Você está na página 1de 10

Human Biology and the Origins of Homo : An Introduction to Supplement 6

Author(s): Leslie C. Aiello and Susan C. Antn


Reviewed work(s):
Source: Current Anthropology, Vol. 53, No. S6, Human Biology and the Origins of Homo
(December 2012), pp. S269-S277
Published by: The University of Chicago Press on behalf of Wenner-Gren Foundation for Anthropological
Research
Stable URL: http://www.jstor.org/stable/10.1086/667693 .
Accessed: 23/01/2013 13:36
Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at .
http://www.jstor.org/page/info/about/policies/terms.jsp

.
JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of
content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms
of scholarship. For more information about JSTOR, please contact support@jstor.org.

The University of Chicago Press and Wenner-Gren Foundation for Anthropological Research are collaborating
with JSTOR to digitize, preserve and extend access to Current Anthropology.

http://www.jstor.org

This content downloaded on Wed, 23 Jan 2013 13:36:28 PM


All use subject to JSTOR Terms and Conditions

Current Anthropology Volume 53, Supplement 6, December 2012

S269

Human Biology and the Origins of Homo


An Introduction to Supplement 6
by Leslie C. Aiello and Susan C. Anton
New fossil discoveries relevant to the origin of Homo have overturned conventional wisdom about the nature of
the australopiths and early Homo, and particularly Homo erectus (including Homo ergaster). They have eroded prior
assumptions about the differences between these genera and complicated interpretations for the origin and evolution
of Homo. This special issue surveys what is now known about the fossil evidence and the environmental context
of early Homo. It also moves beyond the hard evidence and sets the stage for integrated, multidisciplinary studies
to provide a framework for interpretation of the hard evidence. The underlying premise is that to understand the
adaptive shifts at the origin of Homo, it is essential to have a solid understanding of how and why modern humans
and other animals vary. Contributors to this issue include paleoanthropologists, human biologists, behavorialists,
and modelers. We tasked each with bringing her or his special expertise to bear on the question of the origins and
early evolution of Homo. The papers in this collection are a product of a week-long Wenner-Gren symposium held
in March 2011, and this introduction integrates this work and its significance for Homo.

What We Once Knew . . .


The origin of Homo holds particular sway for us and has often
been seen as the point in our evolution when the balance tips
from a more ape-like to a more human-like ancestor. By the
turn of this century, a conventional wisdom had grown up
around the origin of Homo and particularly Homo erectus that
cast this species as the first hominin to take important biological and behavioral steps in the direction of modern humans (Anton 2003; Shipman and Walker 1989). Homo erectus
was envisioned as a large-brained, small-toothed, long-legged,
narrow-hipped, and large-bodied hominin with relatively low
sexual dimorphism. By virtue of a higher-quality, perhaps
animal-based diet, H. erectus is said to have ranged farther,
cooperated more, and quickly dispersed from Africa (Aiello
and Key 2002; Anton, Leonard, and Robertson 2002; McHenry and Coffing 2000; Walker and Leakey 1993). The paucity of early Homo fossils of Homo habilis sensu lato (including
Homo rudolfensis) meant that comparisons of Australopithecus
((Paranthropus) were made to H. erectus (including Homo
ergaster) rather than to other early Homo. And the distinctions
between Australopithecus and Homo were perhaps overemLeslie C. Aiello is President of the Wenner-Gren Foundation for
Anthropological Research (470 Park Avenue South, New York, New
York 10016, U.S.A. [laiello@wennergren.org]). Susan C. Anton is a
Professor in the Center for the Study of Human Origins, Department
of Anthropology, New York University (Rufus D. Smith Hall, 25
Waverly Place, New York, New York 10003, U.S.A. [susan
.anton@nyu.edu]). This paper was submitted 12 XII 11, accepted 8
VII 12, and electronically published 27 IX 12.

phasized by the diminutive size of the most complete Australopithecus skeleton (A.L. 288-1; Lucy), on the one hand,
and the surprisingly large size of the most complete H. erectus
skeleton (KNM-WT 15000; Nariokotome boy), on the other
(e.g., Ruff 1993). The comparisons between H. erectus and
Homo sapiens were so strongly drawn that the inclusion in
the genus of some of the earliest species, such as H. habilis
and H. rudolfensis, was seriously questioned on the basis of
their more australopith-like postcranial skeleton, among other
things (Wood and Baker 2011; Wood and Collard 1999, 2007).
The fossil record never ceases to upset conventional wisdom, and over the past 2 decades, new discoveries from East
and South Africa, Georgia, and even Indonesia have challenged these stark distinctions between Australopithecus and
H. erectus and within non-erectus early Homo. In particular,
new small-bodied and small-brained finds from the Republic
of Georgia and Kenya call to question claims for universally
large size in H. erectus (e.g., Gabunia et al. 2000; Potts et al.
2004; Simpson et al. 2008; Spoor et al. 2007) and focus our
attention instead on the range of variation within that taxon.
This variation in H. erectus has most often been referred to
as sexual dimorphism and/or regional/climatic adaptations
(Anton 2008; Spoor et al. 2007), although short-term accommodations and phenotypic plasticity are likely to have played
an important role (see Anton 2013). And larger-sized, longerlegged Australopithecus have been found (Haile-Selassie et al.
2010), as have members of that genus who may share some
postcranial characteristics with Homo (Asfaw et al. 1999; Berger et al. 2010; Kibii et al. 2011; Kivell et al. 2011; Zipfel et
al. 2011). Additionally, new fossil remains of non-erectus
Homo and new work on previously known remains emphasize

2012 by The Wenner-Gren Foundation for Anthropological Research. All rights reserved. 0011-3204/2012/53S6-0002$10.00. DOI: 10.1086/667693

This content downloaded on Wed, 23 Jan 2013 13:36:28 PM


All use subject to JSTOR Terms and Conditions

S270

Current Anthropology Volume 53, Supplement 6, December 2012

the diversity of the early members of the genus and the ways
in which they differ from Australopithecus (Blumenschine et
al. 2003; Spoor et al. 2007).
Yet despite this increased appreciation of variation in early
fossil Homo, little time has been spent evaluating the relationships between morphology and behaviors in extant taxa,
especially modern humans, in different ecological circumstances. We maintain that these are the data that are essential
to create a more nuanced understanding of the implications
and expectations of anatomical changes at the origin of our
genusan understanding that goes beyond simple assumptions of sexual or climatic variation.

Topic and Rationale


The increasing number of early Homo fossil finds and their
diversity in size and shape suggest that this discussion surrounding the origin and evolution of early Homo is likely to
form a major focus for the next decade of paleoanthropological work. As such, our goal was to bring together human
biologists, behaviorists, modelers, and fossil experts to integrate the rich extant data sets with the new details of the fossil
record. Understanding the adaptive shifts at the origin of
Homo is dependent on a solid understanding of how and why
modern humans vary and particularly on the relationship
between human behavior, human morphology, and human
lifestyle and life history variation.
Among the highly variable features in living humans are
features such as body size and aspects of life history that
separate us from other primates. In many cases, human variation in, for example, growth rates, fertility, and perhaps even
lifespan, can be traced to such environmental or behavioral
factors as nutritional sufficiency and unavoidable (extrinsic)
mortality. Such phenotypic plasticity provides a more rapid
response to environmental challenges than does genetic
change, but the fact that genetic change can follow has been
long suggested in human biology (e.g., Kuzawa and Bragg
2012). Thus, understanding the causes of human phenotypic
plasticity can provide important clues to understanding both
within and between species variation in the morphology of
our hominin ancestors.
Unavoidably, the biology of early Homo will be unlike that
of ourselves, howeverand therefore, primate and mammalian trends are also important to understand. And because
at some point cooperation in hunting or breeding became
important to survival, considering how both carnivory and
cooperation influence life history, body size, and body shape
and whether they leave a detectable signal are important considerations as well.
We set out, then, to probe the meaning of the newly identified ranges of variation in size and shape in early Homo
based on empirical evidence of how extant humans, nonhuman primates, and social carnivores respond energetically,
physiologically, and socially to changes in resource availability
and to stress from climatic, environmental, and other factors.

We argue that understanding the response of extant organisms, especially humans, in shifting environments provides
an ideal basis for understanding the integration of biocultural
responses to environmental constraints. The application of
these data in light of the known fossil record can help us to
understand these past populations, their constraints and adaptive strategies. By mining the rich data sets of our subspecialties, we sought to forge a stronger and more nuanced
understanding of the adaptive shifts that canor cannot
be inferred at the base of our genus and to set out a series
of hypotheses and predictions to be tested against future fossil
and archaeological data. The results of an intense 5-day Wenner-Gren Symposium in Sintra, Portugal, in March 2011 and
our follow-up analyses are presented in this special issue.

Setting the Stage


We begin the volume, as we did the symposium, by reassessing
the fossil foundation of what we now know regarding genus
Homo. Anton (2012) provides an overview of the genus and
its species and the differences between Australopithecus and
Homo. The first recognizable members of the genus Homo
appear at approximately 2.3 Ma, suggesting that the genus
evolved earlier, but substantial fossil evidence does not appear
until about 2.0 Ma. Her paper focuses our attention on the
importance of individual fossil data points for understanding
diversity within and between groups, and she concludes that
a strong case can be made for at least three different morphs
between 2.0 and 1.5 Ma: an 1813-group, a 1470-group, and
Homo erectus (including Homo ergaster). She avoids the use
of taxonomic names for the 1813-group and 1470-group because of uncertainty over group affiliation of type specimens
for early Homo species (e.g., Homo habilis and Homo rudolfensis; Leakey et al. 2012). Her paper also provides an introduction to what is now known about the distinct features
that separate the three different morphs from each other and
from Australopithecus (see Anton 2012, tables 18, and Anton
and Snodgrass 2012, tables 16, for membership of these
morphs and distinguishing morphological and inferred behavioral features). Additionally, she suggests that, on average,
early Homo is larger of body and brain than Australopithecus,
and H. erectus is larger than other early Homo. That said, the
surprising facts, particularly to those who have been involved
in paleoanthropology for a considerable time, are the degree
of diversity within the morphs and that, in some ways, the
morphs are more similar to each other than was previously
imagined. For example, all early Homo, including H. erectus,
may exhibit substantial amounts of sexual dimorphism, and
H. erectus is less fully modern in body proportions than has
been previously claimed. These themes and their implications
are further plumbed in the contributions by Holliday (2012),
Pontzer (2012), and Plavcan (2012).
Fossils cannot be understood and interpreted without their
context, and Potts (2012) provides an overview of the environmental and archaeological background for the evolution

This content downloaded on Wed, 23 Jan 2013 13:36:28 PM


All use subject to JSTOR Terms and Conditions

Aiello and Anton Origins of Homo

S271

of Homo in eastern Africa between 3.0 and 1.5 Ma. He concludes that there were major episodes of moist-arid variability
during this period, superimposed on an overall drying trend.
The first appearance of Homo at approximately 2.3 Ma (Kimbel et al. 1996; and of the Oldowan at approximately 2.58
Ma; Semaw et al. 2003; but see McPherron et al. 2010) as
well as the proliferation of the genus after 2.0 Ma coincide
with particularly high levels of climate variability, suggesting
that adaptive plasticity in its broadest developmental, physiological, and behavioral manifestations was integral to the
evolution of Homo. For example, stone tools, which have a
strong stratigraphic persistence in the archaeological record
after 2.0 Ma, provide an efficient behavioral mechanism to
enhance foraging ability, enabling predictable returns in a
changing environment. But they also pose an energetic challenge of material transport over distances as great as 113 km
by about 2.0 Ma (Braun et al. 2008).

Food, Morphology, and Locomotion


One means of offsetting the energetic cost of tool and raw
material transport as well as increased body and brain size is
dietary expansion to higher-quality food resources, which
might involve access to animal resources (as well as a wider
range of plant food; Aiello and Wells 2002; Aiello and Wheeler
1995; Leonard and Robertson 1997). Such resources would
also serve to buffer environmental instability and resulting
changes of food resources across space and over time (Potts
2012). Some direct evidence for a dietary shift in early Homo
(including even more substantial changes in Homo erectus)
relative to the diet of Australopithecus is provided by Ungar
(2012), who reviews dental macro and micro anatomy and
wear. In particular, all early Homo teeth are most similar to
extant animals that do not use fracture-resistant foods. In
other words, the genus seems not to have used particularly
hard-brittle foods or especially tough foods. However, within
early members of the genus, there are some differences that
suggest a broader subsistence base for H. erectus that included
more tough foods than other early Homo. This dental evidence
is consistent with increased meat eating (or eating other nonbrittle foods) and tool use in food preparation (perhaps even
cooking) over the condition in Australopithecus, with a
broader range of foods eaten by H. erectus than other early
Homo. These results could be a hard tissue signal of dietary
and behavioral plasticity to temper environmental vacillation.
Like the dietary results, other papers suggest that some
adaptations once thought to appear with H. erectus arise at
the origin of the genus or even earlier. Holliday (2012) provides new analyses and an overview of our current knowledge
of body size and body proportions. The unexpected outcome
is that our prior understanding that H. erectus was unique
among the early hominins in having long legs and a narrow,
heat-adapted body is wrong. Leg length scales with body mass,
and large-bodied australopiths have long, human-like legs and
human-like thoraces, while new analyses of the H. erectus

pelvis demonstrate that its bi-illiac breadth was broad and


australopith-like (Simpson et al. 2008; but see Ruff 2010).
Because locomotor efficiency is primarily a function of relative
limb length (Pontzer 2012), and because limb length is allometrically related to body size in all hominins, larger-bodied
individuals of any taxon would be more efficient in walking
and running, with faster optimal speeds and increased absolute speed. Long leg length and arm length also have thermoregulatory advantages in hot climates, which would be a
distinct advantage either during locomotion or at rest.
Given the similar scaling relationships between limb lengths
and body size across hominins, Holliday concludes that there
is little evidence of a major locomotor shift between Australopithecus and early Homo (including H. erectus), a point
that is shown by the analyses presented by Pontzer (2012) as
well. However, they note that a significant difference remains
between these genera in terms of mean body size; early Homo
is approximately 33% larger than Australopithecus, and H.
erectus is approximately 15% larger than other early Homo,
even when the recently discovered small H. erectus fossils (e.g.,
from Georgia, Kenya, Tanzania, and perhaps Ethiopia) and
large Australopithecus are included in estimates.
Despite having similar proportions as earlier hominins, a
number of symposium contributions emphasize that larger
size itself has important energetic, locomotor, and survival
consequences for Homo. Holliday (2012) and Pontzer (2012)
point out that across mammals, larger body size equates with
a larger home range size, which would be exaggerated further
if Homo was also more carnivorous (Anton, Leonard, and
Robertson 2002). Pontzer (2012) develops the implications
of this by demonstrating that across mammals there is no
selection for greater locomotor efficiency (as proxied by
changes in limb proportions) in those species with larger
home range sizes. Instead, species that travel farther adopt a
high-throughput strategy (increased daily energy expenditure
in relation to body size and a correspondingly greater reproductive investment), resulting in greater lifetime reproductive
output. This suggests that in Homo, as in other far-ranging
mammals, there must have been an increased energy budget
to provide for increased brain and body growth and reproduction. Outside of a more calorie-rich diet at the carnivorous
end of the omnivorous spectrum, Pontzer (2012) argues that
an increased daily energy expenditure would suggest greater
food availability, perhaps implying the origins of food sharing.
These results are consistent with the symposium papers on
living humans and extant carnivores by Migliano and Guillon
(2012), Kuzawa and Bragg (2012), and Smith and colleagues
(2012).

Body Size and Growth


While height in humans is influenced by a number of environmental and idiosyncratic factors (see references in Kuzawa and Bragg 2012; Migliano and Guillon 2012), it is
achieved through a combination of speed and duration of

This content downloaded on Wed, 23 Jan 2013 13:36:28 PM


All use subject to JSTOR Terms and Conditions

S272

Current Anthropology Volume 53, Supplement 6, December 2012

growth, which is in turn dependent on resource availability


and mortality probability (Kuzawa and Bragg 2012; Migliano
and Guillon 2012). This may provide a clue for possible interpretations of size variation in Homo. In human populations, the greater the probability of mortality, the earlier is
the age of maturity (and the shorter the period of growth),
to ensure maximum reproductive output. Based on the analysis of small-scale human societies, Migliano and Guillon
(2012) demonstrate that the main determinant in height variability in humans is the probability of mortalitythe lower
the mortality probability, the taller the populations. They also
show that environment has an important effect, although diet
is not significant in their analyses. This probably results from
data limitations in their large comparative analysis.
Kuzawa and Bragg (2012) emphasize this nutritional side
of the equation and argue that nutritional abundance is associated with a faster growth rate, earlier maturity, and larger
adult sizes. Because males are affected more than females,
sexual size dimorphism is increased with greater nutritional
abundance. Nutritional stress has the opposite effectsslower
growth, later maturity, and reduced size dimorphism. If mortality rates were high and precluded later maturity, smaller
body size would be expected.
Pfeiffers (2012) work on the bioarchaeological record of
the small-bodied KhoeSan, however, reminds us of the multifactorial effect on size and of the issue that small size may
be the default in the absence of selective factors for larger
size. That is, bigger may not always be better. In her particular
case study she finds no evidence for the traditional drivers of
small size: nutritional insufficiency, early maturation, high
extrinsic mortality, or climate change. In light of this, she
suggests instead that the long-term relaxation of selection for
large size was allowed due to the relative isolation of the
KhoeSan and therefore an absence of competition with largesized human populations. This might favor increasingly large
male size and shape changes to the pelvis that accommodated
relatively large infants in small mothers, which might otherwise favor large females. It is important, then, to consider
the multiple and sometimes conflicting causes for size change
in light of their influence on developmental plasticity.
At present we lack the detailed data from the fossil record
to assess intraspecific differences in growth and development
with the aim of inferring the possible roles of nutrition, mortality probability, or other factors in shaping the observed size
differences in the hominins. However, because the size variation, particularly in Homo erectus, is similar to that found in
modern humans (Migliano and Guillon 2012), there is every
reason to assume that similar factors were in play and that size
differences in the hominins also reflect an adaptive plasticity
similar to that observed in modern humans. Migliano and
Guillon (2012), Kuzawa and Bragg (2012), and Bribiescas and
colleagues (2012) also emphasize the important role of behavior, and particularly cooperation, in buffering both nutritional
sufficiency and mortality probability, thereby setting the stage
for body size increase. A series of alternative scenarios for con-

sidering especially the regional variation in H. erectus derived


from these principles are developed in detail in the concluding
paper of this issue (Anton and Snodgrass 2012).
Although more data are sorely needed, what we know about
the tempo and mode of growth in the early hominins, based
on rates of dental maturation, is summarized by Schwartz
(2012). These data suggest that both Australopithecus and early
Homo have more rapid maturation than Homo sapiens, which
could reflect environments of higher extrinsic mortality. However, there are probably differences between the genera as well.
Dental eruption in Australopithecus and Paranthropus is comparable to, or faster than, Gorilla gorilla beringei, the fastest
of the living great apes. Dental eruption in H. erectus is equivalent to Pongo pygmaeus pygmaeus, the slowest of the living
great apes, and just below the large range of eruption ages in
modern humans. The somewhat extended developmental
schedule of H. erectus relative to Australopithecus is consistent
with mortality reduction and increased body size. There is
still much to learn, but one definite conclusion of Schwartzs
synthesis is that the full suite of modern human life history
with extended periods of growth and development was not
present in early Homo, including H. erectus, and possibly did
not appear in its modern form until much later in time.
While these papers address size differences between human
populations and what we know about the tempo of hominin
maturation, Plavcan (2012) raises the important issue of sexual size dimorphism. He provides a detailed overview of what
is currently known (and knowable) about sexual size dimorphism in hominins and living primates and provides a number of caveats in relation to interpretation of the evidence. It
has long been assumed that sexual dimorphism is a feature
observable in hominins that can be directly and causally related to social behavior across primates, and particularly to
male competition over mates (Leigh 1992; Plavcan 2001; Plavcan and van Schaik 1997a, 1997b). However, Plavcan cautions
that although all highly dimorphic primate species are polygynous, the inverse relationship is not straightforward, and
any hominin inferences can only be made with extreme caution.
A particularly important aspect of his research is the focus
on female size. He argues that female size represents the optimum for the particular environment, and male size represents a trade-off between the costs of deviating from this
optimum and the benefits of larger size in mate competition.
Female size change alone does not result in marked changes
in dimorphism across extant species. Although there is some
variation within species, without a change in mating competition, which drives male size increase or decrease, an increase in female size should simply be tracked by an equivalent
increase in male size. Beyond this, he notes that size variation
within H. erectus is unremarkable relative to human levels
and that our understanding of sexual size dimorphisms requires better understanding of temporal changes in male and
female size relative to each other in both humans and nonhuman primates. He makes a specific call for further system-

This content downloaded on Wed, 23 Jan 2013 13:36:28 PM


All use subject to JSTOR Terms and Conditions

Aiello and Anton Origins of Homo

S273

atic research on intraspecific geographic and sexual variation


in primates.
The evolution of male life history trade-offs has not been
a major focus in hominin evolution, outside of the relatively
simple association between reduced dimorphism and a move
away from polygynous social systems and strong size-driven
mating competition. However, Bribiescas and colleagues
(2012) argue that it is central to the evolution of Homo.
Consistent with Plavcans interspecific analyses demonstrating
no correspondence between reduced dimorphism and any
particular social system, Bribiescas and colleagues (2012)
point to the fact that humans, with their relatively reduced
dimorphism, are unique among the apes not only in the
degree of their paternal parenting behavior but also in its
variation. There is variation in the amount of time and energy
invested and in the type of offspring care, provisioning, and
other involvement. Paternal parenting behavior is dependent
on context, including the availability of other caregivers such
as grandmothers or siblings.
The important point is the move away from energy investment in large male size that permits not only energy allocation to parenting behavior but also to other aspects of
life history including fertility and longevity. One particularly
novel aspect of their work is the argument that increased male
fertility at older ages may have contributed to the emergence
of female longevity and the evolution of the female postreproductive lifespan and increased female reproductive effort
through grandmothering and child care. This provides an
alternative, or perhaps complementary, explanation to the
Grandmothering Hypothesis (Hawkes et al. 1998, 2011; Kachel, Premo, and Hublin 2011a, 2011b; OConnell, Hawkes,
and Blurton Jones 1999).
It is now clear that we cannot be sure whether there was
any significant difference in dimorphism between Australopithecus and Homo (either early Homo or H. erectus; Anton
2012; Holliday 2012; Plavcan 2012; Pontzer 2012). However,
if we could be sure that dimorphism was reduced in H. erectus,
we would have direct evidence of a change in mating behavior
leading to a reduction of sexual selection acting on male size,
and likely involving a major change in social organization
involving increased levels of cooperation and allocare, and
perhaps increased longevity. Given the importance of understanding the relationship between male and female size, it
would be useful to expend the effort to understand more
critically the relationship between skeletal and body mass dimorphism and how it varies among modern human populations.

Models for Cooperation, Sociality, Life


History, Body Size, and Brain Size
Although we cannot be confident that there was a change in
sexual size dimorphism associated with the evolution of
Homo, the idea that cooperation and food sharing may have
been major distinguishing factors between the Australopithe-

cus and early Homo was a common theme that developed


from a variety of perspectives throughout the symposium.
Many of the previously mentioned papers suggest that the
abilities to maximize food and to limit predation are critical
to increasing body and brain size. Food sharing and consequent group cooperation are means of achieving this. In light
of the several lines of evidence pointing toward the importance of sociality and cooperation, several symposium participants explored the correlates of such behavior in extant
organisms as well as the concept of cooperation as a means
of expanding an organisms capital.
Nonhuman primates provide the logical starting point because of their close phylogenetic relationship to humans. They
demonstrate the roots of human evolutionary plasticity particularly in dietary/niche expansion, extended life history, and
increasing social complexity with extensive cooperation and
communication (Anton and Snodgrass 2012). These abilities
provide the basis for the elaborate niche construction observed in humans that involves accelerating biocultural complexity and an increasing reliance on cooperation in all aspects
of hominin life (Fuentes, Wyczalkowski, and MacKinnon
2010; Odling-Smee, Laland, and Feldman 2003).
However, primates are not the only models for hominin
behavior, and useful insights can be drawn from, for example,
other large-brained animals such as dolphins, cooperative
breeders among all orders, and large-bodied mammal species
that inhabit woodland and savanna environments. In the late
1960s, Schaller and Lowther (1969) wrote a now classic paper
on the relevance of carnivore behavior to the study of early
hominins. Their basic premise was that to understand sociality
in hominins, it would be productive to draw inferences from
animals that are ecologically similar, such as social carnivores,
as well as animals that were closely related, such as the primates. At the time, this work represented a major innovation
in the interpretation of early hominin behavior. Smith and
colleagues (2012) take up where Schaller and Lowther left off
more than 40 years ago to demonstrate commonalities in
behavior and morphology between humans and the Carnivora. Sociality among carnivores is the exception rather than
the rule, but significant features related to sociality and cooperation in the Carnivora include cursorial hunting of large
game in open habitats, a relatively tall body build (shoulder
height in relation to body mass), reduced sexual dimorphism,
larger brains, a high reproductive output (in this case larger
litters), allocare of infants, increased weaning age, and larger
population density. Many of these features (with the possible
exception of reduced sexual dimorphism) are reminiscent of
the morphology of Homo and may help to infer behavior and
life history of early Homo (including Homo erectus). Indeed,
many of these features in social carnivores help to reinforce
the inferred relationships between dietary change, morphology, and cooperation reached by other symposium participants using other data sets.
Brain size expansion in Homo may provide another, independent avenue for inferring the presence of cooperative

This content downloaded on Wed, 23 Jan 2013 13:36:28 PM


All use subject to JSTOR Terms and Conditions

S274

Current Anthropology Volume 53, Supplement 6, December 2012

breeding. Isler and van Schaik (2012) draw on their previous


work on the Expensive Brain Hypothesis (Isler and van Schaik
2009) and on a large comparative mammalian database to
underscore the relationship between brain size increase and
cooperation in the form of allocare. They argue that across
mammals, large brain size is generally correlated with a reduction in population growth rate. This correlation is driven
by the extended ontogenetic periods necessary for the growth
of larger-bodied and larger-brained offspring and the correspondingly longer interbirth intervals required. Allocare provides extra resources to the mother, resulting in early weaning
of the infant and a shorter interbirth interval. An important
aspect of their work is the prediction based on extant primates
that a mean endocranial capacity of 600700 cc would be the
gray ceiling beyond which cooperation in the form of allocare would be essential if the population were to grow fast
enough to replace itself and avoid extinction. The great apes
seem to be at the very farthest extension of this relationship,
just barely allowing population replacement without allocare.
The fact that the mean brain size of the largest-bodied Australopithecus species (478 cc in Australopithecus afarensis; Holloway, Broadfield, and Yuan 2004) converges on that of modern apes suggests that cooperative breeding probably had not
yet appeared in these early hominins. However, by the time
of H. erectus, with brain sizes uniformly over 700 cc, reproductive cooperation would seem to have been a necessity.
Allocare is undoubtedly a key element that enabled hominins to break through the gray ceiling, but it is only one
element of capital in hominin evolution (Kaplan, Lancaster,
and Robson 2003; Kaplan et al. 2000). Wells (2012) sees capital
as a key and integrating concept that brings together many
of the themes of intraspecific and interspecific variation and
plasticity that emerged from the symposium. He defines capital as a generalized energy currency that can be expended in
a variety of ways to increase adaptive flexibility. This results
in the fact that humans are uniquely under-committed to
any specific niche.
Wells (2012) talks about social capital as well as physical
capital. Social capital is facilitated by larger brain sizes, stored
in social relationships and variably expended to achieve predator protection or to enable food security (particularly
through sexual division of labor and allocare). Physical capital
is stored within the body as adipose tissue or extracorporeally
in food hoards and similarly used to avoid predators and
provide sufficient nutrition. Because it is difficult to assess
from the fossil record, adipose tissue has received relatively
little attention in human evolution. Yet it is a major feature
distinguishing us (and our sexual dimorphism; see Anton and
Snodgrass 2012; Plavcan 2012) from nonhuman primates and
one that buffers the costs of reproduction against food shortages in fluctuating environments (Knott 1998; Kuzawa 1998).
It also is the source of signaling molecules responsible for
energy trade-offs between competing biological functions
such as growth, immune function, and reproduction (Wells
2009). It may be one of the major factors responsible for

differences in life history strategies among human populations, and it has been correlated with large brain sizes across
mammals (Navarrete, van Schaik, and Isler 2011). It is thus
important to develop creative ways to infer adiposity from
fossil record.
Storing energy in generalized currencies (social relationships and adipose tissue) means that various aspects of life
history (growth, reproduction, and immunity) can be
funded according to the state of the environment and overall energy availability (Wells 2012). If conditions demand it,
one aspect may be prioritized at the expense of others, resulting in the life history variation and its outcomes in features
such as growth rates, adult body size, fertility, and possibly
lifespan that are observed in modern humans (Bribiescas,
Ellison, and Gray 2012; Kuzawa and Bragg 2012; Migliano
and Guillon 2012). The symposium provided a vehicle for
bringing together the disparate data sets of our subdisciplines
into a framework that suggests ways in which variation is
produced, organized, and interrelated in the extant world.

What We Know Now and What We Hope to


Know in the Future
The final paper of the volume takes up the challenge of using
this framework to generate means of assessing the variation
observed in the fossil record and the biocultural relationship
between hominin morphology, hominin behavior, and the
fluctuating environment of the African Pliocene and Pleistocene (Anton and Snodgrass 2012). What we know is that
early Homo existed in a highly variable environment (a nonequilibrium ecosystem) that may have placed adaptive plasticity at a premium. The type of body size variation observed
in early Homo is consistent with the range observed in modern
humans that is mediated by life history differences in growth
and development that are dependent on energy availability
and mortality probability (Kuzawa and Bragg 2012; Migliano
and Guillon 2012; Migliano, Vinicius, and Lahr 2007). Dietary
differences, involving increased dietary breadth and a more
carnivorous diet, are also evident between these hominins and
are consistent with greater adaptive plasticity than inferred
for the australopiths. Comparative studies suggest that largerbodied hominins would have had to adopt a high energy
throughput strategy, and this, together with the increase in
brain size, would presuppose increased cooperation in the
form of allocare and sexual division of labor. The increased
cooperation and sociality would also be significant in group
protection in a relatively dangerous terrestrial environment
and create a relatively safe niche that would be consistent
with later maturation (and perhaps increased longevity) that
begins to be evident in the dental maturation evidence for
Homo erectus in relation to the other early hominins.
These results provide the basis for a model for the evolution
of Homo involving an integrated feedback loop that drove life
history evolution and contributed to cultural change (Anton
and Snodgrass 2012). The central elements of this model are

This content downloaded on Wed, 23 Jan 2013 13:36:28 PM


All use subject to JSTOR Terms and Conditions

Aiello and Anton Origins of Homo

S275

cooperative behavior, diet, cognitive abilities, and extrinsic


mortality risk. This model also generates a number of testable
hypotheses (e.g., to explain size variation in the hominins),
but each requires additional data from both the fossil and the
modern records to test. Increased body size across genera may
be hypothesized to signal either decreased extrinsic mortality,
increased nutritional sufficiency, or both. On the one hand,
specific predictions are made about rates of growth, timing
of weaning, and timing of growth cessation in each scenario
so additional growth data, especially from dental development
and especially for the earliest Homo, are needed. But beyond
simply finding more fossils, additional means of assessing
hard tissue growth in extant populations in ways that are
comparable to fossil samples are neededand the need to
consider intrapopulational variation among extant primate
and nonprimate mammals to the level that it is done in human
populations is also required. There are issues of scale and
comparability in our data sets that can only be remedied by
long-term analyses of extant populations.
The increase in brain and body size between other early
Homo and H. erectus again suggests decreased extrinsic mortality and/or increased nutrition. These features may reflect
adaptations to the higher mortality rates in terrestrial environments and perhaps cooperative huntingas shown in the
social carnivores. The extended developmental schedule of H.
erectus is consistent with mortality reduction, possibly as the
result of behavioral changes involving some form of cooperation. Here, along with additional fossil data and more comparable data sets, a key endeavor will be mining the rich social
carnivore data sets and creating new ones that consider aspects
of morphology, behavior, and variation. Regarding extrinsic
mortality, new ways can and should be developed to interrogate the archaeological and extant records for predator load
and the degree to which this can be assessed for different
hominin species. Novel means should also be developed for
assessing the ways in which population variation and physical
character development (including size and dimorphism) are
influenced by the high degree of climate variability that characterized this period of evolutionary history.
For example, it has been hypothesized for the small-sized
Dmanisi sample that nutritional insufficiency and perhaps
isolation resulted in short-term accommodations or adaptations (e.g., Anton 2003, 2013; see also Migliano and Guillon
2012). If this is the case, one would expect lower growth rates
and a longer period of maturation in relation to larger-bodied
H. erectus. Alternatively, if the short stature was due to a high
mortality environment, one would expect the smaller specimens to have more rapid growth rates and a shorter period
of maturation. Clearly this requires greater detail than we
currently possess regarding H. erectus growth, but it provides
a place to start.
While we end, as many such symposia do, with a decided
plea for more fossil remains in different localities, we hope
to move beyond that to an agenda of integrated and multidisciplinary studies to provide a framework against which to

test these predictions. For example, to understand the factors


surrounding the evolution of Homo, it will be essential to
target intra- and interpopulational research on energetic and
life history variation in various climatic, nutritional, and mortality environments. We need to tease out the complex interrelationships among these and other variables to understand the fundamental correlates of body size, brain size, and
sexual dimorphism (Kuzawa and Bragg 2012; Smith et al.
2012). Among many other things, we want to know, for example, how skeletal dimorphism tracks body mass dimorphism across populations, especially in humans (Plavcan
2012), and how individual skeletal features are influenced in
males and females in differing circumstances (Bribiescas, Ellison, and Gray 2012; Kuzawa and Bragg 2012).
Our thinking about the origins of Homo has continued to
change since Homo habilis was announced (Leakey, Tobias,
and Napier 1964), the almost complete Nariokotome H.
erectus skeleton was discovered (Brown et al. 1985), and
Dmanisi and other material changed our ideas about variation
in H. erectus (Gabunia et al. 2000; Potts et al. 2004; Simpson
et al. 2008; Spoor et al. 2007). New fossils will undoubtedly
continue to be uncovered. However, this material cannot be
interpreted in a vacuum, and the more we know about intraand interspecific variation in modern humans and other animals, the stronger the foundation we have for a rich understanding of our evolutionary past.

Acknowledgments
We would like to thank the Wenner-Gren Foundation for the
opportunity to hold this symposium and to publish the results
as an open-access supplementary issue of Current Anthropology. We would also like to thank all of the participants for
lively and stimulating discussion and debate over a 6-day
period in March 2011 at the Tivoli Palacio de Seteais Hotel
in Sintra, Portugal. This experience will be fondly remembered for a long time to come. We would like to give special
thanks to Chris Rainwater (New York University), who served
as the rapporteur for the meeting; to Emily Middletown (New
York University), who provided invaluable assistance in helping to edit and prepare all of the manuscripts for publication;
and to Lisa McKamy and the editorial and production staff
at the University of Chicago Press for their help in bringing
this issue to fruition. The meeting would not have been as
successful as it was without the deft organizational skills of
Laurie Obbink, the Wenner-Gren Foundation Conference Associate, and for this we are most grateful.

References Cited
Aiello, Leslie C., and Catherine Key. 2002. Energetic consequences of being
a Homo erectus female. American Journal of Human Biology 14:551565.
Aiello, Leslie C., and Jonathan C. K. Wells. 2002. Energetics and the evolution
of the genus Homo. Annual Review of Anthropology 31:323338.
Aiello, Leslie C., and Peter Wheeler. 1995. The expensive-tissue hypothesis:

This content downloaded on Wed, 23 Jan 2013 13:36:28 PM


All use subject to JSTOR Terms and Conditions

S276

Current Anthropology Volume 53, Supplement 6, December 2012

the brain and the digestive system in human and primate evolution. Current
Anthropology 36:199221.
Anton, Susan C. 2003. Natural history of Homo erectus. Yearbook of Physical
Anthropology 46:126170.
. 2008. Framing the question: diet and evolution in early Homo. In
Primate craniofacial function and biology: papers in honor of Bill Hylander.
Christopher J. Vinyard, Christine E. Wall, and Matthew J. Ravosa, eds. Pp.
443482. New York: Springer.
. 2012. Early Homo: who, when, and where. Current Anthropology
53(suppl. 6):S278S298.
. 2013. Homo erectus and related taxa. In Companion to paleoanthropology. David Begun, ed. Hoboken, NJ: Wiley-Blackwell.
Anton, Susan C., William R. Leonard, and Marcia Robertson. 2002. An ecomorphological model of the initial hominid dispersal from Africa. Journal
of Human Evolution 43:773785.
Anton, Susan C., and J. Josh Snodgrass. 2012. Origins and evolution of genus
Homo: new perspectives. Current Anthropology 53(suppl. 6):S479S496.
Asfaw, Berhane, Tim White, Owen Lovejoy, Bruce Latimer, Scott Simpson,
and Gen Suwa. 1999. Australopithecus garhi: a new species of early hominid
from Ethiopia. Science 284:629635.
Berger, Lee R., Darryl de Ruiter, Steven E. Churchill, Peter Schmid, Kristian
J. Carlson, Paul H. G. M. Dirks, and Job M. Kibii. 2010. Australopithecus
sediba: a new species of Homo-like australopith from South Africa. Science
328:195204.
Blumenschine, Robert J., Charles C. Peters, Fidelis T. Masao, Ron J. Clarke,
Alan L. Deino, Richard L. Hay, Carl C. Swisher, et al. 2003. Late Pliocene
Homo and hominid land use from western Olduvai Gorge, Tanzania. Science
299:12171221.
Braun, David R., Thomas Plummer, Peter Ditchfield, Joseph V. Ferraro, David
Maina, Laura C. Bishop, and Richard Potts. 2008. Oldowan behavior and
raw material transport: perspectives from the Kanjera Formation. Journal
of Archaeological Science 35:23292345.
Bribiescas, Richard G., Peter T. Ellison, and Peter B. Gray. 2012. Male life
history, reproductive effort, and the evolution of the genus Homo. Current
Anthropology 53(suppl. 6):S424S435.
Brown, Frank, John Harris, Richard Leakey, and Alan Walker. 1985. Early
Homo erectus skeleton from west Lake Turkana, Kenya. Nature 316:788
792.
Fuentes, Agustn, Matthew A. Wyczalkowski, and Katherine C. MacKinnon.
2010. Niche construction through cooperation: a nonlinear dynamics contribution to modeling facets of the evolutionary history in the genus Homo.
Current Anthropology 51:435444.
Gabunia, Leo, Abesalom Vekua, David Lordkipanidze, Carl C. Swisher, Reid
Ferring, Antje Justus, Medea Nioradze, et al. 2000. Earliest Pleistocene
cranial remains from Dmanisi, Republic of Georgia: taxonomy, geological
setting, and age. Science 288:10191025.
Haile-Selassie, Yohannes, Bruce M. Latimer, Mulugeta Alene, Alan L. Deino,
Luis Gibert, Stephanie M. Melillo, Beverly Z. Saylor, Gary R. Scott, and C.
Owen Lovejoy. 2010. An early Australopithecus afarensis postcranium from
Woranso-Mille, Ethiopia. Proceedings of the National Academy of Sciences of
the USA 107:1212112126.
Hawkes, Kristen, Peter S. Kim, Brett Kennedy, Ryan Bohlender, and John
Hawks. 2011. A reappraisal of grandmothering and natural selection. Proceedings of the Royal Society B 278:19361938.
Hawkes, Kristen, James F. OConnell, Nicholas G. Blurton Jones, Helen Alvarez, and Eric L. Charnov. 1998. Grandmothering, menopause, and the
evolution of human life histories. Proceedings of the National Academy of
Sciences of the USA 95:13361339.
Holliday, Trenton W. 2012. Body size, body shape, and the circumscription
of the genus Homo. Current Anthropology 53(suppl. 6):S330S345.
Holloway, Ralph L., Douglas C. Broadfield, and Michael S. Yuan. 2004. The
human fossil record: brain endocasts: the paleoneurological evidence. New York:
Wiley.
Isler, Karin, and Carel P. van Schaik. 2009. The expensive brain: a framework
for explaining evolutionary changes in brain size. Journal of Human Evolution 57:392400.
. 2012. How our ancestors broke through the gray ceiling: comparative
evidence for cooperative breeding in early Homo. Current Anthropology
53(suppl. 6):S453S465.
Kachel, A. Friederike, Luke S. Premo, and Jean-Jacques Hublin. 2011a. Grandmothering and natural selection. Proceedings of the Royal Society B 278:384
391.

. 2011b. Invited reply: grandmothering and natural selection revisited.


Proceedings of the Royal Society B 278:19391941.
Kaplan, Hillard, Kim Hill, Jane Lancaster, and Magdalena A. Hurtado. 2000.
A theory of human life history evolution: diet, intelligence, and longevity.
Evolutionary Anthropology 2000:156185.
Kaplan, Hillard, Jane Lancaster, and Arthur Robson. 2003. Embodied capital
and the evolutionary economics of the human life span. In Life span: evolutionary, ecological, and demographic perspectives. James R. Carey and Shripad Tuljapurkar, eds. Pp. 152182. New York: Population Council.
Kibii, Job M., Steven E. Churchill, Peter Schmid, Kristian J. Carlson, Nichelle
D. Reed, Darryl J. de Ruiter, and Lee R. Berger. 2011. A partial pelvis of
Australopithecus sediba. Science 333:14071411.
Kimbel, William H., Robert C. Walter, Donald C. Johanson, Kaye E. Reed,
James L. Aronson, Zelalem Assefa, Curtis W. Marean, et al. 1996. Late
Pliocene Homo and Oldowan tools from the Hadar Formation (Kada Hadar
Member), Ethiopia. Journal of Human Evolution 31:549561.
Kivell, Tracy L., Job M. Kibii, Steven E. Churchill, Peter Schmid, and Lee R.
Berger. 2011. Australopithecus sediba hand demonstrates mosaic evolution
of locomotor and manipulative abilities. Science 333:14111417.
Knott, Cheryl D. 1998. Changes in orangutan caloric intake, energy balance,
and ketones in response to fluctuating fruit availability. International Journal
of Primatology 19:10611079.
Kuzawa, Christopher W. 1998. Adipose tissue in human infancy and childhood: an evolutionary perspective. Yearbook of Physical Anthropology 41:
177209.
Kuzawa, Christopher W., and Jared M. Bragg. 2012. Plasticity in human life
history strategy: implications for contemporary human variation and the
evolution of genus Homo. Current Anthropology 53(suppl. 6):S369S382.
Leakey, Louis S., Philip V. Tobias, and John R. Napier. 1964. A new species
of the genus Homo from Olduvai Gorge. Nature 202:79.
Leakey, Meave G., Fred Spoor, M. Christopher Dean, Craig S. Feibel, Susan
C. Anton, Christopher Kiarie, and Louise N. Leakey. 2012. New fossils from
Koobi Fora in northern Kenya confirm taxonomic diversity in early Homo.
Nature 488:201204, doi:10.1038/nature11322.
Leigh, Steven R. 1992. Patterns of variation in the ontogeny of primate body
size dimorphism. Journal of Human Evolution 23:2750.
Leonard, William R., and Marcia L. Robertson. 1997. Comparative primate
energetics and hominid evolution. American Journal of Physical Anthropology
102:265281.
McHenry, Henry M., and Katherine Coffing. 2000. Australopithecus to Homo:
transformations in body and mind. Annual Review of Anthropology 29:125
146.
McPherron, Shannon P., Zeresenay Alemseged, Curtis W. Marean, Jonathan
G. Wynn, Denne Reed, Denis Geraads, Rene Bobe, and Hamdallah A.
Bearat. 2010. Evidence for stone-tool-assisted consumption of animal tissues
before 3.39 million years ago at Dikika, Ethiopia. Nature 466:857860.
Migliano, Andrea Bamberg, and Myrtille Guillon. 2012. The effects of mortality, subsistence, and ecology on human adult height and implications for
Homo evolution. Current Anthropology 53(suppl. 6):S359S368.
Migliano, Andrea B., Lucio Vinicius, and Marta Mirazon Lahr. 2007. Life
history trade-offs explain the evolution of human pygmies. Proceedings of
the National Academy of Sciences of the USA 104:2021620219.
Navarrete, Ana, Carel P. van Schaik, and Karin Isler. 2011. Energetics and the
evolution of human brain size. Nature 480:9193.
OConnell, James F., Kristin Hawkes, and Nicolas B. Blurton Jones. 1999.
Grandmothering and the evolution of Homo erectus. Journal of Human
Evolution 36:461485.
Odling-Smee, F. John, Kevin N. Laland, and Marcus W. Feldman. 2003. Niche
construction: the neglected process in evolution. Monographs in Population
Biology 37. Princeton, NJ: Princeton University Press.
Pfeiffer, Susan. 2012. Conditions for evolution of small adult body size in
southern Africa. Current Anthropology 53(suppl. 6):S383S394.
Plavcan, J. Michael. 2001. Sexual dimorphism in primate evolution. American
Journal of Physical Anthropology 116:2553.
. 2012. Body size, size variation, and sexual size dimorphism. Current
Anthropology 53(suppl. 6):S409S423.
Plavcan, J. Michael, and Carel P. van Schaik. 1997a. Interpreting hominid
behavior on the basis of sexual dimorphism. Journal of Human Evolution
32:345374.
. 1997b. Intrasexual competition and body weight dimorphism in anthropoid primates. American Journal of Physical Anthropology 103:3768.
Pontzer, Herman. 2012. Ecological energetics in early Homo. Current Anthropology 53(suppl. 6):S346S358.

This content downloaded on Wed, 23 Jan 2013 13:36:28 PM


All use subject to JSTOR Terms and Conditions

Aiello and Anton Origins of Homo

S277

Potts, Richard. 2012. Environmental and behavioral evidence pertaining to


the evolution of early Homo. Current Anthropology 53(suppl. 6):S299S317.
Potts, Richard, Anna K. Behrensmeyer, Alan Deino, Peter Ditchfield, and
Jennifer Clark. 2004. Small mid-Pleistocene hominin associated with East
African Acheulean technology. Science 305:7578.
Ruff, Christopher B. 1993. Climatic adaptation and hominid evolution: the
thermoregulatory imperative. Evolutionary Anthropology 2:5360.
. 2010. Body size and body shape in early Homo: implications of the
Gona pelvis. American Journal of Physical Anthropology 58:166178.
Schaller, George B., and Gordon R. Lowther. 1969. Relevance of carnivore
behavior to study of early hominids. Southwestern Journal of Anthropology
25:307341.
Schwartz, Gary T. 2012. Growth, development, and life history throughout
the evolution of Homo. Current Anthropology 53(suppl. 6):S395S408.
Semaw, Sileshi, Michael J. Rogers, Jay Quade, Paul R. Renne, Robert F. Butler,
Manuel Dominguez-Rodrigo, Dietrich Stout, William S. Hart, Travis Pickering, and Scott W. Simpson. 2003. 2.6-million-year-old stone tools and
associated bones from OGS-6 and OGS-7, Gona, Afar, Ethiopia. Journal of
Human Evolution 45:169177.
Shipman, Pat, and Alan Walker. 1989. The costs of becoming a predator.
Journal of Human Evolution 18:373392.
Simpson, Scott W., Jay Quade, Naomi E. Levin, Robert Butler, Guillaume
Dupont-Nivet, Melanie Everett, and Sileshi Semaw. 2008. A female Homo
erectus pelvis from Gona, Ethiopia. Science 322:10891092.
Smith, Jennifer E., Eli M. Swanson, Daphna Reed, and Kay E. Holekamp.

2012. Evolution of cooperation among mammalian carnivores and its relevance to hominid evolution. Current Anthropology 53(suppl. 6):S436S452.
Spoor, Fred, Meave G. Leakey, Patrick N. Gathogo, Frank H. Brown, Susan
C. Anton, Ian McDougall, Christopher Kiarie, F. Kyalo Manthi, and Louise
N. Leakey. 2007. Implications of new early Homo fossils from Ileret, east
of Lake Turkana, Kenya. Nature 448:688691.
Ungar, Peter S. 2012. Dental evidence for the reconstruction of diet in African
early Homo. Current Anthropology 53(suppl. 6):S318S329.
Walker, Alan, and Richard Leakey, eds. 1993. The Nariokotome Homo erectus
skeleton. Cambridge, MA: Harvard University Press.
Wells, Jonathan C. K. 2009. The evolutionary biology of human body fat: thrift
and control. Cambridge: Cambridge University Press.
. 2012. The capital economy in hominin evolution: how adipose tissue
and social relationships confer phenotypic flexibility and resilience in stochastic environments. Current Anthropology 53(suppl. 6):S466S478.
Wood, Bernard, and Jennifer Baker. 2011. Evolution in the genus Homo.
Annual Review of Ecology, Evolution, and Systematics 42:4769.
Wood, Bernard A., and Mark C. Collard. 1999. The human genus. Science
284:6571.
. 2007. Defining the genus Homo. In Handbook of paleoanthropology.
Winfried Henke, Helen Rothe, and Ian Tattersall, eds. Pp. 15751610. Berlin: Springer.
Zipfel, Bernhard, Jeremy M. DeSilva, Robert S. Kidd, Kristian J. Carlson,
Steven E. Churchill, and Lee R. Berger. 2011. The foot and ankle of Australopithecus sediba. Science 333:14171420.

This content downloaded on Wed, 23 Jan 2013 13:36:28 PM


All use subject to JSTOR Terms and Conditions

Você também pode gostar