
Relevant Bk and geochemical data extracted from previous publica- heseloni of 18 Ma on Rusinga Island, and both the ape Kenyapithecus
tions is tabulated in Supplementary Information Table S1, Kenyan africanus and early monkey Victoriapithecus macinnesi of 15 Ma on
Cenozoic paleosol paleoclimatic data. A compilation of compaction cor- Maboko Island (Fig. 4). In contrast, generalist and apelike adaptations are
rected depth to Bk in paleosols as a paleoprecipitation proxy is from occupied by Ekembo nyanzae and Dendropithecus macinnesi of 18 Ma on
Retallack (1991, 2013), Wynn (2001, 2004a, 2004b), Wynn and Retal- Rusinga Island and Kenyapithecus wickeri of 13.7 Ma at Fort Ternan. The
lack (2001) and Retallack et al. (2002). Relevant Kenyan paleosol morphospace occupied by Dendropithecus macinnesi was unchanged from
chemical data is from Retallack (1991), Retallack et al. (1995, 2002), and 20 to 18 Ma, but change in other lineages toward the tolerator pole
Wynn and Retallack (2001). None of the Kenyan paleosol samples was implies environmental hardship at 18 Ma and again at 15 Ma. Tough-
pretreated to remove carbonate as recommended by Michel et al. (2014), coated fruits are indicated by large procumbent incisors in 18 Ma
so only samples with less than 5% CaO were used to calculate Ekembo nyanzae and 15 Ma Equatorius africanus, as well as Turkanapi-
paleoprecipitation. thecus and Afropithecus from 18 Ma deposits in northern Kenya (Leakey
Chemical weathering also alters the mineral content of soils, espe- et al., 1988a, 1988b; Boschetto et al., 1992; McCrossin and Benefit, 1997;
cially their clay minerals, which begin as smectites, and then lose cationic Leakey and Walker, 1997; McNulty et al., 2015). Hard or fibrous foods
bases with further chemical weathering to become kaolinite. This indi- also are indicated by thick enamel and procumbent incisors of Equatorius
cation of paleoprecipitation works best with non-calcareous soils, which africanus and K. wickeri (Martin, 1985). Tough coated fruits and gourds
are found in climates receiving more than 1000 mm mean annual pre- are also indicated by 18 Ma seeds of Annonaspermum (custard apple) and
cipitation (Sheldon et al., 2002). In Africa today, smectite is dominant in Lagenaria (calabash) on Rusinga Island (Collinson et al. (2009) and 13.7
soils receiving less than 1200 mm mean annual precipitation, and Ma seeds of Annonaspermum (custard apple) at Fort Ternan (Retallack,
kaolinite dominant in wetter climates (Mizota et al., 1988; Dworkin, 1992). Furthermore, both Equatorius africanus and Victoriapithecus mac-
2023). Thus, non-calcareous smectitic soils also define a limited paleo- innesi have postcranial anatomy indicative of facility for ground-walking
climatic window of between 1300 and 1000 mm mean annual precipi- like that of macaques and vervets, respectively (McCrossin et al., 1998).
tation, which can be inferred from clay mineral data of Retallack (1991) These adaptations may have been selected by open and fibrous vegeta-
and Behrensmeyer et al. (2002). tion at 18 and 15 Ma, unlike more mesic vegetation of 20 and 16 Ma.
Paleoprecipitation estimated from foliar physiognomy of three leaf
fossil localities, Kabarsero (12.6 Ma), Waril (8.6 Ma) and Kapturo (6.8

4
G.J. Retallack Evolving Earth 1 (2023) 100005

Fig. 4. Ecological polarity of selected species of Miocene apes from Kenya (C–G) and of their Miocene paleosols (A), as well as of comparable modern primates (B).
Envelopes around the point for each species represents the standard deviation of that measure. These measurements are not for complete specimens but for known
teeth. A generalist ecological role is indicated for fossil and modern primates, but deviations toward molarization are accompanied by evidence of aridity in asso-
ciated paleosols.

5
G.J. Retallack Evolving Earth 1 (2023) 100005

3.2. Ecological polarity of paleosols from subhumid to semiarid climate (Fig. 4). Miocene apes and monkeys
adapted to aridification of their ecosystems and soils in different ways,
Confirmation of environmental changes at these times comes from with apes remaining within woodland soils, but some monkeys abundant
paleosols which entomb these fossil primates. The ecological polarity of within grassland soils.
soils and paleosols can be approximated by three measures: (1) soil
fertility (molar ratio of bases to alumina), (2) soil youth (inverse soil 4. Discussion
development index) and (3) soil hardship such as aridity (shallow depth
of soil carbonates). Competitor plants and animals are on the most fertile 4.1. Extended paleoenvironmental time series for Africa
soils, breeder plants and animals on youthful disturbed soils, and toler-
ator plants and animals on soils presenting hardships of drought or low The time series of early to middle Miocene paleosols from the Nyanza
nutrition (Retallack, 2004). Weakly developed soils have bedding planes Rift and Victoria Lake Basin can be extended with comparable studies in
and other original structures little disturbed by roots and weathering, but the mid-late Miocene Tugen Hills (Jacobs and Deino, 1996; Deino et al.,
as time progresses powdery carbonate accumulates and coalesces into 1990; Behrensmeyer et al., 2002; Pickford and Kunimatsu, 2005; Senut
nodules of increasing size (Retallack, 2005) in a quantitative scale of et al., 2001), and the Miocene to Quaternary Turkana Basin (Wynn, 2001,
development which serves as a breeder polarity proxy (Retallack, 2004). 2004a, 2004b; Liutkus-Pierce et al., 2019). This extended compilation
Tolerator polarity in calcareous soils can be inferred from depth within (Fig. 6) includes data on Kenyan paleosol depth to carbonate, chemical
paleosol profiles to carbonate nodules, which correlates significantly and clay mineral composition. Depth to carbonate (Bk) increases with
with mean annual precipitation (Retallack, 2005), but needs to be cor- mean annual precipitation in soils (Retallack, 2005), and so does the
rected for burial compaction in paleosols (Sheldon and Retallack, 2001). degree of depletion of alkali and alkaline earth elements in subsurface
Competitor polarity can be quantified by measures of chemical soil clayey (Bt) horizons of soils (Sheldon et al., 2002). Precipitation also
fertility, such as bases/alumina ratios of non-calcareous parts of the soil. controls clay minerals of soil Bt horizons: smectite in dry regions (less
The four major soil bases are important nutrients: Ca2þ for plant cell than 1200 mm mean annual precipitation) and kaolinite in wetter re-
walls and animal bones, Mg2þ central to chlorophyll; and Naþ and Kþ as gions (Mizota et al., 1988; Dworkin, 2023). Also included in this paleo-
cytoplasmic electrolytes (Retallack, 2019). Such information from pale- climatic compilation (Fig. 6) are published inferences from size, shape
osols with fossil primates reveals selection pressures on their evolution. and features such as drip tips of fossil leaves (Retallack, 1992; Dugas and
There was clear ecological separation of some species to different Retallack, 1993; Jacobs and Deino, 1996; Jacobs, 2002a, 2002b; Michel
paleosols, but still high diversity within some paleosols (Fig. 5). In the 20 et al., 2014, 2020), which agree with estimates from paleosols where
Ma sites of Koru and Songhor, Ugandapithecus shows little habitat spec- both are found in close proximity. Non-calcareous paleosol mineral
ificity through a variety of tropical dry forest paleosols, but small taxa composition and fossil leaf physiognomy both define isolated peaks in
(Kalepithecus, Micropithecus) are in upland soils and larger taxa (Proconsul, paleoprecipitation, compared with the runs of numerous calcareous
Rangwapithecus, Dendropithecus, Nyanzapithecus) remained in lowland paleosols (Fig. 6). Some of the volatility in this compilation represents
forests close to water. One paleosol type (early successional Kiewo lateral landscape heterogeneity (Fig. 5), revealed in more detail by recent
pedotype: Fig. 5) has six taxa: including likely suspensory feeders (Lim- studies (Michel et al., 2020; Peppe et al., 2023; MacLatchy et al., 2023).
nopithecus, Dendropithecus, Ugandapithecus, from small to large) and likely Vegetation can also be reconstructed from known precipitation
overbranch feeders (Nyanzapithecus, Proconsul, Rangwapithecus, from ranges of modern African ecosystems (Anhuf et al., 1999; Sankaran et al.,
small to large). Thus there was niche partitioning of forest canopy tiers. 2005). This assumption fails for newly evolved ecosystems such as
Diverse sympatric catarrhine communities persisted into the drier grasslands, which can be detected in paleosols from fine root traces and
woodland landscapes of Rusinga Island at 18 Ma, when paleosols with crumb ped structure, as well as from associated grass leaves, pollen, and
the crumb peds and iron–manganese nodules of dambo grasslands (Yom phytoliths (Retallack, 1992; Dugas and Retallack, 1993; Peppe et al.,
pedotype: Fig. 5) appear, but are rare and barren of primate fossils 2023). The changing paleoclimatic range of grasslands has been recon-
(Retallack et al., 1995; Peppe et al., 2023). Other evidence for grasslands structed from climatically sensitive features such as depth to Bk in the
of about the same age are abundant bunch grasses at Ugandan fossil sites same paleosols (Fig. 6). Grasslands and their soils (Mollisols) of uplands
of Bukwa (Pickford, 2002) and Moroto (MacLatchy et al., 2023). Yom (veld) and lowlands (dambo) were a newly coevolved ecosystem engi-
paleosols of dambo grassland are common by 15 Ma on Maboko Island neered by grasses and grazers (Retallack, 2013). The expanded climatic
(Retallack et al., 2002), and contain abundant vervet-like monkeys range of mollic paleosols of grasslands through time (Retallack, 2013) is
(Victoriapithecus: note change of scale for this exceptional collection in revealed by filled symbols in Fig. 6B. In contrast early successional and
Fig. 5). Seasonally inundated grasslands of dry climates were not forest communities had considerable antiquity in Africa, especially a dry
encouraging to fossil apes (Lukens et al., 2017), which were more com- woodland community known as pori which was largely displaced by
mon in riparian woodland paleosols (Nyanzapithecus, Limnopithecus, grasslands in regions between 300 and 500 mm of mean annual precip-
Mabokopithecus and Simiolus of Dhero pedotype). In contrast, Equatorius itation (Retallack, 2007, 2012).
at Maboko and Majiwa was wide ranging in both grassland (Yom, Dhero) The extended paleoclimatic compilation reveals not just one Neogene
and pori woodland (Ratong), which it exploited more effectively than aridification and grassland expansion at about 7 Ma, as has long been
other apes because of its thick enamelled, large molars useful for tough implied by the Miocene pluvial (Leakey, 1952), lake (Kent, 1944), and
foods (Martin, 1985), and its macaque-like limbs and feet (McCrossin rain forest (Andrews, 1996) hypotheses, and late Miocene grassland
et al., 1998). hypothesis (Cerling, 1992; Cerling et al., 1997). These theories had
A similar pattern of forest dependence for Kenyapithecus and other already been discredited by discovery of Miocene desert dunes, alkaline
apes (Simiolus, “Proconsul”, nyanzapithecine: Harrison, 1992) persisted lakes, grasses, grazing mammals, and grassland paleosols in East Africa
within grassland woodland mosaics of Fort Ternan and Kapsibor at 13.7 (Pickford, 1986b, 1995, 2002; Retallack, 1991; Retallack et al., 2002;
Ma, when well-drained short-grass, wooded grassland was widespread Lukens et al., 2017; Peppe et al., 2023; MacLatchy et al., 2023). Instead
(Kappelman, 1991; Retallack, 1992; Dugas and Retallack, 1993). The the new data (Figs. 5–6) reveal a mosaic of habitats and a Neogene pa-
appearance of grasslands so encouraging for victoriapithecine ancestors leoclimatic roller coaster of at least 9 desertifications, of which those at
of vervets, was not so encouraging to apes, which remained rare com- 18 and 15 Ma were among the driest. Of the intervening wet forests,
ponents of the fossil fauna. humidity spikes of 16 and 13 Ma were among the wettest.
Analysis of fossiliferous paleosols at these Kenyan sites demonstrate The relationship between Miocene and living African apes are un-
that molarization increases in apes and monkeys corresponds with times clear, because of the extremely sparse fossil record of chimpanzees
of shallow soil carbonate (Fig. 5), which indicates a paleoclimatic shift (McBrearty and Jablonski, 2005) and gorillas (Suwa et al., 2007). The

6
G.J. Retallack Evolving Earth 1 (2023) 100005

Fig. 5. Paleosols of Miocene primates from southwestern Kenya [data from Retallack (1991) and Pickford (1986a, b), with taxonomy after Harrison (2002, 2010),
Retallack et al. (2002), Ward and Duren (2002), Almecija et al. (2021), for selected stratigraphic levels.

7
G.J. Retallack Evolving Earth 1 (2023) 100005

Fig. 6. Compilation of paleosol data for 20 million


years of vegetation (A) and precipitation (B) variation
in Kenya, compared with carbon (C) and oxygen (D)
isotopic composition of marine foraminifera (Zachos
et al., 2001). Paleoprecipitation data from non-mollic
paleosols (B) is from depth to carbonate (open cir-
cles), clay mineralogy (diamonds), and chemical
composition (squares). Closed circles are those mollic
paleosols with carbonate as well as crumb peds and
fine root traces of grassland soils. Also included are
inferences from fossil floras (open triangles: Jacobs,
2002a, 2002b, Jacobs and Deino, 1996; Collinson
et al., 2009; Michel et al., 2014, 2020) and clay
mineral content (Retallack, 1991; Behrensmeyer
et al., 2002). Precipitation ranges of extant vegetation
types are from Holocene reconstructions (Anhuf et al.,
1999; Sankaran et al., 2005), and expansion of
grasslands from trace fossil and structural studies of
paleosols (Retallack, 2013).

varied Miocene apes of southwest Kenya (Harrison, 2010) may have been australopithecines and humans of our lineage (Almecija et al., 2021)
ancestral to erect-walking hominid fossils, such as Sahelanthropus (Brunet exploited C4 and C3 grassland mosaics of Mollisols, Alfisols, and Incep-
et al., 2002), Orrorin (Senut et al., 2001), and Ardipithecus (White et al., tisols after 4.2 Ma (Wynn, 2004a, 2004b; Wynn et al., 2006).
2009; Lovejoy et al., 2009), which were still in woodland and forest
Alfisols from 6.5 to 4.2 Ma (Retallack, 2007). Starting with Austral- 4.2. Local or global paleoclimatic change
opithecus anamensis our lineage sometimes ventured out onto grassland
Mollisols (Wynn, 2001; White et al., 2006), that had long supported an The extended paleoclimatic compilation is similar to the oxygen
adaptive radiation of baboons and vervets (Frost, 2007). Subsequent isotopic record from the deep-sea (Zachos et al., 2001), which has been

8
G.J. Retallack Evolving Earth 1 (2023) 100005

used as a paleoclimatic proxy for evaluating human evolution in Africa Declaration of interest statement
(de Menocal, 2004), but the match is not precise (Fig. 6B and C). A trend
of extreme and volatile middle Miocene values for paleoprecipitation and The author declares no conflict of interest in publishing this research.
marine stable isotopic composition, but subdued late Miocene to Qua-
ternary values is evident. The paleosol record reveals much greater Data availability
variation in rainfall than would be inferred from carbon isotopic
composition of marine foraminifera (Fig. 6C), which is damped by global Data will be made available on request.
oceanic mixing with time lags of several thousand years (Zachos et al.,
2001). More profound damping is seen in oxygen isotopic composition of Acknowledgements
marine foraminifera, which show a long-term increase in δ13C (‰) unlike
local rainfall and foraminiferal carbon records. This increase is plotted on Fieldwork in 1984 and 1987 was undertaken with a permission letter
reversed axes in Fig. 6D because it has been interpreted as a record of from Ms M.A. Mwango of the Office of the President of Kenya, and R.E.
long-term temperature decline (Zachos et al., 2001; Veizer et al., 2001), Leakey of the Kenyan National Museum. Funding came from Wenner
but is partly an artefact of long-term water recycling due to plate tec- Gren Foundation, and US NSF grants EAR8206183 and EAR8503232. S.
tonics (Kasting et al., 2006). The global oxygen isotope record also shows Gitau, H. Vallianatos, J. Wynn, E. Bestland, D. Dugas, G. Thackray and D.
an increase after 3 Ma due to continental ice cap sequestration of isoto- Retallack aided fieldwork. Invaluable discussions and advice came from
pically light oxygen, in addition to temperature effects (Zachos et al., P. Shipman, M. Pickford, B. Benefit and M. McCrossin.
2001). The African paleosol record and global isotopic records thus
presents a concept of climatic variation experienced by our distant an- Appendix A. Supplementary data
cestors very different from past ideas of a seminal late Miocene climatic
event. Instead of a single origin of humanity at a turning point of envi- Supplementary data to this article can be found online at https://doi.
ronmental change, the new record implies rather that our lineage org/10.1016/j.eve.2023.100005.
responded to a gauntlet of changing conditions with a variety of
adaptations. References
Each fluctuation in climate and vegetation presented new crises and
opportunities to African hominoids. Some specialized features such as Almecija, S., Hammond, A.S., Thompson, N.E., Pugh, K.D., Moya-Sola, S., Alba, D.M.,
2021. Fossil apes and human evolution. Science 372, eabb4363.
procumbent incisors evolved by about 18 Ma in Afropithecus, Ekembo and Andrews, P., 1978. A revision of the Miocene Hominoidea from East Africa. Bull. Brit.
Kenyapithecus for hard-skinned fruits of dry woodlands (McCrossin and Mus. Nat. Hist. 30, 85–224.
Benefit, 1997). Several features of the femur in Ugandapithecus by 20 Ma Andews, P., 1982. Ecological polarity in primate evolution. Zool. J. Linn. Soc. 74,
233–244.
were adapted for climbing in woodlands (Senut, 2016). Yet other likely Andrews, P., 1996. Palaeoecology and hominid palaeoenvironments. Biol. Rev. 71,
forest adaptations such as erect stance in Orrorin and Sahelanthropus by 6 257–300.
Ma (Brunet et al., 2002; Senut et al., 2000), and flat face in Kenyanthropus Andrews, P., Simons, E., 1977. A new African Miocene gibbon-like genus, Dendropithecus
(Hominoidea, Primates) with distinctive postcranial adaptations: its significance to
by 3.5 Ma (Leakey et al., 2001) proved advantageous in the long term, origin of Hylobatidae. Folia Primatol. 28, 161–169.
just as did likely aridland adaptations such as thick enamel in Afropithecus Anhuf, D., Frankenberg, P., Lauer, W., 1999. Die postglaziale Warmphase vor 8000
and Equatorius by 18 Ma (Martin, 1985; McCrossin and Benefit, 1997; Jahren: eine Vegetations Rekonstruction für Afrika. Geogr. Rundsch. 51, 454–461.
Begun, D.R., 2007. Fossil record of Miocene hominoids. In: Henke, W., Tatersall, I. (Eds.),
Smith et al., 2003; Rossie and MacLatchy, 2013), adducted hallux in
Handbook of Paleoanthropology: v.II Primate Evolution and Human Origins.
Kenyapithecus by 15 Ma (McCrossin et al., 1998), and long legs and big Springer, Berlin, pp. 921–977.
brains in Homo by 2.5 Ma (Bramble and Liebermann, 2004). Although Behrensmeyer, A.K., Deino, A.L., Hill, A., Kingston, J.D., 2002. Geology and
each of these ideas could, and should, be debated individually, the geochronology of the middle Miocene Kipsaramon site complex: Muruyur beds,
Tugen Hills, Kenya. J. Hum. Evol. 42, 11–38.
general concept of human evolution as a generalist path through a Benefit, B.R., 1987. The Molar Morphology, Natural History, and Phylogenetic Position of
gauntlet of environmental challenges (Potts, 1996) is now supported by the Middle Miocene Monkey Victoriapithecus. PhD Dissertation. New York University,
the African record of fossil apes and soils. p. 1033.
Benefit, B.R., 1993. The permanent dentition and phylogenetic position of Victoriapithecus
from Maboko Island, Kenya. J. Hum. Evol. 25, 83–172.
5. Conclusions Benefit, B.R., McCrossin, M.L., 1997. Earliest known Old World monkey skull. Nature
388, 368–371.
Benefit, B.R., McCrossin, M.L., 2000. Middle Miocene hominoid origins. Science 287,
Fossil vertebrates in attritional assemblages within paleosols are not 2375a.
only a source of paleoecological information from adaptive features of Bestland, E.A., Krull, E.S., 1999. Palaeoenvironments of early Miocene Kisingiri volcano
their skeletons, but selection pressures on those adaptations can be Proconsul sites: evidence from carbon isotopes, palaeosols and hydromagmatic
deposits. J. Geol. Soc. Lond. 156, 965–976.
inferred from the paleosols as well. Furthermore, paleosol diversity re- Bishop, W.W., Miller, J.A., Fitch, F.J., 1969. New potassium-argon age determinations
veals habitat heterogeneity throughout the evolutionary history of our relevant to the Miocene fossil mammal sequence in East Africa. Am. J. Sci. 267,
apelike ancestors in Africa. Both fossils and paleosols can be considered 669–699.
Bittrich, V., Struck, M., 1989. What is primitive in Mesembryanthemaceae? An analysis of
from the perspective of ecological polarities, which assess the relative
evolutionary polarity of character states. South Afr. J. Bot. 55, 321–331.
emphasis on competition, fecundity, and withstanding. Similarly, soils Boschetto, H.B., Brown, F.H., McDougall, I., 1992. Stratigraphy of the Lothidok Range,
can be ranked for their relative fertility, physical disturbance, and northern Kenya, and K/Ar ages of its Miocene primates. J. Hum. Evol. 22, 47–71.
hardship. Miocene paleosols and primates of southwest Kenya consid- Bramble, D.M., Liebermann, B.E., 2004. Endurance running and the evolution of Homo.
Nature 432, 345–352.
ered from this perspective showed increased molarization with dimin- Brunet, M., Guy, F., Pilbeam, D., Mackaye, H.T., Likius, A., Ahounta, D., Beauvilain, A.,
ished paleoprecipitation. As many as 6 sympatric species of apes in single Blondel, C., Bocherens, H., Boisserie, J.R., de Bonis, L., Coppens, Y., de Jax, J.,
paleosols showed competitive exclusion in polarity ecospace. When Denys, C., Duringer, P., Eisenmann, V., Fanone, G., Fronty, P., Geraads, D.,
Lehmann, T., Lihoreau, F., Louchart, A., Mahamat, A., Merceron, G., Mouchelin, G.,
grasslands first appeared during the middle Miocene, early monkeys such Otero, O., Campomanes, P.P., de Leon, M.P., Rage, J.C., Sapanet, M., Schuster, M.,
as Victoriapithecus, adapted to them just as vervets and baboons live in Sudre, J., Tassy, P., Valentin, X., Vignaud, P., Viriot, L., Zazzo, A., Zollikofer, C., 2002.
grasslands today. However, early hominoid apes, such as Kenyapithecus, A new hominid from the upper Miocene of Chad, central Africa. Nature 418,
145–151.
remained in woodlands and forests. Grasslands played little role in evo- Campanale, J.P., Sun, T.Y., Montell, D.J., 2017. Development and dynamics of cell
lution of upright stance and other human adaptations until the Pliocene polarity at a glance. J. Cell Sci. 130, 1201–1207.
appearance of australopithecines, whereas vervets and baboons had a Cerling, T.E., 1992. Development of grasslands and savannas in East Africa during the
Neogene. Palaeogeogr. Palaeoclimatol. Palaeoecol. 97, 241–247.
long history of adaptation to grasslands back to middle Miocene.

9
G.J. Retallack Evolving Earth 1 (2023) 100005

Cerling, T.E., Harris, J.M., McFadden, B.J., Leakey, M.G., Quade, J., Eisenmann, V., Kropf, D.L., 1992. Establishment and expression of cellular polarity in fucoid zygotes.
Ehleringer, J.R., 1997. Global vegetation change through the Miocene/Pliocene Microbiol. Rev. 56, 316–339.
boundary. Nature 389, 153–158. Leakey, L.S.B., 1952. The environment of the Kenya lower Miocene apes. Actes Congr.
Collinson, M.E., Andrews, P., Bamford, M.K., 2009. Taphonomy of the early Miocene Panafr. Prehist. Comm. 18, 323–324.
flora, Hiwegi formation, Rusinga island, Kenya. J. Hum. Evol. 57, 149–162. Leakey, M.G., Walker, A., 1997. Afropithecus: function and phylogeny. In: Begun, D.R.,
Cote, S., Malit, N., Nengo, I., 2014. Additional mandibles of Rangwapithecus gordoni, an Ward, C.V., Rose, M.D. (Eds.), Function, Phylogeny, and Fossils: Miocene Hominoid
early Miocene catarrhine from the Tinderet localities of Western Kenya. Am. J. Phys. Evolution and Adaptations. Springer, Boston, Boston, MA, pp. 225–239.
Anthropol. 153, 341–352. Leakey, M.G., Spoor, F., Brown, F.H., Gathogo, P.N., Klarie, C., Leakey, L.N.,
Cote, S., McNulty, K.P., Stevens, N.J., Nengo, I.O., 2016. A detailed assessment of the McDougall, I., 2001. New hominin genus from eastern Africa shows diverse middle
maxillary morphology of Limnopithecus evansi with implications for the taxonomy of Pliocene lineages. Nature 410, 433–440.
the genus. J. Hum. Evol. 94, 83–91. Leakey, R.E., Leakey, M.G., Walker, A.C., 1988a. Morphology of Turkanapithecus
Deino, A.L., Tauxe, L., Monaghan, M., Drake, R., 1990. 40Ar/39Ar calibration of the litho- kalakolensis from Kenya. Am. J. Phys. Anthropol. 76, 277–288.
and paleomagnetic- stratigraphies of the Ngorora Formation, Kenya. J. Geol. 96, Leakey, R.E., Leakey, M.G., Walker, A.C., 1988b. Morphology of Afropithecus turkanensis
567–587. from Kenya. Am. J. Phys. Anthropol. 76, 289–307.
de Menocal, P.B., 2004. African climate change and faunal evolution during the Pliocene- Le Gros Clark, W., Leakey, L., 1951. Fossil Mammals of Africa. I. The Miocene
Pleistocene. Earth Planet Sci. Lett. 220, 3–24. Hominoidea of East Africa. Brit. Mus. Nat. Hist., London, p. 117.
Drake, R.E., Van Couvering, J.A., Pickford, M.H., Curtis, G.H., Harris, J.A., 1988. New Liutkus-Pierce, C.M., Takashita-Bynum, K.K., Beane, L.A., Edwards, C.T., Burns, O.E.,
chronology for the early Miocene mammalian faunas of Kisingiri, western Kenya. Mana, S., Hemming, S., Grossman, A., Wright, J.D., Kirera, F.M., 2019.
J. Geol. Soc. London 145, 479–491. Reconstruction of the early Miocene critical zone at Loperot, southwestern Turkana,
Dugas, D.P., Retallack, G.J., 1993. Middle Miocene fossil grasses from Fort Ternan, Kenya. Kenya. Frontiers Ecol. Evol. 7, 44.
J. Paleontol. 67, 113–128. Lovejoy, C.O., Suwa, G., Spurlock, L., Asfaw, B., White, T.D., 2009. The pelvis and femur
Dworkin, S.I., 2023. Using paleosol mineral assemblages to reconstruct evolving changes of Ardipithecus ramidus: the emergence of upright walking. Science 326, 71–71e6.
in climate. Abstract Goldschmit Conference Lyon. https://conf.goldschmidt.info/gold Lukens, W.E., Lehmann, T., Peppe, D.J., Fox, D.L., Driese, S.G., McNulty, K.P., 2017. The
schmidt/2023/meetingapp.cgi/Paper/15142. Early Miocene critical zone at Karungu, western Kenya: an equatorial, open habitat
Feibel, C.S., Brown, F.H., 1991. Age of the primate-bearing deposits on Maboko island, with few primate remains. Front. Earth Sci. 5, 87.
Kenya. J. Hum. Evol. 21, 221–225. MacLatchy, L.M., Cote, S.M., Deino, A.L., Kityo, R.M., Mugume, A.A., Rossie, J.B.,
Fleagle, J.G., 2013. Primate Adaptation and Evolution. Academic Press, New York, p. 486. Sanders, W.J., Cosman, M.N., Driese, S.G., Fox, D.L., Freeman, A.J., 2023. The
Fleagle, J.F., Kay, R.F., 1985. The paleobiology of catarrhines. In: Delson, E. (Ed.), evolution of hominoid locomotor versatility: evidence from Moroto, a 21 Ma site in
Ancestors the Hard Evidence. Alan R. Liss, New York, pp. 23–36. Uganda. Science 380, eabq2835.
Frost, S.R., 2007. African Pliocene and Pleistocene cercopithecid evolution and global Manning, J.T., Chamberlain, A.T., 1993. Fluctuating asymmetry, sexual selection and
climatic change. In: Bobe, R., Alemseged, Z., Behrensmeyer, A.K. (Eds.), Hominin canine teeth in primates. Proc. R. Soc. London B251, pp. 83–87.
Environments in the East African Pliocene: an Assessment of the Faunal Evidence. Martin, R.A., 1990. Estimating body mass and correlated variables in extinct mammals.
Springer, Dordrecht, pp. 51–76. In: Damuth, J., McFadden, B.J. (Eds.), Body Size in Mammalian Paleobiology.
Gommery, D., Senut, B., Pickford, M., Musime, E., 2002. Les nouveaux restes du squelette Columbia University Press, New York, pp. 49–69.
d'Ugandapithecus major (Miocene inferieur du Napak). Ann. Paleontol. 88, 167–186. Martin, L., 1985. Significance of enamel thickness in hominoid evolution. Nature 314,
Grime, J.P., 1979. Plant Strategies and Vegetation Processes. Wiley, Chichester, p. 222. 260–262.
Harrison, T., 1992. A reassessment of the taxonomic and phylogenetic affinities of the McBrearty, S., Jablonski, N.G., 2005. First fossil chimpanzee. Nature 437, 105–108.
fossil catarrhines from Fort Ternan, Kenya. Primates 33, 501–522. McCrossin, M.L., Benefit, B.R., 1997. On the relationships and adaptations of
Harrison, T., 2002. Late Oligocene to middle Miocene catarrhines from Afro-Arabia. In: Kenyapithecus, a large bodied hominoid from the middle Miocene of eastern Africa.
The Primate Fossil Record (Hartwig, W.C. Cambridge Univ. Press, Cambridge, In: Begun, D.R., Ward, C.V., Rose, M.D. (Eds.), Function, Phylogeny and Fossils:
pp. 311–338. Miocene Hominoid Evolution and Adaptations. Plenum, New York, pp. 241–267.
Harrison, T., 2010. Dendropithecoidea, proconsuloidea and hominoidea. In: Werdelin, L., McCrossin, M.L., Benefit, B.R., Gitau, S.R., Palmer, A.K., Blue, K.T., 1998. Fossil evidence
Sanders, W.J. (Eds.), Cenozoic Mammals of Africa. University of California Press, for the origins of terrestriality among Old World higher primates. In: Strasser, E.,
Berkeley, pp. 429–469. Fleagle, J.G., McHenry, H.M., Rosenberger, A.L. (Eds.), Primate Locomotion: Recent
Hill, A., Nengo, I.O., Rossie, J.B., 2013. A Rangwapithecus gordoni mandible from the early Advances. Plenum, New York, pp. 353–376.
Miocene site of Songhor, Kenya. J. Hum. Evol. 65, 490–500. McNulty, K.P., Begun, D.R., Kelley, J., Manthi, F.K., Mbua, E.N., 2015. A systematic
Jacobs, B.F., 2002a. Estimation of low latitude paleoclimates using fossil angiosperm revision of Proconsul with the description of a new genus of early Miocene hominoid.
leaves: examples from the Miocene Tugen Hills, Kenya. Paleobiology 28, 399–421. J. Hum. Evol. 84, 42–61.
Jacobs, B.F., 2002b. Estimation of rainfall variables from leaf characters in tropical Africa. Michel, L.A., Peppe, D.J., Lutz, J.A., Driese, S.G., Dunsworth, H.M., Harcourt-Smith, W.E.,
Palaeogeogr. Palaeoclimatol. Palaeoecol. 145, 231–250. Horner, W.H., Lehmann, T., Nightingale, S., McNulty, K.P., 2014. Remnants of an
Jacobs, B.F., Deino, A.L., 1996. Test of climate-leaf physiognomy regression models, their ancient forest provide ecological context for Early Miocene fossil apes. Nat. Commun.
application to two Miocene floras from Kenya, and 40Ar/39Ar dating of the late 5, 1–9.
Miocene Kapturo site. Palaeogeogr. Palaeoclimatol. Palaeoecol. 123, 359271. Michel, L.A., Lehmann, T., Mcnulty, K.P., Driese, S.G., Dunsworth, H., Fox, D.L., Harcourt-
Janis, C.M., Carrano, M., 1992. Scaling of reproductive turnover in archosaurs and Smith, W.E., Jenkins, K., Peppe, D.J., 2020. Sedimentological and
mammals: why are large terrestrial mammals so rare? Acta Zool. Gennici 26, palaeoenvironmental study from Waregi hill in the Hiwegi Formation (early
201–226. Miocene) on Rusinga Island, Lake Victoria, Kenya. Sedimentology 67, 3567–3594.
Jimenez-Arenas, J.M., Perez-Claros, J.A., Aledo, J.C., Palmqvist, P., 2014. On the Miller, E.R., Benefit, B.R., McCrossin, M.L., Plavcan, J.M., Leakey, M.G., El-
relationships of postcanine tooth size with dietary quality and brain volume in Barkooky, A.N., Hamdan, M.A., Gawad, M.A., Hassan, S.M., Simons, E.L., 2009.
primates: implications for hominin evolution. BioMed Res. Int. 2014, 406507. Systematics of early and middle Miocene old world monkeys. J. Hum. Evol. 57,
Kappeler, P.M., 1996. Intrasexual selection and phylogenetic constraints on the evolution 195–211.
of canine dimorphism in strepsirhine primates. J. Evol. Biol. 9, 43–65. Mizota, C., Kawaski, I., Wakatsuki, T., 1988. Clay mineralogy and chemistry of seven
Kappelman, J., 1991. The paleoenvironment of Kenyapithecus at Fort ternan. J. Hum. Evol. pedons formed in volcanic ash. Geoderma 43, 131–145.
20, 95–129. Peppe, D.J., Cote, S.M., Deino, A.L., Fox, D.L., Kingston, J.D., Kinyanjui, R.N.,
Kasting, J.F., Howard, M.T., Wallmann, K., Veizer, J., Shields, G., Jaffres, J., 2006. Lukens, W.E., MacLatchy, L.M., Novello, A., Str€ omberg, C.A., Driese, S.G., 2023.
Paleoclimates, ocean depth, and the oxygen isotopic composition of seawater. Earth Oldest evidence of abundant C4 grasses and habitat heterogeneity in eastern Africa.
Planet Sci. Lett. 252, 82–93. Science 380, 173–177.
Kay, R.F., 2010. A new primate from the early Miocene of Gran Barranca, Chubut Pickford, M., 1985. A new look at Kenyapithecus based on new discoveries in Kenya.
Province, Argentina: paleoecological implications. In: Madden, R.H., Carlini, A.A., J. Hum. Evol. 14, 113–143.
Vucetich, M.G., Kay, R.F. (Eds.), The Paleontology of Gran Barranca: Evolution and Pickford, M., 1986a. Cenozoic palaeontological sites in western Kenya. Münchner
Environmental Change through the Middle Cenozoic of Patagonia. Cambridge Geowiss. Abhand, Reihe A Geol. Pal€aont 8, 151.
University Press, Cambridge, pp. 220–239. Pickford, M., 1986b. Sedimentation and fossil preservation in the Nyanza Rift system,
Kay, R.F., Plavcan, J.M., Glander, K.E., Wright, P.C., 1988. Sexual selection and canine Kenya. In: Frostick, L.E., Renaut, R.W., Reid, I., Tiercelin, J.J. (Eds.), Sedimentation in
dimorphism in New World monkeys. Am. J. Phys. Anthropol. 77, 385–397. the East African Rifts, vol. 25. Geol. Soc. London Spec. Publ., pp. 345–362
Kelley, J., 1986. Species recognition and sexual dimorphism in Proconsul and Pickford, M., 1995. Fossil land snails of East Africa and their paleoecological significance.
Rangwapithecus. J. Hum. Evol. 15, 461–495. J. Afr. Earth Sci. 20, 167–226.
Kelley, J., 1995. Sex determination in Miocene catarrhine primates. Am. J. Phys. Pickford, M., 2002. Early Miocene grassland ecosystem at Bukwa, Mt Elgon, Uganda.
Anthropol. 96, 391–397. Compt. Rend. Paleovol. 1, 213–219.
Kelley, J., Ward, S., Brown, B., Hill, A., Duren, D.L., 2002. Dental remains of Equatorius Pickford, M., Kunimatsu, Y., 2005. Catarrhines from the middle Miocene (ca. 14.5 Ma) of
africanus from Kipsaramon, Tugen Hills, Baringo district, Kenya. J. Hum. Evol. 42, Kipsaramon, Tugen Hills, Kenya. Anthropol. Sci. 113, 189–224.
39–62. Pickford, M., Senut, B., Gommery, D., 1999. Sexual dimorphism in Morotopithecus bishopi:
Kelley, J., Andrews, P., Alpagut, B., 2008. A new hominoid species from the middle an early Middle Miocene hominoid from Uganda, and a reassessment of its geological
Miocene site of Paşalar, Turkey. J. Hum. Evol. 54, 455–479. and biological contexts. In: Andrews, P., Banham, P. (Eds.), Cenozoic Environments
Kent, P.E., 1944. The Miocene beds of Kavirondo, Kenya. Q. J. Geol. Soc. Lond. 100, and Hominid Evolution: a Tribute to Bill Bishop. Geological Society of London,
85–118. London, pp. 27–38.
Kinzey, W.G., 1992. Dietary and dental adaptations in the Pithecinae. Am. J. Phys.
Anthropol. 88, 499–514.

10
G.J. Retallack Evolving Earth 1 (2023) 100005

Pickford, M., Sawada, Y., Tayama, R., Matsuda, Y.K., Itaya, T., Hyodo, H., Senut, B., 2006. Smith, T.M., Martin, L.B., Leakey, M.G., 2003. Enamel thickness, microstructure and
Refinement of the age of the middle Miocene Fort Ternan beds, western Kenya, and development in Afropithecus turkanensis. J. Hum. Evol. 44, 283–306.
its implications for old world biochronology. Compt. Rendus Geosci. 338, 545–555. Stevens, P.F., 1980. Evolutionary polarity of character states. Annu. Rev. Ecol. Systemat.
Plavcan, J.M., 1998. Correlated response, competition, and female canine size in 11, 333–358.
primates. Am. J. Phys. Anthropol. 107, 401–416. Suwa, G., Kono, R.T., Katoh, S., Asfaw, B., Beyene, Y., 2007. A new species of great ape
Plavcan, J.M., van Schaik, C.P., Kappeler, P.M., 1995. Competition, coalitions and canine from the late Miocene epoch in Ethiopia. Nature 448, 921–924.
size in primates. J. Hum. Evol. 28, 245–276. Thoren, S., Lindenfors, P., Kappeler, P.M., 2006. Phylogenetic analyses of dimorphism in
Potts, R., 1996. Evolution and climate variability. Science 273, 922–923. primates: evidence for stronger selection on canine size than on body size. Am. J.
Retallack, G.J., 1991. Miocene Paleosols and Ape Habitats of Pakistan and Kenya. Oxford Phys. Anthropol. 130, 50–59.
University Press, New York, p. 246. Tuttle, R.H., 2006. Seven decades of East African Miocene anthropoid studies. In:
Retallack, G.J., 1992. Middle Miocene fossil plants from Fort Ternan (Kenya) and Ishida, H., Tuttle, R., Pickford, M., Ogihara, N., Nakatsukasa, M. (Eds.), Human
evolution of African grasslands. Paleobiology 18, 383–400. Origins and Environmental Backgrounds. Springer, Boston, pp. 15–29.
Retallack, G.J., 2004. Ecological polarities of mid-Cenozoic fossil plants and animals from Veizer, J., Godderis, Y., François, L.M., 2001. Evidence for decoupling of atmospheric CO2
central Oregon. Paleobiology 30, 561–588. and global climate during the Phanerozoic. Nature 408, 698–701.
Retallack, G.J., 2005. Pedogenic carbonate proxies for amount and seasonality of Verhaegen, M., Puech, P.F., Munro, S., 2002. Aquarboreal ancestors? Trends Ecol. Evol.
precipitation in paleosols. Geology 33, 333–336. 17, 212–217.
Retallack, G.J., 2007. Paleosols. In: Henke, W., Tatersall, I. (Eds.), Handbook of Walker, A., Shipman, P., 2005. The Ape in the Tree: an Intellectual & Natural History of
Paleoanthropology. V. 1. Principles, Methods and Approaches. Springer, Berlin, Proconsul. Harvard Univ. Press, Cambridge, p. 312.
pp. 383–408. Ward, S.C., Duren, D.L., 2002. Middle to late Miocene African hominoids. In:
Retallack, G.J., 2012. Mallee model for mammal communities of the early Cenozoic and Hartwig, W.C. (Ed.), The Primate Fossil Record. Cambridge University Press,
Mesozoic. Palaeogeogr. Palaeoclimatol. Palaeoecol. 342, 111–129. Cambridge, pp. 385–397.
Retallack, G.J., 2013. Global cooling by grassland soils of the geological past and near Ward, C.V., Begun, D.R., Rose, M.D., 1997. Function and phylogeny in Miocene
future. Annu. Rev. Earth Planet Sci. 41, 69–86. hominoids. In: Begun, D.R., Ward, C.V., Rose, M.F.( (Eds.), Function, Phylogeny, and
Retallack, G.J., 2019. Soils of the Past: an Introduction to Paleopedology. Wiley, Fossils: Miocene Hominoid Evolution and Adaptations. Plenum Press, New York,
Chichester, p. 534. pp. 1–12.
Retallack, G.J., Bestland, E.A., Dugas, D.P., 1995. Miocene paleosols and habitats of Ward, S.C., Brown, B., Hill, A., Kelley, J., Downs, W., 1999. Equatorius: a new hominoid
Proconsul on Rusinga island, Kenya. J. Hum. Evol. 29, 53–91. genus from the middle Miocene of Kenya. Science 285, 1382–1386.
Retallack, G.J., Wynn, J.G., Benefit, B., McCrossin, M., 2002. Paleosols and Westoby, M., Falster, D.S., Moles, A.T., Fesk, P.Q., Wright, I.J., 2002. Plant ecological
paleoenvironments of the middle Miocene Maboko formation, Kenya. J. Hum. Evol. strategies; some leading dimensions of variation between spcies. Annu. Rev. Ecol.
42, 659–703. Systemat. 333, 125–159.
Rossie, J.B., MacLatchy, L., 2013. Dentognathic remains of an Afropithecus individual White, T.D., WoldeGabriel, G., Asfaw, B., Ambrose, S., Beyene, Y., Bernor, R.L.,
from Kalodirr, Kenya. J. Hum. Evol. 65, 199–208. Boisserie, J.R., Currie, B., Gilbert, H., Haile-Selassie, Y., Hart, W.K., 2006. Asa Issie,
Sankaran, M., Hanan, N.P., Scholes, R.J., Ratnam, J., Augustine, D.J., Cade, B.S., Aramis and the origin of Australopithecus. Nature 440, 883–889.
Gignoux, J., Higgins, S.I., Le Roux, X., Ludwig, F., Ardo, J., 2005. Determinants of White, T.D., Asfaw, B., Beyene, Y., Haile-Selassie, Y., Lovejoy, C.O., Suwa, G.,
woody cover in African savannas. Nature 438, 846–849. WoldeGabriel, G., 2009. Ardipithecus ramidus and the paleobiology of early hominids.
Selig, K.R., Silcox, M.T., 2022. Measuring molarization: change through time in premolar Science 326, 64–86.
function in an extinct stem primate lineage. J. Mammalian Evol. 29, 947–956. Wynn, J.G., 2001. Paleosols, stable carbon isotopes, and paleoenvironmental
Senut, B., 2016. Morphology and environment in some fossil hominoids and pedetids interpretation of Kanapoi, northern Kenya. J. Hum. Evol. 39, 411–432.
(mammalia). J. Anat. 228, 700–715. Wynn, J.G., 2004a. Miocene paleosols of Lothagam. In: Harris, J.M., Leakey, M.G. (Eds.),
Senut, B., Pickford, M., Gommery, D., Kunimatsu, Y., 2000. Un nouveau genre Lothagam: the Dawn of Humanity in Eastern Africa. Columbia University Press, New
d’hominoïde du Miocene inferieur d'Afrique orientale: Ugandapithecus major (Le Gros York, pp. 31–42.
Clark & Leakey, 1950). Compt. Rend. Acad. Sci. Ser. IIA Earth Planet. Sci. 331, Wynn, J.G., 2004b. Influence of Plio-Pleistocene aridification on human evolution:
227–233. evidence from paleosols of the Turkana Basin. Am. J. Phys. Anthropol. 123, 106–118.
Senut, B., Pickford, M., Gommery, D., Mein, P., Cheboi, K., Coppens, Y., 2001. First Wynn, J.G., Retallack, G.J., 2001. Paleoenvironmental reconstruction of middle Miocene
hominid from the Miocene (Lukeino Formation, Kenya). C. R. Acad. Sci. Terre Planet. paleosols bearing Kenyapithecus and Victoriapithecus, Nyakach Formation,
Paris 332, 137–144. southwestern Kenya. J. Hum. Evol. 40, 263–288.
Sheldon, N.D., Retallack, G.J., 2001. Equation for compaction of paleosols due to burial. Wynn, J.G., Alemseged, Z., Bobe, R., Geraads, D., Reed, D., Roman, D.C., 2006. Geological
Geology 29, 247–250. and palaeontological context of a Pliocene juvenile hominin at Dikika, Ethiopia.
Sheldon, N.D., Retallack, G.J., Tanaka, S., 2002. Geochemical climofunctions from North Nature 443, 332–336.
American soils and application to paleosols across the Eocene-Oligocene boundary in Zachos, J., Pagani, M., Sloan, L., Thomas, E., Billups, K., 2001. Trends, rhythms, and
Oregon. J. Geol. 110, 687–696. aberrations in global climate. Science 292, 686–693.
Simpson, G.G., 1949. The Meaning of Evolution: a Study of the History of Life and its
Meaning for Man. Yale Univ. Press, New Haven, p. 364.

11

You might also like