ChapterPDF Available

Hominin Footprints in Caves from Romanian Carpathians

Authors:
  • University of South Florida / Babes-Bolyai University
  • Academia Romana Cluj

Abstract

The Romanian karst hosts numerous caves and shelters that over time provided remarkable archaeological and anthropological vestiges. Altogether they show that humans must have entered caves in Romania at least as early as 170,000 years ago. However, ancient human footprints are very rare in the fossil record of East-Central Europe, with only two known locations in the Apuseni Mountains of western Romania. Vârtop Cave site originally preserved three fossil footprints made about 67,800 years ago by a Homo neanderthalensis , whereas Ciur Izbuc Cave was probably home of early H. sapiens that left almost 400 footprints (interspersed with spoors of cave bears), which were indirectly dated to be younger than ~36,500 years.
Reading Prehistoric
Human Tracks
Andreas Pastoors
Tilman Lenssen-Erz
Editors
Methods & Material
Reading Prehistoric Human Tracks
Andreas Pastoors Tilman Lenssen-Erz
Editors
Reading Prehistoric Human
Tracks
Methods & Material
Editors
Andreas Pastoors
Institut für Ur- und Frühgeschichte
Friedrich-Alexander-Universität
Erlangen-Nürnberg
Erlangen, Germany
Tilman Lenssen-Erz
African Archaeology
University of Cologne
Cologne, Germany
ISBN 978-3-030-60405-9 ISBN 978-3-030-60406-6 (eBook)
https://doi.org/10.1007/978-3-030-60406-6
©The Editor(s) (if applicable) and The Author(s) 2021. This book is an open access publication.
Open Access This book is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation,
distribution and reproduction in any medium or format, as long as you give appropriate credit to the
original author(s) and the source, provide a link to the Creative Commons license and indicate if changes
were made.
The images or other third party material in this book are included in the books Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the books Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted
use, you will need to obtain permission directly from the copyright holder.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specic statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional afliations.
Cover illustration: Reading prehistoric human tracks in Tuc dAudoubert (Photo and Copyright holder
Association Louis Bégouën/Tracking in Caves)
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Foreword
It is a great honour and pleasure for me to congratulate the organizers of this
conference and its volume for having brought forward such an innovative approach
and topic. It was a fantastic idea to invite expert trackers for an international
conference on human tracks, to offer them the possibility to meet other trackers
from hunter-gatherer communities around the globe, and to open pathways for
including indigenous experts into archaeological research. This shows that there is
a kind of knowledge beyond the academic knowledge that is able to enrich science.
This conference was somehow an experiment, but a very successful one. To deal
with new categories of knowledge beyond the classical western academic knowl-
edge is extremely challenging, and it is part of the intangible heritage of mankind.
The Humboldt Forum in Berlin will become a place where cultures from all over the
world shall meet and get into exchange, where a new dialogue between cultures can
be developed by cooperation and by co-productions, and where we want to dene a
new understanding of shared heritage and shared history. This is not only a great
challenge, but also a unique chance.
Traditional or indigenous knowledge is so important, because these knowledge
systems are embedded in the cultural traditions of regional, indigenous, or local
communities, it is knowledge acquired over many generations, it is knowledge
mostly about traditional technologies of subsistence, ecological knowledge, tradi-
tional medicine, climate etc., and it is generally based on accumulations of empirical
observation and on interaction with the environment. This traditional knowledge
may distinguish one community from another, it takes on personal and spiritual
meanings, and it can reect the communitys interests.
Communities depend sometimes on their traditional knowledge, especially on
environmental issues, their knowledge is bound to ancestors and ancestral lands, and
it is embedded in a cosmology and therefore has a spiritual component, too.
Communities have strong traditions of ownership or custodianship over knowledge,
the misuse of knowledge may be offensive to traditions, and they prevent the
patenting of traditional knowledge by not expressing consent.
v
vi Foreword
In the broader context traditional knowledge has to be treated in the same way as
other traditional cultural expressions. The World Intellectual Property Organization
(WIPO) interprets traditional knowledge as any form of artistic and literary expres-
sion in which traditional culture and knowledge are embodied. This knowledge is
transmitted from one generation to the next, and it includes handmade textiles,
paintings, stories, legends, ceremonies, music, songs, rhythms and dance.
During the preparation of the Humboldt Forum in Berlin, it is interesting that the
inclusion of indigenous knowledge becomes more and more important and interest-
ing. Years ago we started the project Sharing knowledgewith the Indigenous
University of Tauca in Venezuela, which in the meantime expanded into
neighbouring regions of Brazil and Colombia. This cooperation makes visible the
dynamics and presence of indigenous perspectives on ethnographic objects, it helps
in writing the history of the collections again by including the indigenous perspec-
tive. Through an online-platform the future visitor of the Humboldt Forum gets rst-
hand knowledge from the indigenous perspective on the objects, and not ethnolo-
gists or anthropologists are speaking for the indigenous, but the indigenous speak for
themselves, what we call multivocality. Ethnologists and anthropologists remain
only in an intermediate position. This is a way of decolonizing perspectives by
sharing the power of interpretation.
In these days we talk a lot about decolonizing museums and also decolonizing the
archaeological practice. These questions are addressing issues of power of science
and control of archaeological interpretation. We need participatory approaches, and
we have to develop new methodologies and strategies of community participation.
This kind of community engagement can be a new path into the future of archaeol-
ogy in Africa and beyond. It also can help in reacting towards rapid environmental
changes affecting ecosystems by engaging communities throughout all levels of
research.
But local communities demand to get something back, e.g. the San people in
Southern Africa, Inuit in Alaska, First Nations in Canada, or Aborigines in Australia.
They dened codes of ethics for researchers wishing to study their culture, their
knowledge, their genes or their heritage. They have to be treated respectfully without
publishing insulting information, communities wish to read and check results before
publication to avoid misunderstandings, and they have to have free access to
research data.
Dealing with indigenous knowledge can help us a lot to learn more about a distant
past, but it is also a unique chance to broaden our understanding of the plurality of
cultures today, and that there are very different categories of knowledge. More
knowledge, however, is an important step towards more tolerance and respect for
other cultures and different traditions, what maybe today is more important
than ever.
President of the Stiftung Preussischer
Kulturbesitz, Berlin, Germany
December 2019
Hermann Parzinger
Preface
In May 2017 a conference was hosted at the University of Cologne and the
Neanderthal Museum that covered the topic of prehistoric human tracks in a truly
global perspective: it convened experts from ve continents as well as from various
disciplines for scientic presentations. Besides the usual academic presentations in a
lecture hall a full day was dedicated to discussing with and listening to indigenous
tracking experts from Australia, Canada and Namibia around a re outside. These
talks and practical demonstrations of track reading by the indigenous tracking
experts on a track eld with human footprints aimed at enabling western scholars
to get a glimpse of the methodological basics of expert tracking. For indigenous
trackers it is common practice not only to discriminate male from female footprints
but they can also distinguish age classes of adult persons a differentiation western
science including orthopaedics is unable to achieve. This knowledge now entered
into a discourse with scientic approaches to glean information from human
footprints.
Nearly all projects worldwide investigating human tracks in archaeological
context were present at the conference, covering a time span from the earliest
footprints in Laetoli to Neolithic ones on the Danish coast. Methodological aspects
presented a range from collaboration with indigenous trackers to visualizations
based on state of the art scanning technology. This extraordinary meeting with its
rst time ever encounter of all kinds of human ways of knowing on an archaeological
source material an under-researched one at that called for an dissemination
beyond the closed circle of experts who were present at the conference. The idea of
capturing all this knowledge in a book was cogent and in the process of production it
showed that further aspects that were not represented at the conference, should still
be included so that here we also present authors who did not contribute to the
conference. Through this selection of authors for the rst time the most important
sites which were found worldwide, will be published in a single publication.
This and the broad scope of methodological diversity will make the book a
rewarding read for readers from a wide range of elds of knowing. The analysis of
human tracks by representatives of anthropological, statistical and traditional
vii
approaches feature the multi-layered methods available for the analysis of human
tracks and will appeal to students, scholars and also laypeople with an interest in
archaeologies, anthropology, social anthropology, palaeontology, cognitive science,
cultural science, ichnology and sports science. This book is to show that progress in
science and enlightenment on the one hand requires the development of ever new
methods in order to enhance the ability for ne resolution in measurement and
interpretation of phenomena, but on the other hand it also shows that recourse to
knowledge and skills that may have been our human toolkit throughout out species
history can point out where we should get at with our scientic approaches.
viii Preface
Erlangen, Germany Andreas Pastoors
Cologne, Germany Tilman Lenssen-Erz
Acknowledgments
First of all, we would like to thank the institutions and people who made this book
possible: The authors and the publisher for their collegial and professional cooper-
ation and the sponsors for their generous nancial support. Foremost, we would like
to mention the Volkswagen Foundation. Further funds came from goldschmidt and
the Heinrich Barth Institute. The book is integrated into a research project supported
by the German Science Foundation (UT 41/6-1). Thanks are due to the universities
of Erlangen-Nuremberg and Cologne and the Association Louis Bégouën for their
institutional support and helpfulness. As mentioned above, the book is based on the
international Prehistoric Human Tracks conference which was held in cooperation
with the German Commission for UNESCO in Cologne/Mettmann (May 11-13,
2017). We would like to thank all participants and the donors: Volkswagen Foun-
dation, GO-AIDE Foundation, Neanderthal Museum Foundation, University of
Cologne, Cultures and Societies in Transition (Competence Area IV), Association
Louis Bégouën and Heinrich-Barth-Institute.
Erlangen, Germany Andreas Pastoors
Cologne, Germany Tilman Lenssen-Erz
February 2020
ix
Contents
1 Introduction .......................................... 1
Andreas Pastoors and Tilman Lenssen-Erz
Part I Methodological Diversity in the Analysis of Human Tracks
2 Inferences from Footprints: Archaeological Best Practice ........ 15
Matthew R. Bennett and Sally C. Reynolds
3 Repetition Without Repetition: A Comparison of the Laetoli G1,
Ileret, Namibian Holocene and Modern Human Footprints Using
Pedobarographic Statistical Parametric Mapping .............. 41
Juliet McClymont and Robin H. Crompton
4 Reproduce to Understand: Experimental Approach Based
on Footprints in Cussac Cave (Southwestern France) ........... 67
Lysianna Ledoux, Gilles Berillon, Nathalie Fourment, and Jacques
Jaubert
5 Experimental Re-creation of the Depositional Context
in Which Late Pleistocene Tracks Were Found on the Pacic
Coast of Canada ....................................... 91
Duncan McLaren, Quentin Mackie, and Daryl Fedje
6 Reading Spoor ......................................... 101
Tilman Lenssen-Erz and Andreas Pastoors
Part II Case Studies from Around the Globe
7 Perspectives on Pliocene and Pleistocene Pedal Patterns
and Protection ......................................... 121
Erik Trinkaus, Tea Jashashvili, and Biren A. Patel
xi
xii Contents
8 Frozen in the Ashes ..................................... 133
Marco Cherin, Angelo Barili, Giovanni Boschian,
Elgidius B. Ichumbaki, Dawid A. Iurino, Fidelis T. Masao,
Soa Menconero, Jacopo Moggi Cecchi, Susanna Sarmati,
Nicola Santopuoli, and Giorgio Manzi
9 Steps from History ..................................... 153
Nick Ashton
10 Reconsideration of the Antiquity of the Middle Palaeolithic
Footprints from Theopetra Cave (Thessaly, Greece) ............ 169
Nina Kyparissi-Apostolika and Sotiris K. Manolis
11 On the Tracks of Neandertals: The Ichnological Assemblage
from Le Rozel (Normandy, France) ......................... 183
Jérémy Duveau, Gilles Berillon, and Christine Verna
12 Hominin Footprints in Caves from Romanian Carpathians ....... 201
Bogdan P. Onac, Daniel S. Veres, and Chris Stringer
13 Episodes of Magdalenian Hunter-Gatherers in the Upper
Gallery of Tuc dAudoubert (Ariège, France) ................. 211
Andreas Pastoors, Tilman Lenssen-Erz, Tsamgao Ciqae, /Ui Kxunta,
Thui Thao, Robert Bégouën, and Thorsten Uthmeier
14 Following the Father Steps in the Bowels of the Earth:
The Ichnological Record from the Bàsura Cave
(Upper Palaeolithic, Italy) ................................ 251
Marco Avanzini, Isabella Salvador, Elisabetta Starnini,
Daniele Arobba, Rosanna Caramiello, Marco Romano, Paolo Citton,
Ivano Rellini, Marco Firpo, Marta Zunino, and Fabio Negrino
15 Prehistoric Speleological Exploration in the Cave of Aldène
in Cesseras (Hérault, France): Human Footprint Paths
and Lighting Management ............................... 277
Philippe Galant, Paul Ambert, and Albert Colomer
16 The Mesolithic Footprints Retained in One Bed of the
Former Saltmarshes at Formby Point, Sefton Coast,
North West England .................................... 295
Alison Burns
17 Prehistoric Human Tracks in Ojo Guareña Cave System
(Burgos, Spain): The Sala and Galerías de las Huellas .......... 317
Ana I. Ortega, Francisco Ruiz, Miguel A. Martín,
Alfonso Benito-Calvo, Marco Vidal, Lucía Bermejo,
and Theodoros Karampaglidis
Contents xiii
Part III Experiences with Indigenous Experts
18 Tracking with Batek Hunter-Gatherers of Malaysia ............ 345
Tuck-Po Lye
19 Identify, Search and Monitor by Tracks: Elements of Analysis
of Pastoral Know-How in Saharan-Sahelian Societies ........... 363
Laurent Gagnol
20 TrackersConsensual Talk: Precise Data for Archaeology ....... 385
Megan Biesele
21 An Echo from a Footprint: A Step Too Far ................... 397
Steve Webb
22 Walking Together: Ways of Collaboration in Western-Indigenous
Research on Footprints .................................. 413
Hannah Zwischenberger
Chapter 1
Introduction
Andreas Pastoors and Tilman Lenssen-Erz
Abstract This book explains that after long periods of prehistoric research in which
the importance of the archaeological as well as the natural context of rock art has been
constantly underestimated, research has now begun to take this context into focus for
documentation, analysis, interpretation and understanding. Human footprints are
prominent among the long-time under-researched features of the context in caves
with rock art. In order to compensate for this neglect an innovative research program
has been established several years ago that focuses on the merging of indigenous
knowledge and western archaeological science for the benet of both sides. The book
composes rst the methodological diversity in the analysis of human tracks. Here
major representatives of anthropological, statistical and traditional approaches feature
the multi-layered methods available for the analysis of human tracks. It second
compiles case studies from around the globe of prehistoric human. For the rst time
the most important sites which have been found worldwide are published in a single
publication. The third focus of this book is on rst hand experiences of researchers
with indigenous tracking experts from around the globe, expounding on how archae-
ological science can benet from the ancestral knowledge.
Keywords Prehistoric human tracks · Methodological diversity · Indigenous
tracking
Prehistoric human tracks entered into archaeology on a side track more than
100 years ago when human footprints from the Ice Age were discovered in 1906
in the Palaeolithic cave of Niaux in southern France (Cartailhac and Breuil 1907:
222, 1908: 44; Pales 1976):
A. Pastoors (*)
Institut für Ur- und Frühgeschichte Friedrich-Alexander-Universität Erlangen-Nürnberg,
Erlangen, Germany
e-mail: andreas.pastoors@fau.de
T. Lenssen-Erz
African Archaeology, University of Cologne, Cologne, Germany
e-mail: lenssen.erz@uni-koeln.de
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_1
1
Ajoutons quen deux points épargnés par les pieds des visiteurs modernes, nous avons noté,
à la surface dun sol analogue, mais un peu moins ferme, lempreinte des genoux nus dun
homme qui avait rampé sous une voûte basse, et celles de nombreux pieds également nus,
appartenant à des adultes et à des enfants. (Cartailhac and Breuil 1907: 222)
1
2 A. Pastoors and T. Lenssen-Erz
But the interest in these sources was a rather modest one since only Bégouën (1928)
and Vallois (1928,1931) made scientic studies on them, while many tracks in other
sites were destroyed without recording. Archaeologists treated the remaining tracks
similar to most other sources they deal with: measuring, recording, copying and
casting were the means applied to get at a deepened understanding. Tracking,
i.e. reading of tracks, was not applied so that this realm of knowledge made its
rst appearance in academia only in 1990 with Louis Liebenbergs book The Art of
Tracking, the Origin of Science and yet the insights of this book remained a
dormant potentiality for unjustiably long time. It was only from the rst decade of
the twenty-rst century onwards when more and more scholars and projects turned
their attention towards prehistoric human tracks thus attempting to catch up with
ichnology which for a long time had developed as a specialized eld of research,
mainly coming from the analysis and interpretation of dinosaur tracks (Lockley
1999). Interpretation of tracks in criminal forensics had taken its own, isolated
development (Matthews David 2019) before archaeologists and forensic specialists
pooled their accumulated knowledge and experiences (Bennett and Budka 2019).
With these turns in research strategies, it was acknowledged that human tracks are an
important contextual source for the understanding of peoples behaviour in the past
which previously had mainly concentrated on the sensational footprint nds at
Laetoli in Tanzania (Leaky and Harris 1987). Besides learning from these earliest
footprints about the development of bipedal locomotion the understanding of human
behaviour was of particular interest in the Palaeolithic caves harbouring master-
pieces of prehistoric art such as Niaux, Pech-Merle, Tuc dAudoubert or Chauvet-
Pont dArc Cave. But also many other sites around the globe with fossilized human
tracks gained growing attention (Lockley et al. 2008,2016; Pasda 2013) and
experienced the application of state-of-the-art technology for documentation and
analysis (Bennett et al. 2009,2016; Crompton et al. 2011). However, scientic
methods do not attain much deeper insights than concluding the body height of a
person, where the footprint length represents 15% of body height (the formula is
virtually unchanged since Topinard 1877), but as Bennett and Morse (2014: 148)
point out, there lies vast fuzziness in these results. Nevertheless this estimated size is
from which an educated guess of the age of the person is made (Bennett and Morse
2014: 152154). Because of these shortcomings of scientic methods, some projects
turned to involve indigenous trackers in prehistoric human spoor interpretations
(e.g. Webb et al. 2006; Pastoors et al. 2015), and this conrmed the known but
hitherto neglected ability to glean deeper information from footprints (Liebenberg
1990; Biesele and Barclay 2001; Lowe 2002; Gagnol 2013; for the reliability of
1
Let us add that at two points not affected by the steps of modern visitors, we noted, on the surface
of a similar but slightly less rm ground, the bare knee prints of a man who had crawled under a low
arch, and those of many equally bare feet, belonging to adults and children(translated by the
authors).
indigenous track reading see Stander et al. 1997. Wong et al. 2011). Wherever
indigenous specialists were involved as ichnologists, they were able to considerably
augment the insights about human behaviour at a site thus showing the rich potential
of information resting in these sources, if adequately well preserved. Expectable
critique of these analyses and interpretations points out the lack of testability and
validation (e.g. Bennett and Budka 2019: 155), but scientic methods, for their part,
are presently unable to provide dependable falsication with their proper methods,
which would demonstrate their supremacy. Instead, interpretation with a large team
of scientists of complex tracks seemingly remaining from human/animal interaction
(possibly a hunt) in the Pleistocene (Bustos et al. 2018) eventually has to turn to
speculation about intentions and behaviour of humans and animals in order to nd a
cogent narrative for what the tracks preserve of an event.
1 Introduction 3
Examples of Indigenous Spoor Interpretation
The list of prehistoric sites mentioned in this volume (Happisburgh, Bàsura Cave,
Formby Point, Laetoli, Le Rozel, Calvert Island, Vârtop Cave, Ciur-Izbuc Cave,
Aldène, Theopetra Cave, Ojo Guareña Cave system and Willandra) where scientic
methods have been applied clearly shows that the identication of the trackmakers
by morphometric analyses is not sufcient to capture the potential of the dynamic
processes stored in the spoors. At each of these sites, more or less complex events
were hypothesized but as was to be expected their accuracy and scope varies due to
the personal experience of the respective authors. This procedure constitutes the
unspoken application of the pre-iconographic description in western art according to
Panofsky (Panofsky 1962). Practical experience (familiarity with objects and phe-
nomena) is an absolute prerequisite for a successful application of the
pre-iconographic description, from which a positive correlation between experience
and descriptive accuracy can be derived. In the Tracking in Caves project carried out
in Tuc dAudoubert, the outstanding experience in reading tracks by indigenous
ichnologists was used (Pastoors and Lenssen-Erz 2020; see Lenssen-Erz and
Pastoors Chap. 6; Pastoors et al. Chap. 13). Their expertise was applied not only
to the prehistoric spoors in Tuc dAudoubert but also in the caves of Niaux, Pech-
Merle and Fontanet.
In Niaux Cave (Ariège, France) 38 footprints are known in a small diverticule.
Western academic analysis found some order in an initially seemingly chaotic
distribution of footprints by identifying two to three subjects with an age of
912 years (Pales 1976:9293). The indigenous ichnologists saw an unequal
number of footprints and identied a girl (713 years; age classes according to
Martin 1928) as their sole trackmaker. The spoors were executed in a controlled, not
a chaotic manner and in an upright body posture, which is a puzzle since the ceiling
is too low to stand upright (Pastoors et al. 2015).
The cave of Pech-Merle (Lot, France) reveals a total number of 17 footprints. Last
western academic analysis interpreted the spoors as the result of one single
trackmaker, a big child, adolescent or a small adult (Duday and García 1983). The
4 A. Pastoors and T. Lenssen-Erz
indigenous ichnologists identied ve subjects with an age between infans II
(713 years) to maturus (4160 years) (Pastoors et al. 2017). They saw four adults,
two male and two female and one younger male (713 years) crossing the location
separately independent of each other. Furthermore, they detected two events
deviating from normal walking: subject S5, a female adult, carried additional weight,
and subject S3, a boy of 910, turns left abruptly.
The third cave that has been briey surveyed by the indigenous ichnologists is
Fontanet (Ariège, France). Due to various circumstances, the exact number of
prehistoric spoors is unknown. In any case, no complete western academic analysis
has yet taken place. Currently Lysianna Ledoux is working on a complete inventory
of the spoors. First results are available for three track elds of different sizes
(Ledoux 2019). Accordingly, on the largest of the three areas, plage 1, 62 tracks
were inventoried (identied and measured). Beside footprints there are some hand-
prints and especially numerous slipping marks. The number of trackmakers is
assumed to be between two and six subjects, including children, on the basis of
metric analyses. In addition to recording the identity of the trackmakers, Ledoux is
also concerned with the identication of events. As an example, three tracks suggest
a squatting position, extending on the feet, and the left hand resting back against the
ground (Ledoux 2019: 253). The indigenous ichnologists counted on 2 study areas
(plage 1 and plage 3 at Ledoux 2019) in Fontanet a total 28 prehistoric human traces
(27 footprints and one knee) of 17 subjects that could be combined to a total of
8 trackways, which made up 15 events (Pastoors et al. 2015). Among them there are
six men on plage 1, two women, one boy, three girls and one unspecic male
(covering altogether an age from infans I to maturus). On plage 3 there are four
subjects, all male, between juvenis (1420 years) and maturus (4160 years). In
addition to the information on the identity of the trackmakers, the experienced
trackers were able to identify some special events apart from normal walking. On
plage 1 subject S5 had slipped, subject S6 was going fast and subject S10 was
kneeling. Then a group consisting of four subjects was identied, who were walking
together. These are subject S1, female adultus, subjects S2 (male infans II), S3
(female infans II) and S4 (female infans I). In addition, plage 3 was exclusively
identied as an area of normal walking. No footprint shows a direct relation to the
results of the drawing activities carried out on the ground.
What we have here are results of an analysis by indigenous experts that is part of a
daily practice in many pre-industrialized societies. Information on contemporaries,
their whereabouts and their doings, in these societies is independent of self-observed
evidence or reporting and information. What those who grew up in industrialized
societies with paved and tarred surfaces in most of their life-world may read from the
face of a person (known or unknown) is also frozen in footprints, irrespective of the
wearing of shoes or not (pers. comm. Kxunta; Gagnol 2013; see Gagnol Chap. 19)
and therefore readable for those with developed tracking skills. Since there can be no
doubt that this depth of information gleaned from tracks is being disclosed not by
trancing, dreaming, hallucinating or vision but instead by a positivist approach to the
analysis of hard data i.e. an immediate intuitive assessment of complex measures
and textures as well as of biological, zoological, hydrological, meteorological,
pedological, cultural, social, sedimentological and physical context it should in
1 Introduction 5
the long run also be detectable by scientic means. Tracking is a parascientic
process implying reasoning in an analogous way to western sciences, using induc-
tion, deduction and abduction in order to generate new knowledge (Liebenberg
1990). While western morphometric approaches restrict themselves to inductive
methodology, indigenous trackers, whose approach can be labelled morpho-
classicatory, can imply induction, deduction or abduction depending on data
quality, as was mentioned above.
With this book we want to fathom how far scientic and indigenous ichnology
have advanced towards their meeting point where both can fully and competently
assess the results of the other. Since tracking does not take recourse to alien types of
rationality, logic or causality and by no means includes any esoteric facets, practi-
tioners of scientic ichnology may nd a useful guide in this book for the recogni-
tion of and the advancement towards indigenous ichnology which shows the
potential of what can be gleaned from tracks, while still continuing exchange with
colleagues from around the globe.
On this Book
Tracks are probably the oldest element of human perception that has been the object of
expert analysis ever since humans hunt. Homo sapiens is not a born successful hunter
of any sizeable game since we have a comparatively poor eyesight (also missing the
Tapetum lucidum that makes many animals seeing well in the dark), a rather poor
sense of smelling and we would be too slow, clumsy and harmless for successful
hunting if not intelligence came into play. Everything beyond a turtle poses a true
challenge if we want to get it alive. Therefore reading tracks will probably during the
whole human history have been an important means and advantage for the procure-
ment of fresh meat; it would have been an existential necessity for every adult person
to acquire solid knowledge in all disciplines of environmental sciences. Consequently
Liebenberg (1990)identied tracking as the origin of science.
Adding to this rst appraisal of the analytic and even epistemic value of the
human ability to read tracks, we want this book to provide a state-of-the-art collec-
tion of chapters that represent the best contributions to the eld of track analysis at
the beginning of the third decade of the twenty-rst century. In this digital epoch,
there are sophisticated technological solutions to grasp all attributes that characterize
tracks, and contributions from around the world show how these are being
implemented in many places. Besides this welcome development and enrichment,
there is another, paradoxical development in which many indigenous traditions are
on the verge to disappear, while it is only now that western science understands that
these traditions harbour irretrievable treasures of knowledge for the understanding of
certain archaeological source. The patron of the conference on Prehistoric Human
Tracks in Cologne and Mettmann 2017, Hermann Parzinger, expounds in his
foreword to this book on this point, emphasizing that indigenous knowledges belong
to the toolkit with which people master living not only survival in all kinds of
environments, based on accumulated knowledge inherited from the ancestors, a
topic to which the new Humboldt Forum in Berlin will dedicate considerable space.
It is the aim of this book to give a comprehensive overview of the investigation of
human footprints in terms of methods and of locations and enriching these with
perspectives on tracks from various indigenous groups. Addressing the main ses-
sions of the conference on Prehistoric Human Tracks, this book is divided in the
three parts:
6 A. Pastoors and T. Lenssen-Erz
Part I Methodological diversity in the analysis of human tracks
Part II Case studies from around the globe
Part III Experiences with indigenous experts
Part I, the methodological part of this book, covers three principal aspects of
archaeological research with chapters on technical means, on experimental archae-
ology and on the attempt to open research towards new knowledge systems. Bennett
and Reynolds give a welcome overview of the technical means that are developed
today and additionally provide a useful array of different ways of how to visualize
data or evidence on tracks (Chap. 2). Meritoriously they also provide a checklist for
running eld research on tracks.
Among the ultimate goals and challenges of the various digital methods is the
ability to discriminate tracks of an individual from those of co-occurring individuals.
As McClymont and Crompton point out in their following chapter, two imprints of a
foot of one person are never identical so that the fuzziness of an imprint needs to
become part of the formula by which an individual can be pinned down by his or her
footprint transposed to data (Chap. 3). Besides the information on an individual,
footprints also freeze information about locomotion processes and about the char-
acter of the locomotion.
Important means of archaeology to generate insight into processes and phenom-
ena are experimental renditions. From the working group of Cussac Cave in south-
western France Ledoux and her co-authors report about their endeavours to better
understand taphonomic processes inside a karst cave (Chap. 4). Importantly they
focus on the effects of intermittent oodings which are a common phenomenon in
caves. McLaren and co-authors also describe experiments by which they not only
re-created footprints in clayey ground but also controlled how plant remains and
macrofossils became imprinted in the ground by stepping on (Chap. 5). By covering
the footprints with sand and excavating them experimentally, inferences about the
depositional conditions in the Late Pleistocene were corroborated.
The nal chapter of this rst part of the book by Lenssen-Erz and Pastoors takes
an encompassing epistemological view of the art of tracking as parascientic
practice (Chap. 6). Doubts in indigenous expertsinferences would be very obvious
and justied should they arrive at results that contradict any reasonable expectations
of which people may have entered the caves and how they behaved there. However,
the tracking experts simply augment the depth of exploration of the data, i.e. they
interpret the track with its visible attributes, rening the results and expectations of
scientic researchers. This cannot be characterized as being unscientic simply
because no scientic discipline presently has the means to disprove them, but instead
it is a lack of series of measurements which the sciences will keep on suffering from
before they arrive at an equally dependable resolution.
1 Introduction 7
Part II of the book, dealing primarily with prehistoric track sites from around the
globe, opens with an instructive chapter by Trinkaus and co-authors about how to
analyse and interpret various elements on skeletal foot remains (Chap. 7). It is
through the combined assessment of these accumulated details which makes the
authors conclude that many imprints of bare feet, which are the normal case in
prehistory, yet retain the markers to identify the consistent use of protective foot-
wear. While this chapter covers a wide range of periods of human evolution, the
following chapters are ordered chronologically, beginning with the hominin foot-
prints of Laetoli. Being the prototype of prehistoric human tracks, it is a welcome
contribution of Cherin and co-authors that they review the rather long history of
research on these tracks, connecting it to the present where digital methods and
scanning have become state of the art in research (Chap. 8). What Laetoli is for
Africa, Happisburgh is for Europe, but even though they are considerably younger,
they were an ephemeral phenomenon. While they could not be preserved in place
due to tidal activities, their preservation and afterlife, as it were, are not only secured
in archaeology but also in the arts, as Ashton exemplies with citations from a poem
and a popular book on walking (Chap. 9).
The four human footprints of Theopetra Cave in Greece, according to the authors
Kyparissi-Apostolika and Manolis, are the oldest European tracks that arguably
could be either of Neandertal origin or early Homo sapiens (Chap. 10). They seem
to originate from two young children of whom one is assumed to have worn
footwear, thus supporting the postulate of Trinkaus and co-authors. More and
undisputed Neandertal tracks are reported from Le Rozel from French Normandy
by Duveau and co-authors (Chap. 11). The sheer mass of more than 250 footprints at
this site, sided by a number of handprints, makes this an exceptional site for the
understanding of Neandertal behaviour and group life of about 80,000 BP. Only
some 13,000 years younger and therefore also of Neandertal origin are footprints that
Onac and co-authors present from Vârtop Cave in Romania together with a
plethora of younger footprints of Homo sapiens found in Ciur-Izbuc Cave, also in
the Carpaths (Chap. 12). A cave with an equally large number of Pleistocene
footprints is Tuc dAudoubert in the French Pyrenees, presented by Pastoors and
co-authors (Chap. 13). In this cave the track reading of indigenous trackers was
practiced most meticulously (only the cave of Aldène received equally intense
investigations, but this is still unpublished) and the results, presented in a systema-
tized scheme, allow to follow, as it were, certain individuals through the cave. They
seem to have undertaken a one-time exploration of the cave system during which
some of them procured certain materials, e.g. bear teeth. Interestingly also in the Late
Pleistocene, a similar one-time visit into a deep cave was paid by a small group of
individuals to Bàsura Cave in Italy, as presented by Avanzini and co-authors
(Chap. 14). And again, same as in Tuc dAudoubert, here, too, not only adults but
also adolescents and even very small children were part of the exploring group.
Another parallel between the two cave visits is that even difcult passages where
crawling or dangerous climbing is required did not prevent the groups from bringing
the small children, and none of the visitors in either group who left imprints wore
shoes nor leg clothes. A well comparable case to the two mentioned ones is reported
by Galant et al. from the Mesolithic period from Aldène Cave in Hèrault region of
southern France (Chap. 15). Again, visitors of all ages and both sexes left their
imprints and all seem to have been in the cave only once apparently for the purpose
of exploration only. Interestingly, conditions in the cave have preserved many traces
of the lighting management that the Mesolithic explorers had to implement. Since
with regard to lighting they had no technological advantage over the Late Pleisto-
cene, this practice can be taken as a potential model for comparable explorations
during earlier periods. These insights will in the near future be associated with the
results of the investigations by indigenous experts in this cave.
8 A. Pastoors and T. Lenssen-Erz
More evidence on secular behaviour in the Mesolithic comes from the footprints at
Formby Point on the English Coast on the Irish Sea on which Alison Burns expounds
(Chap. 16). Here human and animal tracks are mixed, and this appears to reveal
consciousness of the tracks in the people who thus articulated in behaviour their
relationship to animals and to the landscape. As other track sites on coasts, the Formby
tracks were bound to disappear once they had been exposed whereas caves can
preserve footprints for millennia. This is the case to a large extent in the huge Ojo
Guareña Cave system in northern Spain, presented by Ortega and co-authors
(Chap. 17). While this cave preserves evidence from the Upper Palaeolithic onwards,
tracks are dated to the Chalcolithic period at around 4300 calBP and that is why this
chapter concludes the chronologically ordered part of the book. The partially sandy
sediment on the oor shows far more than 1000 human footprints in various places,
which constitute vast areas that cannot be explored without destroying tracks. In some
parts of the cave system, tracks show again a one-time exploration as the purpose of
the visit, while other parts convey clear evidence of several visits apparently relating
to the dark zones of the cave as symbolic and social landscape.
Part III of this book intends to encourage an opening of discourse from a closed
academic environment to other ways of knowing in that besides the conventional
realms of academic exchange we also aim at presenting experiences of researchers
from their encounters with indigenous experts, be they hunters or herders. The
chapters clearly show that such encounters instil a rather personal and emotional
but also humble relationship between experts from different worlds of knowing
perhaps partly because the close insight of western scholars into a special eld of
knowledge in a culture without books and formal teaching shows that meticulous
analysis and understanding of phenomena in our world with the human sense can
surpass any technological apparatus. While tracking usually works with the optical
sense for the analysis and interpretation of visual signs, Lye in her contribution on
Batek from the Malayan rainforest shows that the hearing sense can take precedence
over seeing should the environment require this (Chap. 18). But as in visual tracking,
hunters tracking sounds with the ear reach a ne resolution that seems virtually
impossible to the layperson. In addition, in the rainforest also olfactory traces need to
be carefully included. Since the human tracking ability typically has a connotation
with spoor recognition by hunter-gatherers, the contribution of Lye is a welcome
broadening of scope regarding the senses that can be involved.
1 Introduction 9
Another broadening of scope comes from the second chapter of this part by Gagnol
who looks at the tracking abilities of nomadic pastoralists of the Sahara, where he has
done tracking research (Chap. 19). This is the rst source (based on Gagnol 2013)
where the same stunning tracking abilities that are known from hunter-gatherers
are well and comprehensively documented among peoples with other subsistence
strategies. Just as with hunter-gatherers, the camel herders of the Sahara, too, have
equally sophisticated tracking skills regarding animal or human tracks. For the latter
the particularly skilled experts maintain that even social aspects can be read from
tracks. In addition, Gagnol independently found out that expert tracking also requires
the mastering of abductive reasoning (cf. Liebenberg 1990), labelled hodological
strategy in his chapter, thus providing unprejudiced corroboration of Liebenbergs
postulate that tracking at least partly is based on a scienticmindset.
Other quasi-epistemological aspects of tracking are raised in Bieseles chapter
that draws on her decades of experiences with and living among hunter-gatherer
trackers in the Kalahari (Chap. 20). She can report, inter alia, from her observations
on how these experts reach dependable results when reading tracks which as an
exchange of personal insights is embedded in their sharing ideology. But Biesele
also reports on how challenging it was in the beginning to integrate western and San
analytical practice in the Tracking in Caves project.
A pioneering project of the integration of scientic and indigenous knowledge is
the topic of Webbs chapter who after the discovery of the Pleistocene footprints in
the Willandra Lakes in southern Australia was the rst to call indigenous experts to
help understand an archaeological source (Chap. 21). The success of this collabora-
tion is an inspiration to other projects because it showed that the interpretations of the
Aboriginal track experts were completely plausible and generated information that
we could not have obtained in any textbook and even a lifetime in archaeology.
With this Willandra project and the since 2013 ongoing Tracking in Caves
project, practical collaboration between scientic and indigenous experts has gone
some way together, but they had to design routines and practices without having
established models at hand. Therefore the concluding chapter by Zwischenberger
reviews the characteristics of western and indigenous expertise and how such
different knowledge traditions can collaborate on eye level (Chap. 22). Based on
this review and on the analysis of various ethical protocols of indigenous groups, she
compiles guidelines for the collaboration of scientic and indigenous experts. Thus a
circle closes back to the rst chapter in part I where Bennett and Reynolds provide a
checklist for the practical encounter with tracks as archaeological source which
Zwischenbergers contribution complements with an analogue list comprising points
that need to be observed when investigating such sources with indigenous support.
Reading prehistoric human tracks constitutes a perhaps unique kind of discourse
not only for archaeology but for our knowing of the world in general. With the most
sophisticated means of analysis and computing, we of today try to understand and
explain the very same sources, aiming at the very same results as we as a species will
have done millennia ago: we nd tracks of conspecics, and we want to know who
was here before me, what did she or he do and how did she or he feel. Admitting that
whatever apparatus we use, our results remain wanting, we of today fortunately can
call the support of indigenous experts who master reading tracks without the help of
any technology and yet arrive at much deeper understanding. They allow us to get a
glimpse of which information our prehistoric ancestors would have had access to
when encountering human tracks. With reading prehistoric human tracks, we can
liaise our epistemological procedures not only with experts from other cultures but
also with the knowledge even of our Pleistocene ancestors.
10 A. Pastoors and T. Lenssen-Erz
References
Bégouën, H. (1928). Les empreintes de pieds préhistoriques. Congrès International
dAnthropologie et dArchéologie Préhistorique III. Amsterdam, 323328.
Bennett, M. R., & Budka, M. (2019). Digital technology for forensic footware analysis and
vertebrate ichnology. Cham: Springer Nature.
Bennett, M. R., & Morse, S. A. (2014). Human footprints: Fossilised locomotion? Cham: Springer
International Publishing.
Bennett, M. R., Harris, J. W. K., Richmond, B. G., Braun, D. R., Mbua, E., Kiura, P., et al. (2009).
Early hominin foot morphology based on 1.5 Million year old footprints from Ileret, Kenya.
Science, 323, 11971201.
Bennett, M. R., Reynolds, S. C., Morse, S. A., & Budka, M. (2016). Laetolis lost tracks. 3D
generated mean shape and missing footprints. Scientic Reports, 6, 21916. https://doi.org/10.
1038/srep21916.
Biesele, M., & Barclay, S. (2001). Ju/hoan womens tracking knowledge and its contribution to
their husbandshunting success. African Study Monographs, (Suppl. 26), 6784.
Bustos, D., Jakeway, J., Urban, T. M., Holliday, V. T., Fenerty, B., et al. (2018). Footprints preserve
terminal Pleistocene hunt? Human-sloth interactions in North America. Science Advances, 4,
eaar7621.
Cartailhac, E., & Breuil, H. (1907). Une seconde campagne aux cavernes ornées de Niaux (Ariège)
et Gargas (Haute-Pyrénées). Académie des Inscriptions et Belles-Lettres, 213222.
Cartailhac, E., & Breuil, H. (1908). Les peintures et gravures murales des cavernes pyrénéennes.
LAnthropologie, 19,1546.
Crompton, R. H., Pataky, T. C., Savage, R., DAoût, K., Bennett, M. R., Day, M. H., et al. (2011).
Human-like external function of the foot, and fully upright gait, conrmed in the 3.66 million
year old Laetoli hominin footprints by topographic statistics, experimental footprint-formation
and computer simulation. Journal of the Royal Society, Interface, 9(69), 707719. https://doi.
org/10.1098/rsif.2011.0258.
Duday, H., & García, M. A. (1983). Les empreintes de lHomme préhistorique. La grotte du Pech-
Merle à Cabrerets (Lot). Une relecture signicative des traces de pieds humains. Bulletin de la
Société Préhistorique Française, 80, 208215.
Gagnol, L. (2013). Identier, rechercher et surveiller par les traces: Éléments danalyse dun savoir-
faire pastoral dans les sociétés saharo-sahéliennes. Technique & Culture, 61, 166187.
Leakey, M. D., & Harris, J. M. (Eds.). (1987). Laetoli: a pliocene site in Northern Tanzania.
Oxford: Clarendon.
Ledoux, L. (2019). Lichnologie préhistorique et les traces dactivités au sein des cavités ornées.
Les grottes de Fontanet (Ariège) et de Cussac (Dordogne). PhD thesis, Université de Bordeaux.
Liebenberg, L. W. (1990). The art of tracking: the origin of science. Claremont: David Philip.
Lockley, M. (1999). The eternal trail: A tracker looks at evolution. Reading: Perseus Books.
Lockley, M., Roberts, G., & Kim, J. Y. (2008). In the footprints of our ancestors: An overview of
the hominid track record. Ichnos, 15, 106125.
Lockley, M., Meldrum, J., & Kim, J. Y. (2016). Major events in human evolution. In M. G.
Mángano & L. A. Buatois (Eds.), The trace-fossil record of major evolutionary events
(pp. 411448). Dordrecht: Springer.
1 Introduction 11
Lowe, P. (2002). Hunters and trackers of the Australian desert. Surry Hills: Accessible Publishing
Systems.
Martin, R. (1928). Lehrbuch der Anthropologie in systematischer Darstellung mit besonderer
Berücksichtigung der anthropologischen Methoden für Studierende Ärtze und Forschungs-
reisende (2nd ed.). Jena: Fischer.
Matthews David, A. (2019). First impressions: Footprints as forensic evidence in crime in fact and
ction. Costume, 53(1), 4366.
Pales, L. (1976). Les empreintes de pieds humains dans les cavernes: Les empreintes du resau nord
de la caverne de Niaux (Ariège). Archives de lInstitut de Paleontologie Humaine, 36. Paris:
Masson.
Panofsky, E. (1962). Studien zur Ikonologie. Humanistische Themen in der Kunst der
Reanaissance. Köln: DuMont.
Pasda, C. (2013). Zu Fuss Spurenfossilien des eiszeitlichen Menschen. In A. Pastoors &
B. Auffermann (Eds.), Pleistocene Foragers: Their culture and environment: Festschrift in
honour of Gerd-Christian Weniger for his sixtieth birthday (pp. 5170). Mettmann: Neanderthal
Museum.
Pastoors, A., & Lenssen-Erz, T. (2020). Tracking in caves: Indigenous ichnology. In C. L. Smith
(Ed.), Encyclopedia of global archaeology. Cham: Springer. https://doi.org/10.1007/978-3-319-
51726-1_3430-1.
Pastoors, A., Lenssen-Erz, T., Ciqae, T., Kxunta, U., Thao, T., Bégouën, R., Biesele, M., & Clottes,
J. (2015). Tracking in caves: Experience based reading of Pleistocene human footprints in
French caves. Cambridge Archaeological Journal, 25(3), 551564.
Pastoors, A., Lenssen-Erz, T., Breuckmann, B., Ciqae, T., Kxunta, U., Rieke-Zapp, D., & Thao,
T. (2017). Experience based reading of Pleistocene human footprints in Pech-Merle. Quater-
nary International, 430, 155162. https://doi.org/10.1016/j.quaint.2016.02.056.
Stander, P., Ghau, I. I., Tsisaba, D., Oma, I. I., & Vi, I. I. (1997). Tracking and the interpretation of
spoor: a scientically sound method in ecology. Journal of Zoology, 242, 329341.
Topinard, P. (1877). LAnthropologie. Paris: Reinwald.
Vallois, H. V. (1928). Étude des empreintes des pieds humains du Tuc dAudoubert, de Cabrerets et
de Ganties. Congrès International dAnthropologie et dArchéologie Préhistorique, III,
328335.
Vallois, H. V. (1931). Les empreintes de pieds humains des grottes préhistoriques du Midi de la
France. Palaeobiology, 4,7998.
Webb, S., Cupper, M. L., & Robins, R. (2006). Pleistocene human footprints from the Willandra
Lakes, southeastern Australia. Journal of Human Evolution, 50(4), 405413.
Wong, P. B. Y., Van Coeverden, d. G. P., Fekken, C., Smith, H., Pages, M., & Boag, P. T. (2011).
Interpretation of polar bear (Ursus maritimus) tracks by Inuit Hunters: Inter-rater reliability and
inferences concerning accuracy. The Canadian Field-Naturalist, 125, 140153.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Part I
Methodological Diversity in the Analysis of
Human Tracks
Chapter 2
Inferences from Footprints: Archaeological
Best Practice
Matthew R. Bennett and Sally C. Reynolds
Abstract Animal footprints are preserved in the archaeological record with greater
frequency than perhaps previously assumed. This assertion is supported by a rapid
increase in the number of discoveries in recent years. The analysis of such trace
fossils is now being undertaken with an increasing sophistication, and a methodo-
logical revolution is afoot linked to the routine deployment of 3D digital capture.
Much of this development has in recent years been driven by palaeontologists, yet
archaeologists are just as likely to encounter footprints in excavations. It is therefore
timely to review some of the key methodological developments and to focus
attention on the inferences that can and, crucially, cannot be justiably made from
fossil footprints with specic reference to human tracks.
Keywords African palaeoecology · Mammalian ichnology · East African fossil
record · Laetoli · Hominin sites · Pliocene · Ichnology
Introduction
Every contact an animal makes with the ground has the potential to leave a trace, as
set out in Locards famous exchange principle. The average moderately active
person, for example, takes around 7500 steps a day, and if maintained over a lifetime
of 80 years, then they will have left the order of 216,262,500 steps with each step
having a theoretical potential for preservation. Contrast this with the 206 bones in the
human body, and it is not surprising that we frequently uncover fossil footprints. In
fact, it is surprising that we dontnd more. Something we would argue is due to the
lack awareness and prospection, rather than any particular rarity in the geological or
archaeological record. There is a well-documented recent examples where footprints
of Homo heidelbergensis have been recovered (Altamura et al. 2018) but were not
M. R. Bennett (*) · S. C. Reynolds
Institute for Studies in Landscape and Human Evolution, Faculty of Science and Technology,
Bournemouth University, Poole, UK
e-mail: mbennett@bmth.ac.uk
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_2
15
recognised by earlier excavators who destroyed tracks in their quest for more
conventional archaeological materials. Increasing awareness of archaeologists of
the potential to nd footprints in excavations of all ages is an important endeavour,
as is drawing attention to the convergence of methodological approaches, standards
of recovery and best practice. Much of this best practice has been driven in recent
years by the palaeontological community, and it is therefore timely, we believe, to
reviewthese developmentsfor the benet of all, especially for archaeologists.
16 M. R. Bennett and S. C. Reynolds
There has been a rapid growth in recent years in the discovery of human fossil
footprints around the world which questions the long-held and much-stated assump-
tion that footprint preservation is a freak geological event (Bennett and Morse 2014).
To give a avour of the recent discoveries, in 2016 we saw the publication (Masao
et al. 2016) of additional footprints at the famous 3.66-million-year-old footprint site
at Laetoli in northern Tanzania rst reported in 1979 by Leakey and Hay (1979). Not
far from Laetoli, a Late Pleistocene site on the shores of Lake Natron was reported
with hundreds of visible tracks (Balashova et al. 2016; Liutkus-Pierce et al. 2016;
Zimmer et al. 2018). In 2018 the publication of childrens footprints in association
with butchered hippo carcasses was reported from Ethiopia (Altamura et al. 2018),
and there are reports of human tracks in association with giant ground sloth in North
America (Bustos et al. 2018). Footprints preserved in peat have been found on the
Pacic Coast of Canada (McLaren et al. 2018), and a new footprint site in
South Africa is reported by Helm et al. (2018). Footprints have been found in a
diverse range of environments (Fig. 2.1a), and improved awareness by excavators,
continued prospection and a revolution in digital techniques for their capture and
analysis are perhaps responsible for the increasing discovery of new sites. There is
much more to do however.
The aim of this contribution is explore modern tools for the capture and analysis
of fossil footprints and to emphasise some of the challenges archaeologists face in
making inferences from human footprints in particular. We have structured this
review along three stages in what we see as the ichnological pipeline: (1) digital
capture and documentation, (2) analysis and (3) inference. To aid cross-disciplinary
convergence, the key terminology used in vertebrate ichnology is provided in
Table 2.1.
Digital Capture, Documentation and Stratigraphic Context
A quiet revolution during the last decade has transformed human ichnology from an
essentially descriptive discipline into one that is now both data- and hypothesis-
driven. This transformation started with the introduction of optical laser scanners to
capture 3D tracks and has been completed with the routine availability of Structure
from Motion-based photogrammetry (Bennett and Budka 2018). While 3D docu-
mentation is not necessarily new, having been applied in the late 1970s to the Laetoli
tracks (Day and Wickens 1980; Leakey and Harris 1987), it has now become routine
for all modern practitioners, although there remain situations where it has not been
2 Inferences from Footprints: Archaeological Best Practice 17
Fig. 2.1 (a) Sketch showing the typical types of location in which fossil footprints have so far been
found after Bennett and Morse (2014
successfully applied or has failed due to environmental conditions (e.g. Ashton et al.
2014). Collecting 3D data allows the user to test assumptions and inferences that
previously were made simply by assertion, as exemplied by the work of Roberts
et al. (1996). There are those that argue that assertions made by expert trackers
(Pastoors et al. 2015,2016) provide a valid alternative, and while the skill of the
); (b)the power of independent lines of investigation leading
convergence and/or corroboration
18 M. R. Bennett and S. C. Reynolds
Table 2.1 Commonly used terms with respect to footprint or footwear impression. After Marty
et al. (2009)
Term Denition
Track A single footprint or partial impression made by the foot of shod or unshod
animal or human
Trackway A series of tracks made by a single individual
Trail A series of signs or objects left behind by the passage of someone or something.
In this context it might be multiple tracks left by one or more individuals,
forming a path, for example
Trackmaker The animal that made the track
Tracked
surface
The surface on which the trackmaker walked/moved
Over-printing Caused by an individual or animal over-printing an original track
Displacement
rim
A marginal rim to a track formed by the upward displacement of sediment,
sometimes referred to as a push-up structure or a bourrelet
Track ejecta Material ejected by the removal of the trackmakers foot from a track. This often
forms a debris trail in front of a track
Plantar
surface
The base of the trackmakers foot or shoe
Ichnosurface A surface with multiple tracks which may have either formed in an isochronous
(synchronous) or diachronous (time-averaged) fashion
trackers is not in dispute, the inability of a reader to question and/or test those
assertions themselves limits the scientic credibility of such approaches.
The authors have developed, along with colleagues, freeware (www.digtrace.
co.uk) that enables the capture of tracks in 3D using the open source software
OpenMVG (Bennett and Budka 2018). Not only can you create 3D models, but
the freeware provides a series of analytical workbenches with tools specic to the
analysis of footprints. Commercial alternatives exist in the form of such things as
PhotoScan by Agisoft (http://www.agisoft.com), and while they create excellent 3D
models, they do not provide analytical tools specic to footprint analyses, and the
user has to rely on expensive 3D modelling software or freeware tools such as
MeshLab (www.meshlab.net) or CloudCompare (www.danielgm.net/cc). Structure
from Motion relies on multiple oblique digital photographs from which individual
pixel clusters are placed in 3D space. Almost any digital camera can be used
provided its sensor size is known or can be calculated (Bennett and Budka 2018).
The archaeologist who arrives at a eld site or encounters a series of footprints for
the rst time needs a plan. Figure 2.2 lists some of the key elements of any plan and
reviews the things that need to be considered. Assuming that one has the necessary
permits and permissions, the rst step is to consider whether the tracks can be
preserved or whether it is a case of conservation by rescue. Preserving soft-sediment
footprints is challenging, if not impossible (Bennett et al. 2013), and 3D digital
capture is often the only way to create a permanent record of the tracks (Bennett and
Morse 2014). This is especially true of those tracks preserved in coastal exposures
(Bennett et al. 2010; Falkingham et al. 2018; Wiseman and De Groote 2018)or
where the act of excavation disturbs a fragile surface, thereby allowing erosion.
2 Inferences from Footprints: Archaeological Best Practice 19
Fig. 2.2 The basic structure of an ichnological eld survey plan, modied from Bennett and Budka
(2018)
Interestingly, recent work using geophysics (magnetometry and ground-penetrating
radar) shows promise with respect to how tracks may be prospected for without
surface disturbance or excavation (Urban et al. 2018).
20 M. R. Bennett and S. C. Reynolds
The basic information that is required is (1) spatial information showing the
relative position of one track to another (i.e. some form of map); (2) detailed 3D
models of all or a selection of individual tracks; (3) detailed vertical photographs,
written observations and measurements; (4) facies descriptions of the containing
sediments both vertically and spatially; and (5) detailed sampling for datable mate-
rials both below and above the tracked layer. Spatial mapping can be achieved in a
variety of different ways, although traditional eld survey techniques, coupled with
low-level aerial photographs, are now the norm. Traditionally plastic sheets have
been placed over tracked surfaces, and footprint outlines have been traced onto them.
These have to then be reduced to make them manageable, the advantage of vertical
images is that, if the surface is not horizontal, contours can be added using photo-
grammetry or portable LiDAR devices. Working at White Sands National Park
(WHSA), Bustos et al. (2018) used Agisofts PhotoScan to produce photomosaics
of large areas, supplemented with detailed 3D models of individual tracks made with
DigTrace.
Tracks need to be documented individually, or in combination. Deciding what the
sampling strategy should be is critical here. For example, how many tracks should be
excavated out of the total population? What proportion should be conserved, if any,
for future investigators? This clearly depends on the total available track count, but at
sites like those described by Morse et al. (2013) and Bustos et al. (2018) where there
is a surplus of tracks and excavation limits long-term preservation, then these are real
questions for which there are few denitive answers. As working principle excavat-
ing/disturbing the minimum number of tracks is usually the general practice, unless
the site is at risk. Sampling strategies are not necessarily relevant where there are a
small number of endangered tracks. Anything and everything that is excavated needs
to be captured in 3D. Where a whole surface has been scanned, or a 3D mosaic
created, there is a temptation to simply crop this down to create individual track
models. Large-scale models do not always have the resolution, however, for detailed
topological study and rarely deal with undercut edges well. It is therefore advisable
to also make close-up models where lines of sight can be improved. It is also worth
noting that photogrammetry, or laser scanning for that matter, is not always a perfect
solution especially where the tracks are deeply incised in soft ground. The use of an
endoscope can help with this, but it may be necessary to adopt alternative strategies.
For example, Altamura et al. (2017) excavated deep hippo tracks and then inlled
these with plaster before excavating the surrounding surface to reveal the cast. The
casts can be laser scanned subsequently, if required. Once the individual tracks have
been documented, it is then necessary to describe and sample the tracked units and
surrounding lithofacies, using standard procedures.
2 Inferences from Footprints: Archaeological Best Practice 21
Analytical Tools in Ichnology
Traditionally good science requires a separation between description and interpre-
tation. The description will always stand if done well, but the interpretation may
change with time and new ideas or discoveries. In the context of a track, this is the
separation between describing a tracks topological properties and making infer-
ences about the trackmaker from them. There are lots of examples where a
trackmaker is inferred without such description (e.g. Musiba et al. 2008). In describ-
ing and/or measuring a track of whatever origin, we have three main independent
properties to consider. Firstly, we have size dened by a single, or more usually, a
combination of linear measurements. Typically these include measures of length,
area and volume. Secondly, we have shape or form which describes how the track is
dened by intersecting lines, edges or textural boundaries. For example, do the edges
dene a triangle or square? Finally we have topology which is the geometrical
properties, and their spatial relationships to one another, such as the spatial dispo-
sition of different component shapes and depth variations. The morphology
(or anatomy) of a track is the sum of the above properties, and all three dimensions
need to be considered when describing a track (Fig. 2.3a).
Separating aspects of shape and size is an important part of any anatomical
description and is usually achieved by the superposition and transformation of
geometric forms such that size is removed. This is a standard part of most geometric
morphometric analysis and is commonly attained via some form of Generalised
Procrustes Analysis (GPA; Zelditch et al. 2012; Gómez-Robles et al. 2008; Friess
2010). Berge et al. (2006) pioneered the application of GPA to human tracks, an
approach adopted and rened by Bennett et al. (2009) in their analysis of the Ileret
footprints and used by others since (e.g. Wiseman and De Groote 2018). All the
above require the placement of some form of homology-based landmark, that is, a
landmark that relates to a biologically or anatomically homologous structure and
crucially one that can be recognised consistently by observers. Even a single linear
measurement of length requires landmarks to be placed at the start and nish of the
measurement line. Dening landmarks consistently across different studies and
operators is a source of error especially where linear properties such as length are
not clearly dened (Bennett and Morse 2014).
Concern over landmark placement lead Crompton and his team at the University
of Liverpool to develop a whole-foot comparison method in which tracks are
co-registered allowing measures of central tendency to be determined for entire
track populations. Taking their lead from the mathematics behind the analysis of
magnetic resonance imaging (MRI), they developed pedobarographic statistical
parametric mapping (pSPM). It was designed to co-register multiple pressure records
obtained from individuals walking on a treadmill, once co-registered pixels in
similar anatomical positions are compared (Crompton et al. 2012). By substituting
depth for pressure, one can apply it to tracks, although it requires the removal of all
marginal structures for automated registration since it can only match recurrent
plantar surfaces. Manual registration gets around this problem, but in doing so the
22 M. R. Bennett and S. C. Reynolds
Fig. 2.3 (a) Components of track morphology; (b) analytical questions that can be addressed by
quantitative analysis when 3D digital data is collected
objectivity obtained by auto-registration is lost. It is important to note that marginal
deformation structures are as equally valid as the plantar surfaces in interpreting
tracks.
2 Inferences from Footprints: Archaeological Best Practice 23
Bennett et al. (2016a) use an alternative approach in which tracks are
co-registered by matching common structures, via placed landmarks. These matched
points are then used to guide the registration, and this approach has the advantage of
being driven by a simple user interface. Whichever approach is used to register a
population of tracks, once co-registered, one can create measure of central tendency,
pixel by pixel, in the form of a mean/median track and measured of statistical
variability around this (Crompton et al. 2012). This allows both dimensions of
track variability to be explored, namely, (1) intra-trackway variance, that is, the
variation between different tracks made by the same trackmaker in a trackway due to
variation in substrate and inter-step biomechanics, and (2) inter-trackway variance,
that is, the difference between two different trackways that may, or may not, have
been made by the same individual. Both are critical to determining whether a set of
tracks were made by the same trackmaker or not. If the intra-trackway variance is
greater than the inter-trackway variance, then you have a problem in making a
denitive qualitative or quantitative distinction between them. Belvedere et al.
(2018) introduce two new terms, the stat-track and the mediotype (Fig. 2.3b). The
former refers to any statistically produced measure of central tendency for a popu-
lation of tracks, while the latter refers specically to a mean or median track created
from those holotype specimens. They also provide a range of examples of how
whole-track methods can be used to help formalise the description of formal
ichnotaxa (Marty et al. 2009,2016).
Types of Inference from Human Footprints
The discovery and/or excavation of a fossil trackway can be an exciting process,
revealing as it does a captured moment in time when the foot of a human made
contact with the ground. Tracks lead to a range of analytical questions (Fig. 2.3b),
and there are four broad areas of inference that can be drawn from such discoveries:
(1) the trackmaker, their pedal anatomy and inferences about size and body mass;
(2) locomotion style and speed from the depth distribution which is assumed to be a
proxy for pressure; and (3) the palaeobiology of the track assemblage. Different
preservation conditions favour different types of inferences as illustrated in Fig. 2.5
and discussed below.
Anatomical Inferences
An individual track, or more reliably a population of morphologically similar tracks
(Fig. 2.3b), preserves anatomical information about the trackmaker, such as the
shape of the foot and/or the number of digits. This information is inclusive of the soft
tissue that surrounds the bones, material that is rarely preserved and yet essential to
anatomical description and locomotory behaviour. A given animal may produce a
range of different track topologies depending on functional interaction with the
sediment and the sediment itself (Bennett et al. 2014). The next logical step is to
name the trackmaker, although this is not always as simple as it sounds. Within the
Plio-Pleistocene at least the starting point for such interpretation is the
palaeontological record or a modern animal track guide. Over-reliance on the
palaeontological record provides an ever-present risk of missing a species known
only by its tracks at a given site. In a similar way, reliance on modern tracking guides
or local/native trackers when dealing with Pleistocene track sites (Pastoors et al.
2015,2016) assumes a similarity that is not always warranted between past and
present communities. The tendency to t tracks to a known template is also an ever-
present risk. Good science comes from good building blocks forming safe founda-
tions, and in the case of ichnology, this consists of good topological, quantitative, 3D
descriptions of individual tracks and their topological variability as a population of
tracks. This should occur independently of any assessment of the sedimentary facies
and associated palaeontology (Fig. 2.1b).
24 M. R. Bennett and S. C. Reynolds
Assuming the trackmaker can be deduced, the next stage is to consider what
biometric inferences are possible for that given animal. Empirical relationships
between foot size and height are well known for humans and some animals
(Fig. 2.4). For example, Roberts et al. (2008) describe an elephant trackway dated
to around 90 ka from Still Bay in South Africa in aeolinites. The tracks are attributed
to the African elephant (Loxodonta africana), and, interestingly, Roberts et al.
(2008) use the track dimensions to infer both shoulder heights and ages for the
trackmakers using the empirical relationships of Western et al. (1983). There is good
data about the ontology of African elephants (Western et al. 1983; Lee and Moss
1995; Shrader et al. 2006) allowing biometric data to be inferred. It is an approach
that has been applied elsewhere to mammoth tracks in Canada (McNeil et al. 2005)
and Miocene Proboscidea tracks in the United Arab Emirates (Bibi et al. 2012). The
idea is based on tting a growth curve such as a von Bertalanffy growth function, to
available empirical data, despite the fact that the trackmakers sex cannot normally
be determined from a track alone and that the presence of sexual dimorphism
complicates such inferences. To what extent this approach could be applied to
other animals is uncertain but has the potential to be an interesting line for future
research. It has been applied to dinosaur tracks where there are clear morphological
variations with size (e.g. Avanzini and Lockley 2002), and some data on the
ontology of ungulate tracks does exist (e.g. Miller et al. 1986; Musiba et al. 1997;
Cumming and Cumming 2003; Stachurska et al. 2011; Parés-Casanova and
Oosterlinck 2012a,b).
In terms of humans, there is a wealth of empirical relationships for different
ethnic/racial groups which relate foot size to stature (and refs therein: Bennett and
Morse 2014). The founding data is based on different measurement protocols and
either 2D foot imprints or direct foot measurements. As a consequence the value of
this huge body of data to fossil footprint studies is more limited than sometimes
Fig. 2.4 Stature and body mass inferences from footprints; (a) data from a beach showing variation
in foot length due to intra-step variability. A total of 57 tracks where used and then boot-strapped to
give the 2000 data points shown. The 95% condence ellipse is shown by the dashed line; (b)
stature estimates using a range of different published stature equations for a single Holocene
trackway in Namibia N ¼77. A fuller description of this analysis is provided in Bennett and
Morse (2014); (c) foot length, stature and body mass estimates for a sample (N ¼234) of 3D
footprints for a Bournemouth University staff/students
assumed. The rst issue is that foot length varies considerably along a trackway due
to substrate, biomechanical inter-step variability and chance. Consider the data in
Fig. 2.4a which shows the length variation of a single individual walking on a beach
and the potential height variations obtained using a simple 15% foot length to stature
ratio. Similarly for a fossil trackway, you dont normally know the ethnicity/race,
sex or age of the trackmaker, and you are left to make a series of assumptions in
selecting an empirical relationship to model stature from foot length, and this is even
more problematic when dealing with extinct human ancestors. Figure 2.4b shows a
series of height estimations made from the same trackway of 77 tracks in Namibia
(Morse et al. 2013) using different empirical equations. They give a range of
estimates which are generally higher than the more conservative 15% rule of
Topinard (1877). While they give a greater sense of scientic rigour, this is perhaps
an illusionary, and the results depend on the empirical relationship chosen (Fig. 2.4).
Empirical relationships to body mass are available (Fig. 2.4), but they are much
weaker than those for stature, and, while some more sophisticated modelling has
been attempted (e.g. Dingwall et al. 2013; Masao et al. 2016), again the reliability of
such estimates has to be questioned (Bennett and Morse 2014). Finally, while
modest levels of sexual dimorphism are evident in human track length, it is possible
to separate genders on the basis of track measurements within controlled populations
(Bennett and Morse 2014). However, there are simply too many unknowns to do so
in the fossil record. Where small adult tracks exist, it is perhaps possible to
tentatively suggest that they may be those of women, but the possibility that they
are adolescent males can never be ruled out denitively. Despite this, implicit and
explicit gender assignment of footprints is common within the literature. The other
issue to be considered is how representative are size estimates of a population as a
whole. Sampling by substrate and sampling by activity are all issues that are relevant
here (Kinahan 2013). The harsh reality is that unless a large number of tracks are
present, clearly made by different individuals, it is hard to make reliable inferences.
True demographic reconstructions from the basis of human footprints have yet to
attempted, not least because the number of sites where it can be applied is limited. In
terms of track abundance, there is also a ne line between too many tracks leading to
multiple partial impressions due to over-printing and too few to do anything other
than to give data on an individual (Fig. 2.5).
26 M. R. Bennett and S. C. Reynolds
Biomechanical Inferences
As the foot makes contact with the ground, if the shear strength of the substrate is
exceeded, it will deform leaving a record of that interaction (Hatala et al. 2018).
Implicit in all biomechanical inferences from tracks is the idea that distributions in
depth across the surface of a track provide a measure of the plantar force applied by
the foot in different locations and phase of stance. The experimental work of Bates
et al. (2013) shows that there is greatest correlation between plantar pressures and
depth for shallow tracks and that this relationship holds less for deeper examples.
2 Inferences from Footprints: Archaeological Best Practice 27
Fig. 2.5 Factors that control the nature of an ichnosurface and its scientic value
Given that the preservation potential is greater for deeper tracks, this may limit
biomechanical inferences from some types of fossil track site (Fig. 2.5). Speed
estimates are possible using step and stride length records and the empirical relation-
ships developed by Alexander (1984). Biomechanical inferences on hominin tracks
have been extensive and ercely debated (e.g. Meldrum et al. 2011; Bennett et al.
2016a,b; Hatala et al. 2016b)and, as such, are not considered further here.
Palaeobiological Inferences
Lockley (1986) argues that tracks provide palaeobiological data, ranging from the
taxonomic identication of the trackmaker, placing them at given location and time,
to providing information on faunal biodiversity where evidence of multiple
trackmakers is present. This can be used to contribute data on the palaeogeographic
range of a species, its demography and/or population dynamics. Interestingly, in
contemporary environments track density maps are used to assess population num-
bers of specic species, most notably carnivores, and perform better than some other
census methods (e.g. Prins and Reitsma 1989; Silveira et al. 2003; Gompper et al.
2006; Funston et al. 2010; Moreira et al. 2018). The problem with fossil tracks is one
cannot constrain the time interval over which tracks are preserved, and faunal
sampling usually inuences the results. This subject is explored further below.
28 M. R. Bennett and S. C. Reynolds
Inferences from track assemblage with respect to the behavioural ecology of
trackmakers, such as the composition of herds, tendency toward gregarious habits
and even prey-predator relationships, are, in theory, possible. Martin and Pyenson
(2005) argue that trackways contain evidence of an array of behaviours, such as
shifts in speed and direction, lateral movements, obstacle avoidance as well as
gregarious movements. Ostrom (1972) and several others have used directional
data of dinosaur trackways, for example, to argue for gregarious behaviour, while
Bibi et al. (2012) argued for evidence of social structure in Miocene Proboscidea in
the United Arab Emirates on the basis of trackway patterns. Working at Ileret
(Kenya) track site, Roach et al. (2016,2018) argued that hominin tracks show
similar states of deterioration and typically do not cross-cut one another and as
such can be considered at best to be contemporaneous and at worst
penecontemporaneous. They suggest that the parallel hominin trackways (1.5 Ma),
with individual tracks of similar size, indicate that the trackmakers moved as part of
male hunting groups (see also Hatala et al. 2016a,2017). Altamura et al. (2018) point
to the co-association of child tracks with stone tools and the remains of a butchered
hippo carcase from which they infer a more pastoral scene in which young children
are present (0.7 Ma) and presumably learning. Perhaps one of the best examples,
however, of behavioural inference is provided by Bustos et al. (2018). In this work
they show that the tortuosity of trackways made by extinct giant ground sloth from
the terminal Pleistocene increases in the presence of human trackmakers. In fact the
trackways show evidence of evasion with sudden changes in direction and speed.
Bustos et al. (2018) suggest that human hunters were stalking and harassing sloth,
presumably as part of a hunting strategy. The higher the density of tracks and the
more extensive they are in space, the greater the potential to infer palaeobiological
information (Fig. 2.5).
Critical to all these inferences are two fundamental questions which are
referenced in passing in most track-based publications but rarely developed in detail.
They are issues of faunal sampling and of demonstrating co-association within a
track assemblage. Both issues are underpinned by the geological context of the site
and are explored further here due to the fundamental importance of these issues.
Faunal Sampling
Essentially this distils down to the question that troubles all palaeontological studies
eventually, which is to what extent does a proxy count, for example, of the number
of bones or tracks mirror the original faunal population from which they are derived,
in terms of composition and abundance, and to what degree do taphonomic pro-
cesses distort this reection? In the context of tracks, these issues were explored in
the seminal papers by Cohen et al. (1991,1993) which provide one of the few
2 Inferences from Footprints: Archaeological Best Practice 29
Table 2.2 Factors inuencing track survivorship after Cohen et al. (1993)
Substrate susceptibility (strain) Track loading (stress) Secondary reworking rates
Sediment properties (texture,
sorting, bulk density, organic
content)
Pedal anatomy and animal
body mass (mass/shape/
force)
Vertebrate trampling/foraging
(+)
Water content (+/ ) Biomechanics (including
manus vs pes variations)
Invertebrate bioturbation (+)
Degree of cementation inclusive
of salts and algal mats ( )
Kinematics (e.g. direction,
acceleration and
deceleration)
Surface disturbance such as
waves, run-off, deation, desic-
cation (+)
Algal stabilisation/cementation
()
Salt blooms ( /+)
Surface desiccation ( /+)
Burial ( )
The positive and negative signs refer to the direction of a potential inuence. These three grouped
variables are plotted on ternary diagrams in Fig. 2.6
modern analogue studies (neoichnology) relevant to the interpretation of Plio-
Pleistocene track sites.
Cohen et al. (1993) recognise three basic variables at play in track preservation
(summarised in Table 2.2), namely, (1) animal load (force/area) (2) sediment sus-
ceptibility (strain/stress) and (3) secondary reworking. These three summary vari-
ables are not completely independent of one another. For example, a surface with a
high susceptibility to deformation will also be one that is easily reworked. However,
it does provide a means of exploring the interplay of these variables in track
formation and the immediate taphonomy of those tracks. Here we conceptualised
these variables on a ternary diagram to dene a track-forming window for a given
animal (load), at a specic time and place (Fig. 2.6). It denes the Goldilocks track-
forming zone, as modelled numerically by Falkingham et al. (2011). These variables
and the track-forming windows they dene vary spatially across a site such as a lake
margin, or river, and temporally, as pore water and surface water conditions change
over time. The track-forming window, present at one location and moment in time,
samples the fauna that passes; too little fauna and it wont be sampled, too much and
it will not be preserved at all due to self-inicted bioturbation (Fig. 2.5). Take the
example shown in Fig. 2.7, which shows a hypothetical model based on a lake or
river margin. Rainfall events are linked to rises in water level, and track preservation
occurs primarily during falls in water levels when the maximum printable surface is
revealed. Animals are not continuously present but visit periodically, as indicated by
the green bars. For tracks to be preserved, there need animals to be coincident with a
track-forming window. The site is effectively trackmaker limited, and the tracks
preserved do not reect the whole faunal community, simply part of it (Fig. 2.7).
The corollary to the track-forming window is the surface relaxation time. That is,
the time it takes for the surface structure (i.e. track topology) to decay through
surface reworking, whether due to bioturbation or by surface processes, along with
30 M. R. Bennett and S. C. Reynolds
Fig. 2.6 Track-forming window as dened by the interplay of animal load, substrate reworking
and substrate susceptibility. Different animals have different track-forming windows. For example,
compare (b)to(c) while track-forming windows may vary through time, as shown in (c)
sediment plasticity during enhanced moisture or brecciation caused by desiccation.
Brecciation of desiccated tracks was documented in modern analogue studies from
Amboseli (Kenya) by Bennett and Morse (2014) and is discussed by Cohen et al.
(1993). There are factors which can delay relaxation, such as the growth of algal
mats, associated sediment trapping (Marty et al. 2009) and cementation by salts. The
distribution of track-forming windows will vary spatially around a given environ-
ment. On the margins of the saline Lake Manyara (alkaline ats) in Tanzania, Cohen
et al. (1991,1993) recognised three ichnological zones:
Zone One: This was an onshore zone where the sediment is dry at the surface and
subject to salt blooms and associated deation. Insect bioturbation is limited, and
biotic reworking is achieved by animal trampling. While groundwater uctua-
tions do occur, the surface is sufciently rm to only take the tracks of the
2 Inferences from Footprints: Archaeological Best Practice 31
Fig. 2.7 Hypothetical model of a lake or river margin. For tracks to be preserved as shown on the
right, animals need to be coincident with a track-forming window. The isolated tracks shown are
associated with occasional preservation during water level rises, while the broader ichnosurfaces are
revealed during water-level regression
heaviest animals when groundwater rises. Desiccation often renders tracks indis-
tinct, and they are inlled by breccia and wind-blown mud. Long-term survivor-
ship of tracks can be high.
Zone Two: This is the shore zone where sediments are usually wet and salt crusts
are minimal. Insect bioturbation is pronounced, as is animal reworking, ground-
water uctuations and shore face reworking. The soft sediment in this zone is
ideal for the preservation of small mammals and birds in conjunction with larger
ones, but survivorship, and therefore preservation potential, is far less.
Zone Three: This is the subaqueous zone where the sediment is saturated and
often, as a consequence, unstable. Larger animals may leave abundant tracks in
this zone, but they quickly become indistinct and have low survivorship potential.
The way in which animals traverse these zones, and thereby leave their tracks,
depends in part on water quality. In the case of Lake Manyara, most of the trackways
follow the shoreline due to its salinity. No animals drink here. At less saline lakes,
animal trackways may be more perpendicular to the shore, and carnivores (including
humans) may adopt a more shore-parallel strategy in order to intersect prey (Roach
et al. 2016). The point here is that each zone has different track-capturing properties
and will sample the fauna differently, and different zones also have different
preservation potentials. Clearly the deeper tracks in Zone One may have greater
preservation potential than those in Zone Three. Small mammals and birds, for
example, may be under-represented in some faunal samples.
Animal behaviour in each of these zones may also be relevant; one animal
repeatedly walking in a small area can generate a lot of tracks! Cohen et al. (1993)
make a distinction between faunal estimates based on milling and directional
behaviour, tracks left by the former greatly inate animal abundances, and they
suggest that randomly orientated tracks should be avoided in making abundance
estimates. Similarly, game trails may be recognisable at some sites, but because of
their composite nature, they speak only to the presence of multiple animals, not to
the exact number, even if individual ichnotaxa can be discerned (e.g. Ashley and
Liutkus 2003). In terms of broad time-averaged faunal assessment, the risk of
sampling across diachronous surfaces is unlikely to be a signicant issue, unless
we are dealing with large time intervals; after all, skeletal and tool assemblages are
time-averaged phenomena. However, demonstrating co-association is an issue for
inferences around behavioural ecology.
32 M. R. Bennett and S. C. Reynolds
Problems of Co-association
That a track assemblage was imprinted in close temporal association is an assump-
tion that is implicit at most footprints sites but one which should, in truth, be
demonstrated every time. We can explore some of the issues by theoretically
modelling lake-level uctuations around a lake such as Lake Manyara. Figure 2.8
shows a water-level curve, deduced from a stacked sequence of lithofacies, with
tracked surfaces marking lake-level regressions. Lake regressions create the maxi-
mum spatial track-forming window; transgressions will tend to erode tracks, as shore
process advances over an area and may compress the space for track-forming zones.
This may correspond, however, to different rates of regression, when the data is
re-plotted with time as the vertical (Fig. 2.8). Only one of the ichnosurfaces shown is
isochronous, being associated with a rapid fall in lake levels, while the others record
diachronous assemblages. On the margins of this hypothetical water body, the rates
of sedimentation and therefore track burial will likely vary seasonally, and multiple
track-forming windows may become superimposed (Fig. 2.9). We can develop the
model rst presented in Fig. 2.7 to explore this further, although in this case animal
abundance is continuous (i.e. track formation is window limited). In the rst
Fig. 2.8 Hypothetical model of lake-level uctuations around a lake or river margin with available
trackmakers. The lithofacies associated with such a scenario is indicated on the left. These are likely
to be ning-upward sequences, associated with the transgressions, with tracks on the unconfor-
mities associated with regression. When time is substituted for depth on the right, the importance of
a rapid fall in water level in creating an isochronous faunal track sample is indicated
2 Inferences from Footprints: Archaeological Best Practice 33
Fig. 2.9 Three hypothetical scenarios around the margin of a lake or river, with a link between
rainfall and water level. Animals are present throughout the time shown, and we assume that the
optimum track-forming window occurs during a lake-level fall and that with exposure, the tracks
degrade through desiccation and brecciation; (a) in this scenario only the deepest tracks (Zone One)
will be preserved; (b) in this scenario the three track-forming windows are essentially superimposed
on one another to form a diachronous assemblage; (c)nally, in this scenario, the rising lake level is
assumed to be associated with sedimentation, and the tracks become more separated. With greater
rates of sedimentation, separation between the track-forming surfaces will increase. The presence of
isolated tracks represents the possibility of deeper tracks being preserved during a transgressive
episode. Note the use of window limited and trackmaker limited in Fig. 2.7
scenario, periodic water-level highs are separated by desiccation events, leading to a
tracked assemblage dominated by desiccation, in which only the deepest tracks,
associated with the heaviest animals, are preserved, and then only poorly (Fig. 2.9a).
In scenario two, we have shorter periods of desiccation between track-forming
windows, but little sedimentation and the resulting tracks are superimposed one
into the other, such that the surface is really a composite of several track-forming
windows. This is typical, for example, of some of the assemblages at Ileret. Only in
the third scenario, where rising water levels and associated sedimentation are
assumed, do we get tracked surfaces that separate out one from another and verge
toward being isochronous (Fig. 2.9c). Only on these types of surfaces is there
potential to deduce behavioural interactions.
34 M. R. Bennett and S. C. Reynolds
Co-association is usually argued for on the basis of cross-cutting relationships
between different trackways and on similar levels of track freshness. That is, tracks
which show a similar degree of degradation are assumed to be
penecontemporaneous, especially if they are associated similar deformation struc-
tures (i.e. rim and sub-track structures) indicating similar pore water conditions. The
ever-present risk is that a track-forming surface may have been reactivated several
times. These models demonstrate that, in truth, there is no simple answer to the issue
of co-association, other than care is needed when making such assumptions. These
assumptions need to be justied and explained more clearly than is often the case in
the literature. Ultimately the only way of demonstrating true co-association is to look
for evidence of behavioural interaction between one or more trackmakers
(e.g. Bustos et al. 2018). If two animals where present on the landscape at the
same point and moment in time, there should be some interaction, either active
(i.e. prey-predator) or passive (i.e. scenting or avoiding via trackway adjustment).
Conclusions
The study of human ichnology, along with other associated animals, is rapidly
advancing with the discovery of new sites. The advent of 3D digital capture allows
increasing research sophistication with quantitative hypothesis-led testing. There is
much to do to replace the assertion-based approaches of the past with data-driven
observations and inferences. In this review, we have explored some of the issues
associated with the ichnological pipeline from data collection to the inferences that
can be made, advocating throughout for the power of 3D digital data capture and
analyses. In the later part of the review, we have emphasised the importance of
geological context to the interpretation of track-based assemblages, and to assump-
tions of co-association, and around issues of faunal sampling. There is greater rigour
needed here in ichnological research, and we would argue that the geological context
of an assemblage is critical to its interpretation, regardless of the level of sophisti-
cation of the tools used to document it.
References
Alexander, R. M. C. N. (1984). Stride length and speed for adults, children, and fossil hominids.
American Journal of Physical Anthropology, 63,2327.
Altamura, F., Melis, R. T., & Mussi, M. (2017). A middle Pleistocene hippo tracksite at Gombore
II-2 (Melka Kunture, upper Awash, Ethiopia). Palaeogeography, Palaeoclimatology,
Palaeoecology, 470, 122131.
Altamura, F., Bennett, M. R., DAoût, K., Gaudzinski-Windheuser, S., Melis, R. T., Reynolds,
S. C., & Mussi, M. (2018). Archaeology and ichnology at Gombore II-2, Melka Kunture,
Ethiopia: Everyday life of a mixed-age hominin group 700,000 years ago. Science Reports, 8,
2815. (2018). https://doi.org/10.1038/s41598-018-21158-7.
2 Inferences from Footprints: Archaeological Best Practice 35
Ashley, G., & Liutkus, C. M. (2003). Tracks, trails and trampling by large vertebrates in a rift valley
paleo-wetland, lowermost Bed II, Olduvai Gorge, Tanzania. Ichnos, 9,2332.
Ashton, N., Lewis, S. G., De Groote, I., Duffy, S. M., Bates, M., Bates, R., Hoare, P., Lewis, M.,
Partt, S. A., Peglar, S., & Williams, C. (2014). Hominin footprints from early Pleistocene
deposits at Happisburgh, UK. PLoS One, 9(2), e88329.
Avanzini, M., & Lockley, M. (2002). Middle Triassic archosaur population structure: Interpretation
based on Isochirotherium delicatum fossil footprints (Southern Alps, Italy). Palaeogeography,
Palaeoclimatology, Palaeoecology, 185(34), 391402.
Balashova, A., Mattsson, H. B., Hirt, A. M., & Almqvist, B. S. (2016). The Lake Natron footprint
tuff (northern Tanzania): Volcanic source, depositional processes and age constraints from eld
relations. Journal of Quaternary Science, 31(5), 526537.
Bates, K. T., Savage, R., Pataky, T. C., Morse, S. A., Webster, E., Falkingham, P. L., Ren, L., Qian,
Z., Collins, D., Bennett, M. R., & McClymont, J. (2013). Does footprint depth correlate with
foot motion and pressure? Journal of the Royal Society Interface, 10, 20130009. https://doi.org/
10.1098/rsif.2013.0009.
Belvedere, M., Bennett, M. R., Marty, D., Budka, M., Reynolds, S. C., & Bakirov, R. (2018). Stat-
tracks and mediotypes: powerful tools for modern ichnology based on 3D models. PeerJ, 6,
e4247. https://doi.org/10.7717/peerj.4247.
Bennett, M. R., & Budka, M. (2018). Digital technology for forensic footwear analysis and
vertebrate ichnology. Dordrecht: Springer.
Bennett, M. R., & Morse, S. A. (2014). Human footprints: Fossilised locomotion? Dordrecht:
Springer.
Bennett, M. R., Harris, J. W. K., Richmond, B. G., Braun, D. R., Mbua, E., Kiura, P., Olago, D.,
Kibunjia, M., Omuombo, C., Behrensmeyer, A. K., Huddart, D., & Gonzalez, S. (2009). Early
Hominin foot morphology based on 1.5 million year old footprints from Ileret, Kenya. Science,
323, 11971201.
Bennett, M. R., Gonzalez, S., Huddart, D., Kirby, J., & Toole, E. (2010). Probable Neolithic
footprints preserved in inter-tidal peat at Keng, South Wales (UK). Proceedings of the
GeologistsAssociation, 121,6676.
Bennett, M. R., Falkingham, P., Morse, S. A., Bates, K., & Crompton, R. H. (2013). Preserving the
impossible: Conservation of soft-sediment hominin footprint sites and strategies for three-
dimensional digital data capture. PloS One, 8, e60755. https://doi.org/10.1371/journal.pone.
0060755.
Bennett, M. R., Morse, S. A., & Falkingham, P. L. (2014). Tracks made by swimming hippopotami:
An example from Koobi Fora (Turkana Basin, Kenya). Palaeogeography, Palaeoclimatology,
Palaeoecology, 409,923.
Bennett, M. R., Reynolds, S. C., Morse, S. A., & Budka, M. (2016a). Laetolis lost tracks: 3D
generated mean shape and missing footprints. Scientic Reports, 6, 21916. (2016). https://doi.
org/10.1038/srep21916.
Bennett, M. R., Reynolds, S. C., Morse, S. A., & Budka, M. (2016b). Footprints and human
evolution: Homeostasis in foot function? Palaeogeography, Palaeoclimatology,
Palaeoecology, 461, 214223.
Berge, C., Penin, X., & Pellé, É. (2006). New interpretation of Laetoli footprints using an
experimental approach and Procrustes analysis: Preliminary results. Comptes Rendus Palevol,
5, 561569.
Bibi, F., Kraatz, B., Craig, N., Beech, M., Schuster, M., & Hill, A. (2012). Early evidence for
complex social structure in Proboscidea from a late Miocene trackway site in the United Arab
Emirates. Biology Letters, 8, 670673.
Bustos, D., Jakeway, J., Urban, T. M., Holliday, V. T., Fenerty, B., Raichlen, D. A., Budka, M.,
Reynolds, S. C., Allen, B. D., Love, D. W., Santucci, V. L., Odess, D., Willey, P., McDonald,
G., & Bennett, M. R. (2018). Footprints preserve terminal Pleistocene hunt? Human-sloth
interactions in North America. Science Advances, 4, eaar7621. https://doi.org/10.1126/sciadv.
aar7621.
36 M. R. Bennett and S. C. Reynolds
Cohen, A., Lockley, M., Halfpenny, J., & Michel, A. E. (1991). Modern vertebrate track taphonomy
at Lake Manyara, Tanzania. PALAIOS, 6, 371389.
Cohen, A. S., Halfpenny, J., Lockley, M., & Michel, E. (1993). Modern vertebrate tracks from Lake
Manyara, Tanzania and their paleobiological implications. Paleobiology, 19, 433458.
Crompton, R. H., Pataky, T. C., Savage, R., DAoût, K., Bennett, M. R., Day, M. H., Bates, K.,
Morse, S. A., & Sellers, W. I. (2012). Human-like external function of the foot, and fully upright
gait, conrmed in the 3.66 million year old Laetoli hominin footprints by topographic statistics,
experimental footprint-formation and computer simulation. Journal of the Royal Society Inter-
face, 9, 707719. https://doi.org/10.1098/rsif.2011.0258.
Cumming, D. H., & Cumming, G. S. (2003). Ungulate community structure and ecological
processes: Body size, hoof area and trampling in African savannas. Oecologia, 134(4),
560568.
Day, M. H., & Wickens, E. H. (1980). Laetoli Pliocene hominid footprints and bipedalism. Nature,
286, 385.
Dingwall, H. L., Hatala, K. G., Wunderlich, R. E., & Richnmond, B. G. (2013). Hominin stature,
body mass, and walking speed estimates based on 1.5 million-year-old fossil footprints at Ileret,
Kenya. Journal of Human Evolution, 64, 556568. https://doi.org/10.1016/j.jhevol.2013.02.
004.
Falkingham, P. L., Bates, K. T., Margetts, L., & Manning, P. L. (2011). The Goldilockseffect:
Preservation bias in vertebrate track assemblages. Journal of the Royal Society Interface, 8,
11421154.
Falkingham, P. L., Bates, K. T., Avanzine, M., Bennett, M., Bordy, E., Breithaupt, B. H., Castanera,
D., Citton, P., Díaz-Marinez, I., Farlow, J. O., Fiorillo, A. R., Gatesy, S. M., Getty, P., Hatala,
K. G., Hornung, J. J., Hyatt, J. A., Klein, H., Lallensack, J. N., Martin, A. J., Marty, D.,
Matthew, N. A., Meyer, C. A., Milan, J., Minter, N. J., Razzolini, N. L., Romilio, A., Salisbury,
S. W., Scicio, L., Tanaka, I., Wiseman, A. L. A., Xing, L. D., & Belvedere, M. (2018). A
standard protocol for documenting modern and fossil ichnolgoical data. Palaeontology, 61, 469.
https://doi.org/10.1111/pala.12373.
Friess, M. (2010). Calvarial shape variation among Middle Pleistocene hominins: An application of
surface scanning in palaeoanthropology. Comptes Rendus Palevol, 9(6), 435443.
Funston, P. J., Frank, L., Stephens, T., Davidson, Z., Loveridge, A., Macdonald, D. M., Durant, S.,
Packer, C., Mosser, A., & Ferreira, S. M. (2010). Substrate and species constraints on the use of
track incidences to estimate African large carnivore abundance. Journal of Zoology, 281,5665.
Gómez-Robles, A., Martinón-Torres, M., Bermúdez de Castro, J. M., et al. (2008). Geometric
morphometric analysis of the crown morphology of the lower rst premolar of hominins, with
special attention to Pleistocene Homo. Journal of Human Evolution, 55, 627638.
Gompper, M. E., Kays, R. W., Ray, J. C., Lapoint, S. D., Bogan, D. A., & Cryan, J. R. (2006). A
comparison of noninvasive techniques to survey carnivore communities in northeastern North
America. Wildlife Society Bulletin, 34, 11421151.
Hatala, K. G., Demes, B., & Richmond, B. G. (2016a). Laetoli footprints reveal bipedal gait
biomechanics different from those of modern humans and chimpanzees. Proceedings of the
Royal Society B, 283, 20160235. (2016). https://doi.org/10.1098/rspb.2016.0235.
Hatala, K. G., Roach, N. T., Ostrofsky, K. R., Wunderlich, R. E., Dingwall, H. L., Villmoare, B. A.,
Green, D. J., Braun, D. R., & Richmond, B. G. (2016b). Footprint reveal direct evidence of
group behavior and locomotion of Homo erectus. Scientic Reports, 6, 28766. https://doi.org/
10.1038/srep28766.
Hatala, K. G., Roach, N. T., Ostrofsky, K. R., Wunderlich, R. E., Dingwall, H. L., Villmoare, B. A.,
Green, D. J., Braun, D. R., Harris, J. W., Behrensmeyer, A. K., & Richmond, B. G. (2017).
Hominin track assemblages from Okote member deposits near Ileret, Kenya, and their implica-
tions for understanding fossil hominin paleobiology at 1.5 Ma. Journal of Human Evolution,
112,93104.
Hatala, K. G., Perry, D. A., & Gatesy, S. M. (2018). A biplanar X-ray approach for studying the 3D
dynamics of human track formation. Journal of Human Evolution, 121, 104118.
Helm, C. W., McCrea, R. T., Cawthra, H. C., Lockley, M. G., Cowling, R. M., Marean, C. W.,
Thesen, G. H., Pigeon, T. S., & Hattingh, S. (2018). A new Pleistocene Hominin Tracksite from
the Cape South Coast, South Africa. ScienticReports, 8, 3772. (2018). https://doi.org/10.1038/
s41598-018-22059-5.
2 Inferences from Footprints: Archaeological Best Practice 37
Kinahan, J. (2013). The use of skeletal and complementary evidence to estimate human stature and
identify the presence of women in the recent archaeological record of the Namib desert.
South African Archaeological Bulletin, 68,7278.
Leakey, M. D., & Harris, J. M. (Eds.). (1987). Laetoli: A Pliocene site in Northern Tanzania.
Oxford: Clarendon Press.
Leakey, M. D., & Hay, R. L. (1979). Pliocene footprints in the Laetoli beds at Laetoli, Northern
Tanzania. Nature, 278, 317.
Lee, P. C., & Moss, C. J. (1995). Statural growth in known-age African elephants (Loxodonta
africana). Journal of Zoology, 236,2941.
Liutkus-Pierce, C. M., Zimmer, B. W., Carmichael, S. K., McIntosh, W., Deino, A., Hewitt, S. M.,
McGinnis, K. J., Hartney, T., Brett, J., Mana, S., & Deocampo, D. (2016). Radioisotopic age,
formation, and preservation of late Pleistocene human footprints at Engare Sero, Tanzania.
Palaeogeography, Palaeoclimatology, Palaeoecololgy, 463,6882.
Lockley, M. G. (1986). The paleobiological and paleoenvironmental importance of dinosaur
footprints. PALAIOS, 1,3747.
Martin, A. J., & Pyenson, N. D. (2005). Behavioral signicance of vertebrate trace fossils from the
union chapel site. Pennsylvanian footprints in the black Warrior Basin of Alabama. Alabama
Paleontological Society Monograph, 1,5973.
Marty, D., Strasser, A., & Meyer, C. A. (2009). Formation and taphonomy of human footprints in
microbial mats of present-day tidal-at environments: Implications for the study of fossil
footprints. Ichnos, 16, 127142.
Marty, D., Falkingham, P. L., & Richter, A. (2016). Dinosaurs track terminology: A glossary of
terms. In P. L. Falkingham, D. Marty, & A. Richter (Eds.), Dinosaur tracks: The next steps
(pp. 399400). Bloomington/Indianapolis: Indiana University Press.
Masao, F. T., Ichumbaki, E. B., Cherin, M., Barili, A., Boschian, G., Iurino, D. A., Menconero, S.,
Moggi-Cecchi, J., & Manzi, G. (2016). New footprints from Laetoli (Tanzania) provide
evidence for marked body size variation in early hominins. eLife, 5.https://doi.org/10.7554/
eLife.19568.
McLaren, D., Fedje, D., Dyck, A., Mackie, Q., Gauvreau, A., & Cohen, J. (2018). Terminal
Pleistocene epoch human footprints from the Pacic coast of Canada. PLoS One, 0193522.
https://doi.org/10.1371/journal.pone.0193522.
McNeil, P., Hills, L. V., Kooyman, B., & Tolman, S. M. (2005). Mammoth tracks indicate a
declining late Pleistocene population in southwestern Alberta, Canada. Quaternary Science
Review, 24, 12531259.
Meldrum, D. J., Lockley, M. G., Lucas, S. G., & Musiba, C. (2011). Ichnotaxonomy of the Laetoli
trackways: The earliest hominin footprints. Journal of African Earth Science, 60,112.
Miller, K. V., Marchinton, R. L., & Nettles, V. F. (1986). The growth rate of hooves of white-tailed
deer. Journal of Wildlife Diseases, 22, 129131.
Moreira, D. O., Alibhai, S. K., Jewell, Z. C., da Cunha, C. J., Seibert, J. B., & Gatti, A. (2018).
Determining the numbers of a landscape architect species (Tapirus terrestris), using footprints.
PeerJ, 6,e4591.
Morse, S. A., Bennett, M. R., Liutkus-Pierce, C., Thackeray, F., McClymont, J., Savage, R., &
Crompton, R. H. (2013). Holocene footprints in Namibia: The inuence of substrate on footprint
variability. American Journal of Physical Anthropology, 151, 265279.
Musiba, C. M., Tuttle, R. H., & Hallgrímsson, B. (1997). Swift and sure-footed on the Savanna: A
study of Hadzabe gaits and feet in Northern Tanzania. American Journal of Human Biology, 9,
303321.
Musiba, C. M., Mabula, A., Selvaggio, M., & Magori, C. C. (2008). Pliocene animal trackways at
Laetoli: Research and conservation potentials. Ichnos, 15, 166178.
Ostrom, J. H. (1972). Were some dinosaurs gregarious? Palaeogeography, Palaeoclimatology,
Palaeoecology, 11, 287301.
38 M. R. Bennett and S. C. Reynolds
Parés-Casanova, P. M., & Oosterlinck, M. (2012a). Hoof size and symmetry in young catalan
pyrenean horses reared under semi-extensive conditions. Journal of Equine Veterinary Science,
32, 231234.
Parés-Casanova, P. M., & Oosterlinck, M. (2012b). Relation between hoof area and body mass in
ungulates reared under semi-extensive conditions in the Spanish Pyrenees. Journal of Animal
Science Advances, 2, 374379.
Pastoors, A., Lenssen-Erz, T., Ciqae, T., Kxunta, U., Thao, T., Bégouën, R., Biesele, M., & Clottes,
J. (2015). Tracking in Caves: Experience based reading of Pleistocene human footprints in
French caves. Cambridge Archaeological Journal, 25, 551564.
Pastoors, A., Lenssen-Erz, T., Breuckmann, B., Ciqae, T., Kxunta, U., Rieke-Zapp, D., & Thao,
T. (2016). Experience based reading of Pleistocene human footprints in Pech-Merle. Quater-
nary International.https://doi.org/10.1016/j.quaint.2016.02.056.
Prins, H. H. T., & Reitsma, J. M. (1989). Mammalian biomass in an African equatorial rain forest.
Journal of Animal Ecology, 58, 851861.
Roach, N. T., Hatala, K. G., Ostrofsky, K. R., Villmoare, B., Reeves, J. S., Du, A., Braun, B. R.,
Harris, J. W. K., Behrensmeyer, A. K., & Richmond, B. G. (2016). Pleistocene footprints show
intensive use of lake margin habitats by Homo erectus groups. Scientic Reports, 6, 26374.
https://doi.org/10.1038/srep26374.
Roach, N. T., Du, A., Hatala, K. G., Ostrofsky, K. R., Reeves, J. S., Braun, D. R., Harris, J. W.,
Behrensmeyer, A. K., & Richmond, B. G. (2018). Pleistocene animal communities of a 1.5
million-year-old lake margin grassland and their relationship to Homo erectus paleoecology.
Journal of Human Evolution, 122,7083.
Roberts, G., Gonzalez, S., & Huddart, D. (1996). Intertidal Holocene footprints and their archae-
ological signicance. Antiquity, 70, 647651.
Roberts, D. L., Bateman, M. D., Murray-Wallace, C. V., Carr, A. S., & Holmes, P. J. (2008). Last
interglacial fossil elephant trackways dated by OSL/AAR in coastal aeolianites, still bay,
South Africa. Palaeogeography,. Palaeoclimatology, Palaeoecology, 257, 261279.
Shrader, A. M., Ferreira, S. M., McElveen, M. E., Lee, P. C., Moss, C. J., & Van Aarde, R. J.
(2006). Growth and age determination of African savanna elephants. Journal of Zoology, 270,
4048.
Silveira, L., Jacomo, A. T., & Diniz-Filho, J. A. F. (2003). Camera trap, line transect census and
track surveys: A comparative evaluation. Biological Conservation, 114, 351355.
Stachurska, A., Kolstrung, R., Pięta, M., & Silmanowicz, P. (2011). Hoof size as related to body
size in the horse (Equus caballus). Animal Science Papers Reports, 29, 213222.
Topinard, P. (1877). Anthropologie. Paris: C. Reinwald.
Urban, T. M., Bustos, D., Jakeway, J., Manning, S. W., & Bennett, M. R. (2018). Use of
magnetometry for detecting and documenting multi-species Pleistocene megafauna tracks at
White Sands National Monument, New Mexico, USA (Vol. 199, pp. 206213). Quaternary
Science Reviews.
Western, D., Moss, C., & Georgiadis, N. (1983). Age estimation and population age structure of
elephants from footprint dimensions. Journal of Wildlife Management, 47, 11921197.
Wiseman, A. L., & De Groote, I. (2018). A three-dimensional geometric morphometric study of the
effects of erosion on the morphologies of modern and prehistoric footprints. Journal of
Archaeological Science Reports, 17,93102.
Zelditch, M. L., Swiderski, D. L., & Sheets, H. D. (2012). Geometric morphometrics for biologists:
Aprimer. Amsterdam: Academic.
Zimmer, B., Liutkus-Pierce, C., Marshall, S. T., Hatala, K. G., Metallo, A., & Rossi, V. (2018).
Using differential structure-from-motion photogrammetry to quantify erosion at the Engare Sero
footprint site, Tanzania. Quaternary Science Reviews, 198, 226241.
2 Inferences from Footprints: Archaeological Best Practice 39
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Chapter 3
Repetition Without Repetition: A
Comparison of the Laetoli G1, Ileret,
Namibian Holocene and Modern Human
Footprints Using Pedobarographic
Statistical Parametric Mapping
Juliet McClymont and Robin H. Crompton
Abstract It is traditionally held that early hominins of the genus Australopithecus
had a foot transitional in function between that of the other great apes and our own
but that the appearance of genus Homo was marked by evolution of an essentially
biomechanically modern foot, as well as modern body proportions. Here, we report
the application of whole foot, pixel-wise topological statistical analysis, to compare
four populations of footprints from across evolutionary time: Australopithecus at
Laetoli (3.66 Ma, Tanzania), early African Homo from Ileret (1.5 Ma, Kenya) and
recent modern (presumptively habitually barefoot) pastoralist Homo sapiens from
Namibia (Holocene), with footprints from modern Western humans. Contrary to
some previous analyses, we nd that only limited areas of the footprints show any
statistically signicant difference in footprint depth (used here as an analogy for
plantar pressure). A need for this comparison was highlighted by recent studies using
the same statistical approach, to examine variability in the distribution of foot
pressure in modern Western humans. This study revealed very high intra-variability
(mean square error) step-to-step in over 500 steps. This result exemplies the
fundamental movement characteristic of dynamic biological systems, whereby
regardless of the repetition in motor patterns for stepping, and even when
constrained by experimental conditions, each step is unique or non-repetitive;
hence, repetition without repetition. Thus, the small sample sizes predominant in
the fossil and ichnofossil record do not reveal the fundamental neurobiological
driver of locomotion (variability), essentially limiting our ability to make reliable
interpretations which might be extrapolated to interpret hominin foot function at a
population level. However, our need for conservatism in our conclusions does not
equate with a conclusion that there has been functional stasis in the evolution of the
hominin foot.
J. McClymont · R. H. Crompton (*)
University of Liverpool, Liverpool, UK
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_3
41
Keywords Australopithecus · Barefoot · Neurobiological degeneracy ·
Pedobarography · Redundancy · Variation
Introduction
42 J. McClymont and R. H. Crompton
The origins and evolution of human striding bipedalism have long been a focus of
human palaeontology and evolutionary biomechanics. However, the fossil evidence
for the evolution of the postcranial skeleton has not been an unambiguous source of
information. Claims that morphological features taken as human adaptations for
terrestrial bipedalism reduce effectiveness in arboreal climbing are challenged by the
combination of both capabilities in several indigenous modern human populations
(Venkataraman et al. 2013a,b). This is a clear demonstration of neurobiological
degeneracy (Seifert et al. 2016) and the high variability necessary for dynamic
systems (Davids et al. 2003). These two theories underpinning biomechanical
movement and locomotor adaptation are reviewed in the references provided but
are not discussed in great detail here due to the nature of this publication. Briey,
however, both concepts can be illustrated by the everyday phrase, there are many
ways to skin a cat. Thus, the long femoral necks and small femoral heads together
with aring iliac crests found in many australopiths (McHenry 1975) (but due to
high biological variability, emphatically not all), versus large femoral heads and
short femoral necks with sigmoid, non-aring iliac crests in ourselves, must be
interpreted as reecting naturally high variation in biological forms. Undoubtedly,
the mechanics of hip adduction and abduction must have been different in some
australopiths; however, variation in morphology may act to achieve the same
biomechanical effect on joint systems in different ways. Take, for example, clear
evidence of facultative upright bipedal behaviour in Gorilla gorilla (see, e.g. Watson
et al. 2009). Gorilla morphology isnt designed specically for upright bipedalism,
but natural biological variation permits it. Very few papers in primatology, hominin
palaeontology and human and non-human ape ichnopalaeontology address variation
in bone morphology and footprint topology within the context of the locomotor
system. Small fossil samples are often described as key for understanding human
locomotor evolution or morphofunctional behaviour. Such relationships are often
claimed from the evidence of a single individual without including either morpho-
logical variation or facultative capabilities. It is more reasonable to recognize that
high natural intra-individual facultative variation elicits high inter-individual
morphofunctional variation at a population level and vice versa. Both are a conse-
quence of complex neurobiological evolution in biological systems. A noteworthy
exception to a general lack of investigation of variability, however, is Dunn et al.
(2014) on the Gorilla talus.
Leading cognitive and ecological motor skills specialists address biomechanical
variation anchored by the theoretical paradigms of dynamical systems (Thelen 1995,
2005; Davids et al. 2003; Bartlett et al. 2007) and neurobiological degeneracy
(Edelman and Gally 2001; Seifert et al. 2016). These paradigms have been increas-
ingly inuential in biomechanics, especially sports science. However, hominin
palaeontology has not yet taken these advances on board, while further being
hampered by small sample size, which make it difcult to incorporate an under-
standing of the impact on variability in biological systems on evolutionary interpre-
tations. For example, the biomechanical complexity involved in taking a step begins
at the hip joint, a simple, single ball joint articulation, while the knee joint comprises
the biarticular joint between femur and tibia and a third, sliding joint between
the tibia and bula. Thus, the knee joint is kinematically complex, with sliding
(essentially planar), as well as rotatory motion. However, the structure predomi-
nantly concerned in transferring forces to the substrate (whether ground, branches or
any other surface) is the foot. Here we face 26 bones (excluding sesamoids) which
form 33 joints and are controlled by over 100 muscles, tendons and ligaments. Thus,
given variation in the complexity of biomechanical forms, and a shortage of fossils,
it is no wonder that functional interpretations of the evolution of the foot have been
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 43
and still are interpreted in different and contradicting ways.
For example, the OH-8 Homo habilis foot has been described as (non-human)
ape-like in some joints but not others (e.g. Kidd et al. 1996; and see Harcourt-Smith
and Aiello 2004) but elsewhere more or less entirely humanlike in function (e.g. Day
and Napier 1964). What humanlike in function implies for gait is obvious (habitually
upright, striding steps), but what is the signicance for gait of describing a foot as a
mosaic of non-human ape-like and humanlike joints, when there are 33 joints to
consider? In engineering parlance, such a complex system is described both as
functionally redundant (there are many neuromechanical pathways to achieve a
consistent motor pattern) and that its determinacy is low (it will be difcult to predict
how the system will act in different iterations of the same task). Perhaps the most
accessible review of these concepts as applied to biological systems can be found in
Alexander (2003).
Since foot structure is so complex, and redundancy therefore so high, our
condence in eliciting functional information at a population level about species
gait from comparisons of individual foot bones (e.g. Jungers et al. 2009;Wa
rdetal.
2011) must be fairly low. There is some potential however to interpret biomechan-
ical variability retrospectively from topological features of fossil footprint trails.
1
The basis of this potential is that a natural relationship between forces exerted on the
ground by the foot, to balance, propel and control walking, and the consequent
deformation of the ground could logically exist, given that
F
p¼A
1
Acaveat is required for the present paper: given our brief discussion here on variability and sample
size, interpretations of the variability presented herein are made purely for the trackmaker as they took
the 11 steps in this sample and cannot be extrapolated to predictions regarding a population locomotor
mode. Furthermore, this is not an assessment of functional variability as would be required for inferring
stability and balance behaviour: that would require thousands of steps. For an excellent tutorial on
functional variability, equations and analysis techniques, see Bruijn et al. 2013.
where pressure (p) equals the amount of force (F) (the scalar of which is measured in
Pascals (Pa)), acting per unit area (A) (Giancoli 2004). However, substrates, no
matter their composition, will always reach a point at which maximum body weight
is fully supported, and thus, the substrate stops deforming even though pressure is
still being applied. Substrate composition at and below ground surface, substrate
moisture content, even the electrical charge at the surface of substrate particles and a
host of other factors can, however, be expected to interact with deforming forces
delivered by the plantar surface of the foot. Thus, the shape of a footprint does not
mirror the foot that made it due to substrate effects. We explored some of these
interactions in Bates et al. (2013) (and see Supplementary Material) and concluded
that the relationship of foot pressure and print depth varies with substrate compli-
44 J. McClymont and R. H. Crompton
ance: substrate moisture and presence and depth of subsurface compaction levels,
but also the mechanical requirements at toe off, inuencing print topology.
A crucial recent discovery that of StW 573 Australopithecus prometheus,
3.67 Ma, which is over 90% complete (see e.g. Clarke 2019;C
romptonetal.
2018) crucially, closely similar in date to the Laetoli footprint trails. The best
known and most complete early human ancestor was Australopithecus afarensis,
represented by the diminutive AL-288-1 Lucy skeleton, 3.4 Ma. Thus, despite
her relatively late date, most locomotor interpretations of the Laetoli footprint
trails have been based on the AL-288-1 Lucy skeleton (e.g. Crompton et al. 1998,
2012).
Since discovery of this partial (circa 30% complete) skeleton, her combination
of humanlike knees with reconstructed limb proportions that were thought to
indicate long arms and short legs, and her clear digital curvature, has fostered a
long-term dispute on her locomotor behaviour. Some (e.g. Stern and Susman 1983,
1991;Sternetal.1984) assert that these traits would have compromised her
terrestrial bipedalism, so that she would have walked with a bent hip and knee
(BHBK) posture and even a somewhat shufing gait. Opposing are those who
argue that she was an effective terrestrial biped and that the arboreal features are
simply retained anachronisms (Latimer et al. 1987; Latimer and Lovejoy 1989;
Latimer 1991).
Computer simulation and experimental studies have since shown that even
her proportions as rst reconstructed with effective fully upright bipedalism
(e.g. Crompton et al. 1998;K
ramer1999; Kramer and Eck 2000) and that a BHBK
gait in humans causes an unsustainable rise in core body temperature within 5 min of
walking (Carey and Crompton 2005). Given the general similarity of muscular
physiology in placental mammals, and all other things being equal, forwards dynamic
modelling has, similarly, predicted BHBK gait to near double the metabolic costs of
transport in Au. afarensis (Sellers et al.2005;Naganoetal.2005). More recently, other
partial skeletons of this species, most notably KSD/VP 1-1 from Woranso-Mille
(Lovejoy et al. 2016), but also other isolated bones or partial material from Afar
(including material referred to AL-333, see, e.g. McHenry 1986), have shown that
Lucys small stature and long forelimbs are the exception rather than the rule. In fact,
analysis of the StW 573 longbones now suggests that AL-288-1 probably did not have
particularly short legs in relation to arm length (Heaton et al. 2019). Au. afarensis now
appears to be a very variable species both in stature and postcranial morphology, and
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 45
this is further attested to by KSD/VP 1-1 being assigned to this species by Lovejoy
et al. (2016).
To gain further enlightenment on the mode of locomotion in Au. afarensis,
several groups have analysed the penecontemporaneous Laetoli footprint trails,
holding that on the basis of the above-cited equation, footprint depth and topograph-
ical features must at the very least reect foot-ground interactions. Early attempts
focussed on features of single footprints chosen from the clearer G1 trail, and some
declared that the footprints are essentially modern in character (Day and Wickens
1980; White1980). Others argued that features such as a relatively abducted hallux
with limited hallux print depth support a BHBK model for gait (Stern and Susman
1983). However, White and Suwa (1987) regard these features as taphonomic
artifacts. This more than 30-year-old debate continues today.
Although it is now broadly accepted that selection of single prints for study is
inappropriate, Meldrum and colleagues, as late as 2011, claim that a line in one
footprint shows a chimpanzee-like mid-tarsal break, so claiming that Au. afarensis
lacked a medial longitudinal arch (Meldrum et al. 2011). An opposing argument
based on discussions with the chief taphonomist of the Laetoli footprint trails (pers.
comm. Craig Feibel to RHC) suggests that this topographical feature is simply a
product of natural sedimentological fracture in the substrate over time although this
would require micro-sedimentological analysis to conrm. Statistical and biome-
chanical approaches to the Laetoli footprint trails have predicted the stride length,
foot shape, body proportions and speed of the trackmaker (Alexander 1984; Reyn-
olds 1987; Raichlen et al. 2010), and spatio-temporal characteristics of the same trail
have been used as an analogy to predict speed of walking and energetic costs in Au.
afarensis (to date, primarily AL-288-1) (Kramer and Eck 2000; Sellers et al. 2005).
Hatala et al. (2016)compared just 5 of the 11 taphonomically usable footprints
from Laetoli G1 using an inappropriate regionalized (and hence anatomically biased,
see, e.g. Pataky et al. 2011) topological analysis, to prints made by modern humans
and bipedally walking chimpanzees. The authors concluded that topological features
from the Laetoli G1 prints are evidence for a functionally unique locomotor mode.
Specically, the authors claim to be able to identify kinematic distinctions in foot
and lower limb function and that the trackmaker probably walked with a more exed
knee posture, describing it as a form of bipedalism that was well developed but not
equivalent (Hatala et al. 2016) to that of modern humans. Raichlen et al. (2010)
found that a simple whole foot statistical comparison of heel and toe depths in the
11 usable prints indicated a fully upright posture. Crompton et al. (2012) used a
rigorous combination of topographical whole foot statistical analysis and computer
modelling to compare the mean tendency of the 11 usable G1 prints, predicting foot
pressure in upright and BHBK gait. The authors conclude, similarly to Raichlen and
colleagues, that they cannot have been made by an individual walking BHBK and
were more likely left by an upright striding biped. Raichlen and Gordons(2017)
preliminary statistical comparison of heel and toe depth conrms these ndings for
the more extensive Laetoli S series, which were made by individuals of greatly
varying stature.
) made a broader comparison of the Laetoli G1 trail
both to prints from Ileret, made presumptively by early African Homo erectus and to
Holocene pastoralist footprints from Namibia. On the basis of footprint depth,
substrate conditions at the time of footprint formation of the Namibian trails,
which cross from drier sandy bank sediments, through muds, and back to drier
bank deposits in an ancient streambed, bridged those at Laetoli (relatively shallow
and only slightly moist) and Ileret (deep and wet muds). The Namibian trackways,
made in the Holocene age, were made by presumptively habitually barefoot indi-
viduals, and given the alteration that footwear induces in human plantar pressure,
they are an invaluable control. Using third-party open-source code to derive mean
and median tendencies of the tracks, they conclude that there is functional stasis
between the 3.66 Ma (Crompton et al.
Recently, Bennett et al. (
trate, registration and re-registration), together with all data processing and method-
ological sensitivity checks.
2016
2012) Laetoli G1 trails and the circa 1.5 Ma
(Bennett et al. 2009) Ileret trails. Inasmuch as this implies a fully upright gait at the
time of footprint formation at Laetoli G1, this study is in accord with both that of
Raichlen et al. (2010) and Crompton et al. (2012). Each of these studies used at least
twice as many G1 footprints as did Hatala et al. (2016), raising the possibility that
sample size, and further possible loss of variability between footprints due to their
46 J. McClymont and R. H. Crompton
subjective selection of only ve footprints, could account for their very different
interpretations.
Because of this contradiction, an extended statistical analysis of variability in
footprint topology using pedobarographic statistical parametric analysis (pSPM) of
the Laetoli G1, Ileret and Namibian fossil footprint trails in comparison with
experimental modern human plantar pressure records is presented here.
Methods
We employ the robust method of statistical parametric mapping (SPM), a topograph-
ical statistical approach rst developed by Friston et al. (1995) for functional brain
imaging and extensively validated by that group (open source). The algorithms have
been further developed for foot pressure studies and incorporated into our open-
source software pedobarographic statistical parametric mapping (pSPM), which has
been further and extensively validated (e.g. Pataky 2010; Pataky and Goulermas
2008; Pataky et al. 2008,2011). In its analogous extension to footprint depth,
statistical comparison of the samples also uses pixel-level pairwise t-tests (Pataky
and Goulermas 2008; Pataky et al. 2008) but here after normalization by plantar
surface maximum depths. Methods follow Crompton et al. (2012) and Bates et al.
(2013) except that we now employ automated registration where possible and use an
enhanced method for isolating and normalizing prints. These changes and a full
description of the method are presented in Supplementary Material (Fig. 3.4, data
processing prior to registration. Figures 3.5,3.6,3.7 and 3.8 diagramatically illus-
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 47
Results and Interpretations
Figure 3.1 presents the mean footprints and the results of topological statistical
comparisons of 11 prints from the G1 trail at Laetoli (Leakey and Hay 1979; Leakey
and Harris 1987), with 9 prints from the upper surface at Ileret (FwJj14E; Kenya)
(Bennett et al. 2009), a 32 print sample from the Holocene trail at Walvis Bay,
Namibia (Kinahan 1996; Morse et al. 2013), and a modern Western (thus habitually
shoe wearing) dataset collected on a treadmill (N ¼100 pressure records registered
to create one mean print each from 10 individuals).
Habitually shod Western modern and presumably habitually unshod Holocene
modern human footprints (Fig. 3.1a) show no areas of signicant difference. The
Ileret dataset differs signicantly from that for modern Western humans ( p¼0.000),
(Fig. 3.1b) showing a deeper medial arch. The Laetoli mean shows signicantly
deeper medial arch and anterior heel impressions than the modern Western human
mean (Fig. 3.1c) and signicantly shallower hallucal impressions. The Ileret foot-
prints are signicantly different from the Holocene modern human mean (Fig. 3.1d),
having a shallower medial arch and deeper distal toes, albeit under a small area in the
midfoot (p¼0.044) and restricted to print edges. The latter could be the result of
imperfect registration due to minor overall shape differences in the two populations
or the subject dragging the foot from where it would have been sunken into the soft
sediment. Most notably, the Holocene modern humans from Namibia and Laetoli
means differ signicantly (Fig. 3.1e) in only a very small area under the hallux.
Fig. 3.1 Statistical whole foot comparison of four footprint populations. The rst and second
footprint images in each comparison (af) are the mean image from the two species being
compared. The third footprint image is the comparison of the two population means. The fourth
footprint image is the inference plot determining the pixel-level locations of statistically signicant
differences between the two populations being compared; (a) habitually shod anatomically modern
H. s. sapiens vs. habitually unshod H. s. sapiens;(b) habitually shod H. s. sapiens vs. early African
Homo erectus footprints from Ileret; (c) habitually shod H. s. sapiens vs. Laetoli footprints; (d)
presumptively habitually unshod (Namibia) H. s. sapiens vs. Ileret Homo erectus footprints; (e)
habitually unshod H. s. sapiens vs. Laetoli footprints; (f) Ileret Homo erectus vs. Laetoli footprints.
Inference plots represent the probability values for areas with statistically signicant differences in
footprint depth and are represented in the far-right column of each comparison. This column is
blank in (a) designating no statistically signicant differences in topology in this comparison
Finally, statistically signicant differences between the Ileret and Laetoli means
(Fig. 3.1f) exist in deeper impressions under small areas of MTH1, the hallux and the
posterior medial heel.
48 J. McClymont and R. H. Crompton
Visual inspection of the experimental footprints provided in the contribution by
Hatala et al. (2016) reveals that similarities in the forefoot and hallux depths of their
modern human and selected Laetoli footprints more than exist with their selected
chimpanzee footprint. The statistically signicant deeper medial midfoot impression
in Laetoli ( p¼0.000) than in Western modern humans (Fig. 3.1c), also reported by
Hatala et al. (2016), could be attributed to the effect of habitual shoe wearing in
Western modern humans, creating a higher medial arch in this group (Stolwijk et al.
2013). We have shown through experimental studies of the relationship of footprint
depth to footprint morphology (Bates et al. 2013) that there is a clear tendency for
deeper prints to have relatively deeper forefoot impressions. It is therefore likely that
the statistically signicant differences between Laetoli and Ileret (Fig. 3.1f), and
Ileret and Holocene modern human footprints (Fig. 3.1d) sampled here, are attrib-
utable to the greater overall footprint depth at Ileret. (The Laetoli sample showed a
mean of 31 mm, range 2637 mm; for Ileret the mean maximum plantar depth was
49 mm, the range 2494 mm; see Supplementary Material in additional footprint
discussion.) Here, moisture content was likely higher, based on sidewall suction
against the foot producing long narrow tracks; by this interpretation, the moisture
content likely weakened the sediment in which the tracks were made (Craig 1997;
Bennett et al. 2016). Similarly, the relatively greater number of deep prints from
Holocene modern humans (from a wetter substrate, mean maximum plantar depth
was 45 mm and range 2377 mm) compared to Laetoli could readily account for
deeper hallux impressions in Holocene human footprints. Crompton et al. (2012)
used computer modelling to simulate contact pressures under the foot in upright and
exed knee walking. They showed that bent knee or exed knee walking produced
higher forefoot than hindfoot pressures because of the anterior shift of the centre of
mass (CoM). Their analysis of a larger dataset including all of those prints analysed
by Hatala et al. (2016) revealed consistently deeper hindfoot than forefoot impres-
sions, indicating full extension at the knee during upright walking (Ferris et al. 1998;
see Crompton et al. 2003,2008 on the relationship between the heel-strike transient
and extended knee postures in orangutans). While any comparison of human and Au.
afarensis postcrania strongly suggests that the locomotor systems of the Laetoli
trackmaker and modern humans form biomechanically distinct kinematic chains,
this does not necessarily imply dramatically different external ability and function
(Bock 1965,1994; Laland et al. 2015; Seifert et al. 2016). This interpretation follows
the expectations from effects due to high functional redundancy (Latash et al. 2002)
and high degrees of freedom in the foot (e.g. Wolf et al. 2008), both natural and
essential components to be considered in analyses of fossil footprint trails and
explaining the difference between prints and between populations via high
variability.
Figure 3.2 represents all prints in the Laetoli G1 sample used in this analysis
alongside 11 consecutive p-images collected during treadmill walking from a
healthy human at 1.1 m/s (McClymont et al. 2016). This gure is not a statistical
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 49
25
20
15
10
5
Fig. 3.2 A visual representation of the variability in relative foot print depth (a) and plantar
pressure (b); (a) footprint of whole foot pixel-level statistical topographical depth map of the
Laetoli footprints used in this analysis, showing variable distribution of maximum depth step-to-
step. Greatest depth is evident in combinations of heel, lateral midfoot and the lateral MTH5-2 step-
to-step; (b) 11 consecutive registered whole foot pixel-level statistical pressure images collected
from a treadmill walking trial at 1.1 m/s in H. s. sapiens. Areas in yellow are areas of highest
variability (mean square error (MSE)) step-to-step. Greatest pressure is consistently evident in
combinations of heel and lateral MTH4-1
comparison of relative depths (Laetoli) and plantar pressures (modern human) as the
two samples were collected under completely different conditions. It is simply
presented to visually demonstrate the variability in each step or uctuations in
foot-ground interactions during just 11 steps. While there are undeniable differences
in foot shape and topography, they share similar variation in topology step-to-step.
The Laetoli prints (Fig. 3.2a) are consistently deeper under the heel and lateral
forefoot with steps 10 and 11 showing deeper depth under the whole forefoot than
in previous steps. The human prints (Fig. 3.2b) despite being more consistent step-
to-step due to the normalizing effects of the treadmill (Kang and Dingwell 2008) also
show consistently high variability in pressure under the heel and MTH4, 3 and 1. We
should note that the human subject does not represent the most variable subject from
our Western human sample but instead the average. Again, this gure is not intended
as a statistical comparison of relative depths and variability in pressure between the
two subjects or across deep time. It is simply presented to visually illustrate the step-
to-step variability of foot-ground interactions during locomotion despite two very
different substrates and in a relatively tiny interpretive sample of only 11 footprints
out of the thousands the individual took the day they were made. Tudor-Locke et al.
(2017)showed that the average 20-year-old American male walks between 2247 and
12,334 and females 1755 and 9824 steps per day. This underlines the requirement of
steps necessary to interpret the natural and characteristic patterns of variability that
contribute to morphofunctional interpretations of the individual making fossil foot-
print trails.
The variation in relative depth in the sequence of 11 footprints reects similar and
thus normal biomechanical variation in stride dynamics in both the G1 trackmaker
(Fig. 3.2a) following Alexanders(2003) work (and see Wainwright 1991;
Wainwright et al. 2002) and reecting Bernsteins(1967) classic description of
human movement as repetition without repetition. That is, each movement task
(e.g. step) is driven by a unique set of neural and motor patterns, temporarily
assembled to produce a task outcome (Latash et al. 2002) based on the unique
mechanics of each step and the substrate upon and environment in which it is taken.
Thus, while each step is programmed by the suit of bipedal evolutionary traits, each
step is unique. Thus, not only have we shown that there is intra-species variation in
foot pressure within the great apes, sufcient for both Asian and African apes
(orangutan and bonobo) to overlap in midfoot pressure patterns with habitually
shoe-wearing humans (Bates et al. 2013), but our interpretations here are founded
on intra-individual variation step-to-step, as predicted by Bernstein (1967) and
Latash et al. (2002).
50 J. McClymont and R. H. Crompton
As mentioned earlier a very serious limitation in the analysis of fossil footprint
data is sample size. Figure 3.3 (below) simulates the possible effects of Hatala et al.s
(2016) subjective selection of just 5 G1 footprints, by sampling the rst, middle and
last group of 5 prints from the 11 used here and available for ready analysis
(Raichlen et al. 2010; Crompton et al. 2012). Indeed, as Hatala et al. (2016)
observed, modern humans consistently have a deeper impression in the forefoot
than that of the Laetoli hominin, irrespective of which sample is selected. However,
the last nal comparison in set (Fig. 3.2c) shows (considerably smaller) areas of the
midfoot and anterior heel, where it is the Laetoli prints which are deeper. Thus,
interpreting topological differences between the G1 prints is not immune to a
subjective choice of prints, raising concerns about the conclusions of Hatala et al.
(2016). On the assumption that footprints are correlated with foot pressure, even
given the interactions with substrate characteristics alluded to above and which we
dealt with in detail in Bates et al. (2013), our concerns are very deeply amplied by
new data concerning the sample size required to reliably characterize human gait.
Arts and Bus (2011) recommend only 12 steps per foot for clinical assessment of
plantar pressure. Sample sizes of as little as 10, and at most 50, are commonly used
to assess gait parameters including pressure and kinematics in clinical practice. The
higher value of 50 slightly mitigates the effect of step-to-step variability that would
otherwise perhaps lead to false interpretations, due to the high variability step-to-step
(McClymont et al. 2016). Owings and Grabiner (2003) however have demonstrated
that sample sizes of over 100 steps are needed to reliably characterize an individuals
kinematics, to within 95% condence. Equally, McClymont (2017) showed through
a Monte Carlo subsampling analysis of random samples of >2000 footprints per
subject that the individual trial N typically collected in plantar pressure studies in the
literature (N ¼1050 p-images) produces MSE ranges that are more than 50%
higher than from when sampled from a larger total individual N of >500. At N ¼<10
records this increases to more than 75%, indicating a high probability that such a
small individual trial N would not reect either the range of variation or the habitual
mean pressure that would be represented by a larger dataset of consecutive foot print
records. Acquiring a sample of more than 500 is clearly unfeasible for footprints and
even for foot pressure in the inrm. However, samples of 100138 would deliver
around 95% condence and might be achievable in the future at Laetoli S. But even
assuming a close link between pressure and footprint depth, a sample of ve, as used
by Hatala et al. (2016), offers well under 25% condence of assessing pressure
characteristics, i.e. a probability of unreliable assessments. Indeed, our own sample
of 11 would allow no better than 50% reliability; however, we are accounting for the
effects of variability in our interpretations and not making condent claims for a new
locomotor mode, from what we know to be a biomechanically unsatisfactory
sample size.
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 51
Discussion
Based on only very minimal statistically signicant differences (just a few pixels)
between Laetoli and unshod (Holocene) modern humans (Fig. 3.1e), and the com-
monality of stride-to-stride and step-to-step uctuations illustrated in the Laetoli G-1
trail and the modern human example (Fig. 3.3a, b), we cannot nd any evidence that
the Laetoli trackmaker utilized a exed knee posture at the time of formation of the
prints examined, supporting previous results (Sellers et al. 2005; Raichlen et al.
2010; Crompton et al. 2012). The extended evolutionary synthesis (EES) (Laland
et al. 2015), and the unifying theory of dynamical biological systems (Davids et al.
2003) and neurobiological degeneracy (Seifert et al. 2016), all predict high variabil-
ity in adaptive biological systems, permitting rapid evolutionary change (Laland
et al. 2015) and stable, functional movement (Bernstein 1967; Seifert et al. 2016).
The high redundancy present in the anatomically complex structure of the foot is
likely to be employed to control step-to-step dynamic variability in walking
(Dingwell et al. 2010), and activation patterns are substantially subject to stochastic
processes, reecting neurobiological degeneracy (Seifert et al. 2016). We (Pataky
et al. 2013) also found in a large sample of foot pressure records (N ¼5243) that
autocorrelation in maximal plantar pressure between steps is very weak, such that
statistical power calculations found that a null hypothesis that local plantar pressure
values are uncorrelated in short gait bouts is likely true with an average probability of
78.9%. This is both consistent with dynamical systems theory and very worrying for
the analysis of short/discontinuous trails such as those at Ileret and similarly for the
Hatala et al. G1 sample of ve. While we have not attempted to quantify dynamic
behaviour here, when pressure is taken as analogous to, but not equivalent to, depth
in the Laetoli footprints, we can infer a exible, upright hominin gait variably
resisting perturbations to stabilize the CoM across the hot, damp ash (Dingwell
et al. 2010). Recent evidence from a variety of substrates found that forefoot depth
increased with moisture content in a modern human sample (Bates et al. 2013),
leading to a requirement for increased forefoot forces to clear the foot from the
substrate. We conclude that the differences in relative forefoot depths are a product
of substrate, specically of high moisture content in the modern human experimental
sample, Ileret, and part of the Walvis Bay trail, versus relatively low moisture
content (Craig 1997) at Laetoli.
Further visual inspection of the experimental footprints provided in the contribu-
tion by Hatala et al. (2016) reveals similarities in the forefoot and hallux depths of
modern human and Laetoli footprints, more so than with the selected chimpanzee
52 J. McClymont and R. H. Crompton
5.00
2.50
0.00
-2.50
-5.00
5.00
2.50
0.00
-2.50
-5.00
5.00
2.50
0.00
-2.50
-5.00
p = 0.000
p = 0.000
p = 0.000
p = 0.000
p = 0.000 p = 0.453
p = 0.522 p = 0.579
p = 0.762
p = 0.000
p = 0.000
p = 0.006
p = 0.002
p = 0.003
p = 0.350
p = 0.173
Fig. 3.3 Subtraction of the H. s. sapiens mean p-images from the dataset collected in soft sandy
sediments from Bates et al. (2013) and 3 alternative set means of 5 registered prints from the
11-print dataset of the Laetoli G1 trail; (a) the rst set of ve Laetoli footprints registered and the
mean image compared with H. s. sapiens mean p-images following registration; (b) the second set
of ve Laetoli footprints registered and the mean image compared with H. s. sapiens mean p-images
following registration; (c) the third set of ve Laetoli footprints registered and the mean image,
compared with H. s. sapiens mean p-images following registration
footprint. The relatively deeper experimental toe depths observed in the human and
chimpanzee prints are likely due to the substrate effects described extensively by
Bates et al. (2013), while the human and Laetoli hallucal impression indicating toe
off is also clear. While not directly measuring dynamic behaviour, the unique case of
fossil footprint trails is a reection of dynamic behaviour that occurred at one time.
The reliability of assessment, putting aside substrate characteristics, is a major
issue in interpreting gait from footprints: very loosely, Raichlen et al. (2010) and
Crompton et al. (2012) have at best a 50% chance that their conclusions that the G1
trackmaker walked upright are correct, and similarly Bennett et al. (2016) have at
best a 50% chance that their conclusion of functional stasis between Laetoli and
Ileret is correct. Hatala et al. (2016) have at best a one-in-four chance of having
drawn a correct conclusion in claiming that the ve prints chosen shown that the G1
trackmaker walked with a more exed posture than ourselves. But, should Masao
et al. (2016)discover more extensive footprints at Laetoli S, we may well get
50 continuous steps and up to near 90% reliability (again always given a good
relationship of footprint depth with foot pressure), which would make the analysis of
gait from footprint depth much more meaningful and promising.
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 53
The reconstruction of gait from the postcrania of early hominins however will
require a different and more indirect strategy, as even in single bones, and not taking
into account the functional redundancy of distal segments, we can expect high intra-
and inter-taxon variability of trait morphology, some of which variation will not be
functional (Bock and von Wahlert 1965), since motor control patterns adapt loco-
motor behaviour step-to-step based on each interaction between the body and the
environment (Bernstein 1967; Riley and Turvey 2002). This is a primary tenet of
dynamical systems theory, which has now matured into the concept of neurobiolog-
ical degeneracy (see, e.g. Seifert et al. 2016), on which current biomechanical studies
of gait variability are now almost always based. The prediction of overall gait
patterns in early hominins such as Au. afarensis from morphology of proximal
bony elements such as long-bone shaft and femoral neck cross-sectional geometry,
as attempted by Ruff et al. (2016), is hazardous enough unless dynamic modelling is
used to assess summed forces applied to the foot.
Conclusion
Based on the lack of statistically signicant differences between Laetoli and unshod
modern humans from Namibian footprint trails (Fig. 3.1e), and considering the
commonality of stride-to-stride and step-to-step uctuations in both trails
(Fig. 3.1a, b), we nd no evidence from this analysis that indicates the Laetoli
trackmaker utilized a exed knee posture beyond the range of variation in modern
humans today (given the sediment characteristics and small sample of footprints).
This supports previous ndings for footprint analyses (Sellers et al. 2005; Raichlen
et al. 2010; Crompton et al. 2012) and is consonant with studies showing that
Au. afarensis was biomechanically capable of, and therefore likely to have
performed, erect bipedality. It is possible that some or most australopith populations
engaged in substantial arboreality, as suggested for AL-288-1 Lucy (Ruff et al.
2016), based on cross-sectional geometry of her long bones and femoral neck. It
does not however follow that selection for arboreal activity reduced effectiveness in
terrestrial bipedalism in australopiths. We have shown that footprints in the habitu-
ally unshod Holocene Namibian population and in the maker of Laetoli G1 could be
very similar.
54 J. McClymont and R. H. Crompton
The primary theoretical argument underpinning our conclusions is that this is
possible due to high degrees of variability expected in all aspects of morphology and
locomotor behaviour across all biological species, and quantiable by functional
variability during movement (Bruijn et al. 2013). However, as is typical of paleon-
tology, we are restricted by small sample sizes, and hence unable to capture the full
biomechanical variation in movement which would have been present in the
trackmaker at the time footprints were made. Thus, without the inclusion of, or
reference to, the known variability in fossil populations, and the functional variabil-
ity in locomotion in analogous, extant non-human ape populations, interpretations
should only be made for the trackmaker and not used to predict species level
behavior, or to suggest unique locomotor modes at he species level. Variability in
morphology and behavior, and ontogenetic plasticity, although challenging for our
understanding of human and non-human ape fossils, should equally be seen as key to
our success in dealing with environmental change and expanding into a very wide
range of new environments. Plasticity, in a real sense, is key to evolutionary
biological success of all species.
Acknowledgements We dedicate this contribution to Mr. Russell Savage, our dear friend and
colleague, who contributed extensively to this work before his sudden and untimely death in late
2016.
We thank Drs. Karl Bates and Todd Pataky for their contributions to core analysis of footprint
depth and pressure. We thank Dr. Craig Feibel for his advice on the taphonomy of Laetoli G1, and
the extent and identity of usable G1 prints, and Prof. Ronald Clarke for related advice from his
perspective as one of the chief excavators of Laetoli G1. We thank our senior colleague Prof.
Michael Day for his kindness in entrusting his photogrammetric records of G1 and G2/3, made
shortly after excavation and prior to subaerial erosion, to RHC. We thank Prof. Matthew Bennett for
giving us access to his Namibia and Ileret data. The participation of JMcC in both excavations at
Ileret and Walvis Bay, Namibia, has enabled us to work on these scans with condence. The
research was funded by the Leverhulme Trust, the UK Natural Environment Research Council and
the Institute of Ageing and Chronic Disease at the University of Liverpool.
Supplementary Material
Detailed Materials and Methods
Ileret prints used in this analysis are from the upper footprint surface at FwJj14E
(4
1804400 N; 36
1601600 E) and include ve prints from the longest trail (FUT1-1,
FUT1-3, FUT1-5, FUT1-6, FUT1-7), two prints from a shorter trail, (FUT3-1,
FUT3-2) and four individual prints (FUI1, FUI2, FUI6, FUI7). They were imprinted
in ne-grained tuffaceous silt and ne sand deposited as overbank ood deposits and
assigned to Homo erectus on the basis of biometric inferences of body mass and
stature (Bennett et al. 2009). The Laetoli prints (Leakey and Harris 1987) (Trail G1)
used here are scans of rst-generation casts of the Laetoli G1 prints at the National
Museum of Kenya, laser-scanned using a Konica Minolta VI900 with a vertical
resolution of 90 μm. Access to Days photogrammetric data provided a vital check
on print morphology. Prints G1/28 and G1/30 were omitted due to excessive erosion
and vegetation damaged and also G1/38 as the posterior heel imprint is missing
through faulting (Crompton et al. 2012). On the edge of the Namib Sand Sea (Walvis
Bay, Namibia), unshod footprints from the Holocene (11,500 ka) (Kinahan 1996;
Morse et al. 2013) occur on silt surfaces, deposited as overbank ood deposits from
the Kuiseb River and exposed between sand dunes (23
0002500 S; 14
2902600 E).
These prints were excavated in 2010 and scanned using a Konica Minolta VI900
(Morse et al. 2013). The print makers are assumed to have been habitually unshod
due to their African context and date, as well as the presence of skin (callus) texture
visible in the footprints (Kinahan 1996; Morse et al. 2013). If footwear was worn on
occasions, it is unlikely to have been laterally constrictive. Optical laser scans of
100 prints from 10 living, Western individuals, made in a laboratory tray lled with
ne, moist sand, were recorded using an LDI PS-400, and the 10 subject means
combined into an overall modern human mean (Crompton et al. 2012). Photographs
and stereopairs of individual Laetoli prints are available in the Laetoli monograph
(Leakey and Harris 1987), scans of both the Laetoli and Ileret prints have been
previously published (Bennett et al. 2009; Crompton et al. 2012), and the prints from
Namibia have also been well documented (Kinahan 1996; Morse et al. 2013).
Consequently the replication of individual print images here would be redundant.
Many are freely available online via Bennetts Bournemouth University website.
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 55
All footprint scans were rectied to the orthogonal plane and cropped so that only
the plantar surface of each footprint was retained (Fig. 3.4). After removal of any
additional surrounding sediment, the data was imported as XYZ point clouds into
Matlab and processed using Liverpools in-house software pedobarographic statis-
tical parametric mapping (pSPM) (Pataky and Goulermas 2008; Pataky et al. 2008).
This software was designed to compute measures of central tendency across multiple
foot pressure images (Friston et al. 1995; Pataky and Goulermas 2008); however, by
substituting pressure for depth, it has here been applied to footprint trails (Crompton
et al. 2012). This substitution does not imply that we believe the relationship
between foot pressure and footprint depth to be linear, permitting a direct and simple
(yet biomechanically incorrect) interpretation of gait from footprint depth. Never-
theless, a natural relationship must exist given p¼F
A, where pressure ( p) is the
amount of force (F) acting per unit area (A). Following this nomological premise, we
trust this analogue to more robustly underpin interpretation from statistical compar-
isons and inferences on multiple records.
The pSPM software co-registers the entire plantar surface of a sample of foot-
prints such that each pixel (footprint depth) corresponds to the equal anatomical
location in all co-registered images. To achieve standardized comparisons, all point
clouds were down-sampled into images of 1 mm
2
pixel dimensions. To enable
standardized comparison of footprints of different absolute depths, each image was
normalized by its own maximum depth such that pixel values ranged 01, with
0 corresponding to shallowest depth and 1 the point of maximum depth of the
footprint as in our previous study (Bates et al. 2013). Registration of images within
pSPM can be undertaken using a number of automated algorithms or through
56 J. McClymont and R. H. Crompton
Fig. 3.4 Diagrammatic explanation of the data processing carried out prior to registration and
topological statistical analysis; (a) surfaced laser scan of modern human footprint; (b) ten points
were selected on the under-formed surface surrounding the print (i.e. outside any displacement rims,
fractured areas, etc.). A plane was subsequently tted through these points, and (c) the rotation
required to align this plane with the horizontal was applied to the footprint, thereby aligning print
depth with the vertical axis; (de) the same horizontal plane was then lowered until it reached the
highest topological point (i.e. shallowest depth) on the plantar surface of the footprint. All pixels
above this plane were then cropped out leaving only the plantar surface (occasionally small
surrounding areas of sediment were manually removed). Depth normalization was then carried
out using the range of depths present across the plantar surface, culminated with a scaled depth
range of 01, with the shallowest point (within the midfoot in e) having a value of 0 and the deepest
point (in the hallux in e) having a value of 1
manual manipulation that involves the rotation and scaling of individual images to a
common template image (Pataky et al. 2008). A previous study tested the accuracy
and repeatability of manual registration and showed that it produces comparable and
in some cases better results than various registration algorithms (Pataky et al. 2008).
Where a higher level of divergence in topology occurs, such as with inter-species
comparisons, manual registration has been found to give better results (Crompton
et al. 2012).
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 57
In this analysis topological variation required that the 9 Ileret prints were man-
ually registered to each other, as were the 11 prints used from the G1 trail (Crompton
et al. 2012). The Walvis Bay and modern human footprints were internally registered
using an automated algorithm that minimized the root mean square error of pixels
globally across pressure images (Pataky and Goulermas 2008). Registrations
between populations of prints, facilitating cross-site (i.e. cross-species) comparisons
were all performed manually. Here, all manual registrations were repeated three
times to observe any impact of operator subjectivity on subsequent statistical tests.
Once registered, measures of central tendency can then be calculated to create
statistical parametric maps (SPMs) and compared pixel by pixel using pairwise
t-tests. Statistical comparison between print populations (i.e. different trails) is
possible since probability values are available for every pixel in the SPM. Pixel-
wise two-sample t-tests can be used to create a statistical image known as an SPM
{t}(Pataky and Goulermas 2008; Pataky et al. 2008; Crompton et al. 2012) that
provides a statistical comparison between two print populations. The large pixel
numbers pose a potential problem since large t values (e.g. t > 3) are likely to occur
simply by chance, and in a footprint or plantar pressure (which are the product of
interaction between two continuous media) neighbouring pixels are clearly not
independent.
However, neighbouring pixels tend to behave in a similar way due to the smooth
outline, or boarder, of a print, and their t values form a generally smooth SPM, which
can be shown to be topologically characteristic of a thresholded SPM (e.g. cluster
size, number of clusters, etc.). Specically, random eld theory (RFT) is used to
determine the t-threshold at which alpha ¼5% of the pixels would be expected to
reach, simply by chance, based on the smoothness and on the foot shape which is
parameterized by pixel connectivity with the plantar surface. Shape information is
necessary because a square eld, for example, would be expected to produce fewer
suprathreshold clusters than would a long, narrow rectangular eld of the same area
and same smoothness. The SPM is then thresholded based on this critical t value, and
one is left with some suprathreshold clusters of pixels that have survived the
threshold. RFT then uses analytical probability density functions to compute the
likelihood that clusters of the given size could have been produced by chance
(Friston et al. 1995; Pataky and Goulermas 2008).
Figure 3.1 presents the mean footprints and the results of statistical comparisons
of 9 prints from the upper surface at Ileret (FwJj14E; Kenya) (Bennett et al. 2009),
with 11 prints from the G1 Trail at Laetoli (Leakey and Hay 1979; Leakey and Harris
1987), a 32-print sample from the Holocene trail at Walvis Bay, Namibia (Morse
et al. 2013), and the modern Western, habitually shoe-wearing dataset (N ¼100 foot-
prints from 10 individuals). The rst two images in each set are the site means, the
third is their subtraction to show where they differ, while the fourth identies those
areas where the difference is statistically different using pixel-level pairwise t-tests
(Pataky and Goulermas 2008; Pataky et al. 2008) after normalization by plantar
surface maximum depths and their probability level These tests were carried out
using the same methods used in previous studies (Crompton et al. 2012; Bates et al.
2013).
58 J. McClymont and R. H. Crompton
Additional Footprint Discussion
Comparison of footprint topology between sites with different substrates and geo-
logical properties is potentially difcult since the biomechanical signature of a
trackmaker is mediated through the geotechnical properties of the substrate, and
the substrate may also inuence taphonomic modication (Craig 1997; Ditcheld
Fig. 3.5 An example of within-subject registration using ten prints from the modern, Western human
dataset; (a) the ten prints are preprocessed as explained in Fig. 3.4. An initial registration is then
performed that individually (i.e. one at a time) aligns the last nine prints with the rst print in the
dataset (not depicted above); (b) subsequently a second registration is performed in which all ten
prints are individually (i.e. one at a time) aligned with their mean image; (c) the mean image itself. For
the modern Western human and Namibian prints, this was carried out using automated algorithms
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 59
Fig. 3.6 Depiction of the same registration process shown in Fig. 3.5 for the Laetoli prints, in
which registration was carried out manually (operator rotation and scaling of images) rather than
using automated algorithms. The same manual registration was necessary on the Ileret prints
and Harrison 2011). In this analysis, at a macroscale we compared print populations
from three natural environments, two from silt-rich ood/overbank deposits (Walvis
Bay, Namibia and Ileret) (Crompton et al. 2012; Morse et al. 2013) and one from
volcanic ash deposited via air-fall at Laetoli (Leakey and Harris 1987), with a sample
of modern prints collected from ne sand in the laboratory. At the microscale,
variation also exists within each depositional environment, dependent on local
variations in grain size, moisture content, vertical stratigraphy and, signicantly
(especially the case at Laetoli), the degree of turbation by animal trampling
(Morse et al. 2013). Substrate affects are particularly obvious in the Ileret prints,
whereby withdrawal of the heel from soft, wet substrates causes side wall suction,
naturally decreasing the macro-shape, specically the width of the print (Craig 1997;
Bennett et al. 2009; Morse et al. 2013). The enhanced longitudinal asymmetry
deeper forefoot (MTH1-3) than the heel is also a feature of a softer substrate and is
a visible feature in the mean Ileret print.
Technically, the substrate rst holds the weight of the individual during the rst
phase of stance, only to fail further during the second phase associated with higher
plantar pressures during toe off. The lack of clarity of toe impressions is a feature of
deeper prints where foot withdrawal often modies the impressions left by phalanges
(Crompton et al. 2012). This is particularly evident at Ileret where toe drag is clear,
associated with higher forces required to pull the toes out of deeper substrate. The
medial longitudinal arch is also modied in softer substrates by the proximal
movement of sediment under rotation of the ball of the foot, potentially producing
a tendency towards a atter arch in deeper prints.
60 J. McClymont and R. H. Crompton
Fig. 3.7 Non-linear registration of (a) modern human; (b) Ileret; (c) Laetoli; (d) Holocene human
means registered to the Namibian mean
Fig. 3.8 Stage after registration to the reference mean, here Namibia, when individual prints are
re-registered to the non-linear mean templates and the means regenerated
However, the methodology used in this analysis helps mitigate these inuences.
Principally, we are able to compare whole footprint populations on the basis of
measures of central tendency rather than by comparing individual prints, which may
show strong individual substrate inuences (Leakey and Harris 1987; Bennett et al.
2016; Morse et al. 2013). For cross-site comparisons it subsequently becomes
important that the range of sedimentological properties exhibit overlap (i.e. in
terms of their geomechanical strength), thereby isolating biological (anatomy and
gait) similarities and differences that impact on footprint form. It is important to note
that these sedimentological conditions may not directly or obviously translate into
sediment characteristics that are easily measurable in the geological record, such as
average grain size, sorting or composition. Broadly similar geomechanical properties
(e.g. bearing capacity, Poisson ratio, etc.) may be produced by different combina-
tions of physical sediment characteristics (Craig 1997). There is no doubt that further
experimental work is needed to explore the inuence of sedimentology on footprint
form (and the range of variables that dene a sediments rheology). However, we
suggest that in the absence of this experimental work, and a detailed mechanistic
understanding, it is perhaps most appropriate to ensure comparisons are made on
prints of overlapping depths since depth does appear to correlate with substrate
strength (Bates et al. 2013; Morse et al. 2013).
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 61
References
Alexander, R. M. N. (1984). Stride length and speed for adults, children, and fossil hominids.
American Journal of Physical Anthropology, 63,227.
Alexander, R. M. N. (2003). Principles of animal locomotion. Princeton: Princeton University
Press.
Arts, M., & Bus, S. (2011). Twelve steps per foot are recommended for valid and reliable in-shoe
plantar pressure data in neuropathic diabetic patients wearing custom made footwear. Clinical
biomechanics, 26, 880884.
Bartlett, R., Wheat, J., & Robins, M. (2007). Is movement variability important for sports
biomechanists? Sports Biomechanics, 6(2), 224243.
Bates, K.T., Savage, R., Pataky, T. C., Morse, S. A., Webster, E., Falkingham, P. L., Ren,L.
Qian, Z, Collins, D., Bennett, M.R. McClymont, J., Crompton R.H. (2013). Does footprint
depth correlate with foot motion and pressure? Journal of the Royal Society Inteface, 10.https://
doi.org/10.1098/rsif.2013.0009.
Bennett, M. R., Harris, J. W. K., Richmond, B. G., Braun, D. R., Mbua, E., Kiura, P., Olago, D.,
Kibunjia, M., Omuombo, C., Behrensmeyer, A. K., Huddart, D., & Gonzalez, S. (2009). Early
hominin foot morphology based on 1.5-million-year-old footprints from Ileret, Kenya. Science,
323(5918), 11971201. https://doi.org/10.1126/science.1168132.
Bennett, M. R., Reynolds, S. C., Morse, S. A., & Budka, M. (2016). Footprints and human
evolution: Homeostasis in foot function? Palaeogeography, Palaeoclimatology,
Palaeoecology.https://doi.org/10.1016/j.palaeo.2016.08.026.
Bernstein, N. (1967). The co-ordination and regulation of movements. Oxford: Pergamon.
Bock, W. (1965). The role of adaptive mechanisms in the origin of higher levels of organisation.
Systematic Zoology, 14(4), 272287.
Bock, W. (1994). Concepts and methods in ecomorphology. Journal of Biosciences, 19, 403413.
Bock, W., & von Wahlert, G. (1965). Adaptation and the form-function complex. Evolution, 19(3),
269.
Bruijn, S. M., Meijer, O. G., Beek, P. J., & Van Dieën, J. H. (2013). Assessing the stability of
human locomotion: A review of current measures. Journal of the Royal Society Interface,
10(83), 20120999.
62 J. McClymont and R. H. Crompton
Carey, T. S., & Crompton, R. H. (2005). The metabolic costs of bent-hip, bent kneewalking in
humans. Journal of Human Evolution, 48,2544.
Clarke, R. J. (2019). Excavation, reconstruction and taphonomy of the StW 573 Australopithecus
prometheus skeleton from Sterkfontein Caves, South Africa. Journal of Human Evolution,
127, 41.
Craig, R. (1997). Soil mechanics. London: Chapman & Hall.
Crompton, R. H., Li, Y., Wang, W., Günther, M. M., & Savage, R. (1998). The mechanical
effectiveness of erect and bent-hip, bent-kneebipedal walking in Australopithecus afarensis.
Journal of Human Evolution, 35,5574.
Crompton, R. H., Thorpe, S. K. S., Wang, W., Li, Y., Payne, R., Savage, R., Carey, T., Aerts, P.,
Van Elsacker, L., Hofstetter, A., Günther, M., & Richardson, R. (2003). The biomechanical
evolution of erect bipedality. Courier Forschuungs-Institut Senckenberg, 243, 115126.
Crompton, R. H., Vereecke, E. E., & Thorpe, S. (2008). Locomotion and posture from the common
hominoid ancestor to fully modern hominins, with special reference to the last common panin/
hominin ancestor. Journal of Anatomy, 212(4), 501543.
Crompton, R. H., Pataky, T. C., Savage, R., DAout, K., Bennett, M. R., Day, M. H., Bates, K.,
Morse, S., & Sellers, W. I. (2012). Human-like external function of the foot, and fully upright
gait, conrmed in the 3.66-million-year-old Laetoli hominin footprints by topographic statistics,
experimental footprint- formation and computer simulation. Journal of the Royal Society
Interface, 9, 707719.
Crompton, R. H., McClymont, J., Thorpe, S., Sellers, S., Heaton, S., Pickering, T. R., Pataky, T.,
Stratford, D., Carlson, K., Jashashvili, T., Beaudet, A., Bruxelles, L., Goh, C., Kuman, K., &
Clarke, R. (2018). Functional anatomy, biomechanical performance capabilities and potential
Niche of StW 573: An Australopithecus Skeleton (circa 3.67 Ma) From Sterkfontein Member
2, 3 and its signicance for The Last Common Ancestor of the African Apes and for Hominin
Origins. Bioxriv.https://doi.org/10.1101/481556.
Davids, K., Glazier, P., Araujo, D., & Bartlett, R. (2003). Movement systems as dynamical systems.
Sports Medicine, 33(4), 245260.
Day, M. H., & Napier, J. R. (1964). Fossil foot bones. Nature, 201, 969970.
Day, M. H., & Wickens, E. H. (1980). Laetoli Pliocene hominid footprints and bipedalism. Nature,
286, 385387.
Dingwell, J., John, J., & Cusumano, J. (2010). Do humans optimally exploit redundancy to control
step variability in walking? PLoS Computational Biology, 6(7). https://doi.org/10.1371/journal.
pcbi.1000856.
Ditcheld, P., & Harrison, T. (2011). Sedimentology, lithostratigraphy and depositional history of
the Laetoli area. In T. Harrison (Ed.), Paleontology and geology of Laetoli: Human evolution in
context (Geology, geochronology, paleoecology and paleoenvironment, vertebrate paleobiology
and paleoanthropology) (Vol. 1, pp. 4776). Dordrecht: Springer.
Dunn, R. H., Tocheri, M. W., Orr, C. M., & Jungers, W. L. (2014). Ecological divergence and talar
morphology in gorillas. American Journal of Physical Anthropology, 153(4), 526541.
Edelman, G. M., & Gally, J. A. (2001). Degeneracy and complexity in biological systems. Pro-
ceedings of the National Academy of Sciences, 98(24), 1376313768.
Ferris, D., Louie, M., & Farley, C. (1998). Running in the real world: Adjusting leg stiffness for
different surfaces. Proceedings of the Royal Society of London Series B: Biological Sciences,
265(1400), 989994.
Friston, K. J., Holmes, A. P., Worsley, K. J., Poline, J. P., Frith, C. D., & Frackowiak, R. S. (1995).
Statistical parametric maps in functional imaging: A general linear approach. Human Brain
Mapping, 3, 165189.
Giancoli, D. (2004). Physics: Principles with applications. Upper Saddle River: Pearson Education.
Harcourt-Smith, W. E. H., & Aiello, L. C. (2004). Fossils, feet and the evolution of human bipedal
locomotion. Journal of Anatomy, 204, 403416.
Hatala, K., Demes, B., & Richmond, B. (2016). Laetoli footprints reveal bipedal gait biomechanics
different from those of modern humans and chimpanzees. Proceedings of the Royal Society
Series B, 283(1836), 20160235.
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 63
Heaton, J. L., Pickering, T. R., Beaudet, A., Clarke, R. J., Bruxelles, L., Carlson, K. J., Crompton,
R., Stratford, D. T., Kuman, K., McClymont, J., & Jashashvili, T. (2019). The long limb bones
of the StW 573 Australopithecus skeleton from Sterkfontein Member 2: Descriptions and
proportions. Journal of Human Evolution, 133, 167197.
Jungers, W. L., Harcourt-Smith, W. E. H., Wunderlich, R. E., Tocheri, M. W., Larson, S. G.,
Sutikna, T., Due, R. A., & Morwood, M. J. (2009). The foot of Homo oresiensis.Nature, 459,
8184. https://doi.org/10.1038/nature07989.
Kang, H. G., & Dingwell, J. B. (2008). Effects of walking speed, strength and range of motion on
gait stability in healthy older adults. Journal of Biomechanics, 41(14), 28992905.
Kidd, R. S., OHiggins, P. O., & Oxnard, C. E. (1996). The OH8 foot: A reappraisal of the hindfoot
utilizing a multivariate analysis. Journal of Human Evolution, 31, 269291.
Kinahan, J. (1996). Human and domestic animal tracks in an archaeological lagoon deposit on the
coast of Namibia. South African Archaeological Bulletin, 51,9498.
Kramer, P. A. (1999). Modeling the locomotor energetics of extinct hominids. Journal of Exper-
imental Biology, 202, 28072818.
Kramer, P. A., & Eck, G. G. (2000). Locomotor energetics and leg length in hominid bipedality.
Journal of Human Evolution, 38, 651666.
Laland, K. N., Uller, T., Feldman, M. W., Sterelny, K., Müller, G. B., Moczek, A., Jablonka, E., &
Odling-Smee, J. (2015). The extended evolutionary synthesis: Its structure, assumptions and
predictions. Proceedings of the Royal Society B: Biological Sciences, 282(1813), 20151019.
Latash, M., Scholz, J., & Schöner, G. (2002). Motor control strategies revealed in the structure of
motor variability. Exercise and Sports Science Review, 30,2631.
Latimer, B. (1991). Locomotor adaptations in Australopithecus afarensis: The issue of arboreality.
In B. Senut & Y. Coppens (Eds.), Origine(s) de la bipédie chez les Hominidés (pp. 169176).
Paris: CNRS.
Latimer, B., & Lovejoy, C. O. (1989). The calcaneus of Australopithecus afarensis and its
implications for the evolution of bipedality. American Journal of Physical Anthropology, 78,
369386.
Latimer, B., Ohman, J. C., & Lovejoy, C. O. (1987). Talocrural joint in African hominoids:
Implications for Australopithecus afarensis.American Journal of Physical Anthropology, 74,
155175.
Leakey, M., & Harris, J. (1987). Laetoli: A Pliocene site in northern Tanzania. Oxford: Clarendon
Press.
Leakey, M., & Hay, R. (1979). Pliocene footprints in the Laetoli beds at Laetoli, northern Tanzania.
Nature, 278(5702), 317323.
Lovejoy, C. O., Latimer, B. M., Spurlock, L., & Haile-Selaissie, Y. (2016). The pelvic girdle and
limb bones of KSD/VP/1-1. Ethiopia. In Y. Haile-Selasssie & D. F. Su (Eds.), The postcranial
anatomy of Australopithecus afarensis: 1051 New Insights from KSD-VP-1/1 (pp. 155177).
Dordrecht: Springer.
Masao, F. T., Ichumbaki, E. B., Cherin, M., Barili, A., Boschian, G., Iurino, D. A., Menconero, S.,
Moggi-Cecchi, J., & Manzi, G. (2016). New footprints from Laetoli (Tanzania) provide
evidence for marked body size variation in early hominins. eLife, 5, e19568. https://doi.org/
10.7554/eLife.19568.
McClymont, J. (2017). Foot pressure variability and locomotor plasticity in hominins (Doctoral
dissertation, University of Liverpool).
McClymont, J., Pataky, T. C., Crompton, R. H., Savage, R., & Bates, K. T. (2016). The nature of
functional variability in plantar pressure during a range of controlled walking speeds. Royal
Society Open Science, 3(8), 160369.
McHenry, H. M. (1975). Biomechanical interpretation of the early hominid hip. Journal of Human
Evolution, 4(5), 343355.
McHenry, H. M. (1986). Size variation in the postcranium of Australopithecus afarensis and extant
species of Hominoidea. Human Evolution, 1(2), 149155.
64 J. McClymont and R. H. Crompton
Meldrum, D. J., Lockley, M. G., Lucas, S. G., & Musiba, C. (2011). Ichnotaxonomy of the Laetoli
trackways: The earliest hominin footprints. Journal of African Earth Sciences, 60(12), 112.
Morse, S. A., Bennett, M. R., Liutkus-Pierce, C., Thackeray, F., McClymont, J., Savage, R., &
Crompton, R. H. (2013). Holocene footprints in Namibia: The inuence of substrate on footprint
variability. American Journal of Physical Anthropology, 151(2), 265279.
Nagano, A., Umberger, B. R., Marzke, M. W., & Gerritsen, K. G. M. (2005). Neuromusculoskeletal
computer modeling and simulation of upright, straight-legged, bipedal locomotion of
Australopithecus afarensis (A.L. 288-1). American Journal of Physical Anthropology, 126,
213.
Owings, T., & Grabiner, M. (2003). Measuring step kinematic variability on an instrumented
treadmill: How many steps are enough? Journal of Biomechanics, 36, 12151218.
Pataky, T. (2010). Generalized n-dimensional biomechanical eld analysis using statistical para-
metric mapping. Journal of Biomechanics, 43(10), 19761982.
Pataky, T., & Goulermas, J. (2008). Pedobarographic statistical parametric mapping (pSPM): A
pixel-level approach to plantar pressures image analysis. Journal of Biomechanics, 41,
21362143.
Pataky, T. C., Caravaggi, P., Savage, R., Parker, D., Goulermas, J. Y., Sellers, W. I., & Crompton,
R. H. (2008). New insights into the plantar pressure correlates of walking speed using
pedobarographic statistical parametric mapping (pSPM). Journal of Biomechanics, 41(9),
19871994.
Pataky, T. C., Bosch, K., Mu, T., Keijsers, N. L., Segers, V., Rosenbaum, D., & Goulermas, J. Y.
(2011). An anatomically unbiased foot template for inter-subject plantar pressure evaluation.
Gait & Posture, 33(3), 418422.
Pataky, T. C., Savage, R., Bates, K. T., Sellers, W. I., & Crompton, R. H. (2013). Short-term step-
to-step correlation in plantar pressure distributions during treadmill walking, and implications
for footprint trail analysis. Gait & Posture, 38(4), 10541057.
Raichlen, D. A., & Gordon, A. D. (2017). Interpretation of footprints from site S conrms human-
like bipedal biomechanics in Laetoli hominins. Journal of Human Evolution, 107, 134138.
Raichlen, D. A., Gordon, A. D., Harcourt-Smith, W. E., Foster, A. D. H., Jr., & W.R. (2010).
Laetoli footprints preserve earliest direct evidence of human-like bipedal biomechanics. PLoS
One, 5(3), e9769.
Reynolds, T. (1987). Stride length and its determinants in humans, early hominids, primates, and
mammals. American Journal of Physical Anthropology, 72, 101115.
Riley, M., & Turvey, M. (2002). Variability and determinism in motor behavior. Journal of Motor
Behaviour, 34,99125.
Ruff, C. B., Burgess, M. L., Ketcham, R. A., & Kappelman, J. (2016). Limb bone structural
proportions and locomotor behavior in A.L. 288-1 (Lucy). PLoS ONE, 11(11), e0166095.
https://doi.org/10.1371/journal.pone.0166095.
Seifert, L., Komar, J., Araújo, D., & Davids, K. (2016). Neurobiological degeneracy: A key
property for functional adaptations of perception and action to constraints. Neuroscience &
Biobehavioral Reviews, 69, 159165. https://doi.org/10.1016/j.neubiorev.2016.08.006.
Sellers, W. I., Cain, G., Wang, W. J., & Crompton, R. H. (2005). Stride lengths, speed and energy
costs in walking of Australopithecus afarensis: Using evolutionary robotics to predict locomo-
tion of early human ancestors. Royal Society Interface, 2, 431442.
Stern, J. T., & Susman, R. L. (1983). The locomotor anatomy of Australopithecus afarensis.
American Journal of Physical Anthropology, 60, 279317.
Stern, J. T., & Susman, R. L. (1991). Total morphological patternversus the magic trait:
Conicting approaches to the study of early hominid bipedalism. In B. Senut & Y. Coppens
(Eds.), Origine(s) de la bipédie chez les Hominidés (pp. 99112). Paris: CNRS.
Stern, J. T., Susman, R. L., & Jungers, W. (1984). Arboreality and bipedality in the Hadar hominids.
Folia Primatologica, 43, 113156.
Stolwijk, N. M., Duysens, J., Louwerens, J. W. K., van de Ven, Y. H., & Keijsers, N. L. (2013). Flat
feet, happy feet? Comparison of the dynamic plantar pressure distribution and static medial foot
geometry between Malawian and Dutch adults. PloS One, 8(2), e57209.
Thelen, E. (1995). Motor development: A new synthesis. American Psychologist, 50(2), 79.
3 Repetition Without Repetition: A Comparison of the Laetoli G1, Ileret, Namibian... 65
Thelen, E. (2005). Dynamic systems theory and the complexity of change. Psychoanalytic Dia-
logues, 15(2), 255283.
Tudor-Locke, C., Schuna, J. M., Jr., Han, H. O., Aguiar, E. J., Green, M. A., Busa, M. A., Larrivee,
S., & Johnson, W. D. (2017). Step-based physical activity metrics and cardiometabolic risk:
NHANES 2005-06. Medicine and Science in Sports and Exercise, 49(2), 283.
Venkataraman, V. V., Kraft, T. S., & Dominy, N. J. (2013a). Tree climbing and human evolution.
Proceedings of the National Academy of Sciences, 110, 12371242.
Venkataraman, V. V., Kraft, T. S., & Dominy, N. J. (2013b). Phenotypic plasticity of climbing-
related traits in the ankle joint of great apes and rainforest hunter-gatherers. Human Biology, 85,
309328.
Wainwright, P. C. (1991). Ecomorphology: Experimental functional anatomy for ecological prob-
lems. American Zoologist, 31(4), 680693.
Wainwright, P., Bellwood, D., & Westneat, M. (2002). Ecomorphology of locomotion in labrid
shes. Environmental Biology of Fishes, 65(1), 4762.
Ward, C. V., Kimbel, W. H., & Johanson, D. C. (2011). Complete fourth metatarsal and arches in
the foot of Australopithecus afarensis.Science, 331, 750753.
Watson, J., Payne, R., Chamberlain, A., Jones, R., & Sellers, W. I. (2009). The kinematics of load
carrying in humans and great apes: Implications for the evolution of human bipedalism. Folia
Primtalogica, 80(5), 309328.
White, T. D. (1980). Evolutionary implications of Pliocene hominid footprints. Science, 208,
175176.
White, T. D., & Suwa, G. (1987). Hominid footprints at Laetoli: Facts and interpretations. American
Journal of Physical Anthropology, 72, 485514.
Wolf, P., Stacoff, A., Liu, A., Nester, C., Arndt, A., Lundberg, A., & Stuessi, E. (2008). Functional
units of the human foot. Gait & Posture, 28(3), 434441.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Chapter 4
Reproduce to Understand: Experimental
Approach Based on Footprints in Cussac
Cave (Southwestern France)
Lysianna Ledoux, Gilles Berillon, Nathalie Fourment, and Jacques Jaubert
Abstract The morphology of a track depends on many factors that must be con-
sidered when interpreting it. An experimental approach is often required to under-
stand the inuence of each of these factors, both at the time of the track formation
and after its formation. These aspects, which are fairly well documented for tracks
found in open-air settings, are much more limited for those found in karst settings.
Although caves are stable environments enabling the preservation of archaeological
remains, many taphonomical processes can alter the grounds and the walls. Based on
the observations made on footprints found in Cussac Cave (Dordogne region of
southwestern France), this study focuses on one of these natural phenomena and
tests the impact of ooding episodes and the resulting clay deposits on the tracks
morphology and topography. Our experiments show that although the general
morphology of footprints and some details such as digits are preserved, their
topography is altered by successive ooding episodes and clay deposits. The loss
of denition of the footprints due to ooding episodes can also lead to misinterpre-
tation. This work sheds new light on the Cussac footprints, while the further
development of such experiments will allow us to improve our results and apply
them to other settings and sites.
Keywords Taphonomy · Footprint · Cave · Palaeolithic · Experimentation
L. Ledoux (*) · J. Jaubert
Université de Bordeaux, Pessac, France
e-mail: ledouxlysianna@gmail.com
G. Berillon
UMR7194 MNHNCNRS/Département Homme et Environnement, Musée de lHomme, Palais
de Chaillot, Paris, France
N. Fourment
SRA, DRAC Nouvelle Aquitaine 54 rue Magendie, Bordeaux Cedex, France
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_4
67
Introduction
Tracks are among the most fragile and underestimated of archaeological remains, yet
they provide a valuable source of information on site frequentation. They are the
direct representation of a particular event in an individuals life, the marker and proof
of human or non-human passage through a place. They offer an essential insight into
the biology, locomotion, behaviour or activities of trackmakers. In the absence of
other remains, tracks may even be the only elements enabling an exploration of the
frequentation of a site.
68 L. Ledoux et al.
Over the last few decades, and recently, palaeoichnological studies have been
regularly conducted in open-air settings and have beneted from the development of
new recording and analysis techniques (Mietto et al. 2003; Berge et al. 2006; Webb
et al. 2006; Aramayo 2009; Bennett et al. 2009; Raichlen et al. 2010; Felstead et al.
2014; Ashton et al. 2014; Burns 2014; Bennett et al. 2016; Masao et al. 2016;
Panerello et al. 2017; Wiseman and De Groote 2018; Altamura et al. 2018; McLaren
et al. 2018; Smith et al. 2019; Moreno et al. 2019). In cave settings, they reached
their peak between the 1970s and the early 2000s but were less developed (Duday
and Garcia 1985,1986; Garcia 1986). Human and non-human tracks have been
documented and studied in the caves of Niaux (Clottes and Simonnet 1972; Pales
1976; Garcia et al. 1990), Pech Merle (Duday and Garcia 1983), Foissac (Garcia and
Duday 1983), Aldène (Ambert et al. 2000; Ambert et al. 2001) and Chauvet Pont
dArc (Garcia 2001,2005). Recently, interest in ichnology in the karst setting has
re-emerged among prehistorians, who have resumed the study of tracks in several
ornated caves such as the Tuc dAudoubert (Bégouën et al. 2009; Pastoors et al.
2015; see Pastoors et al. Chap. 13), Pech Merle (Pastoors et al. 2017), Aldène
(Pastoors et al. 2015; see Galant et al. Chap. 15), Cussac (Ledoux et al. 2017;
Ledoux 2019), Fontanet (Pastoors et al. 2015; Ledoux 2019), Bàsura (Citton et al.
2017;Romano et al. 2019; see Avanzini et al. Chap. 14) and Ojo Guareña (Ortega
Martinez et al. 2014; see Ortega et al. Chap. 17).
Given the variety of factors that are likely to have impacted the morphology of
hominin tracks (from the biology of the trackmakers to the nature of the substrate
and taphonomic agents), experimental approaches have been developed, especially
over the last decade, inspired by the work done on non-hominin tracks (Sollas 1879;
Brand 1996; Gatesy 2003; Milàn and Bromley 2007).
The rst studies were those conducted by Léon Pales, who observed variations in
the footprints of the same trackmaker according to the sediment and the foot
dynamic (Pales 1976). These works related to karst were pioneering and have no
equivalent in this type of setting. Subsequent experiments focused on the footprints
of early hominins in open-air settings and were developed within comparative and
functional perspectives. Since the year 2000, an increasing number of experimental
works have been conducted in response to the development of new tools (pressure
pad for the recording of plantar pressure and 3D surface recording techniques such
photogrammetry or optical laser scanning). The properties of the formation sediment
are also central (Pataky et al. 2008a; Pataky and Goulermas 2008; Pataky et al.
2008b;DAoût et al. 2010; Crompton et al. 2012; Bennett et al. 2013; Morse et al.
2013; Hatala et al. 2013; Bennett and Morse 2014; Hatala et al. 2018; Zimmer et al.
2018). However, the potential impact of taphonomical agents remains poorly inves-
tigated and has been examined in open-air settings (Marty et al. 2009;DAoût et al.
2010; Bennett and Morse 2014; Roach et al. 2016; Panerello et al. 2017; Hatala et al.
2018; Wiseman and De Groote 2018).
4 Reproduce to Understand: Experimental Approach Based on Footprints in Cussac... 69
Althoughvaried intheir objectives and contexts, these studies demonstrate that
our interpretations of tracks require a better understanding of both the formation and
the conservation setting. Each track is unique, and the objective of the ichnological
study is to understand the factors behind this uniqueness. Despite increasing interest
in the study of tracks in caves (Ortega Martinez et al. 2014; Pastoors et al. 2015;
Pastoors et al. 2017; Citton et al. 2017; Ledoux 2019; Romano et al. 2019),
ichnology in Palaeolithic caves is still little known and the formation and conserva-
tion context of these caves poorly studied. Here we present our rst results drawn
from experiments focusing on the impact of ooding on human footprints. This
natural phenomenon has been observed in Cussac Cave (Ledoux 2019) and is
recurrent in the cave setting.
The Karst Setting
Formation
As with the open-air setting, the morphology of a track produced in cave depends on
the sediment and the trackmaker. Tracks are the result of the compression of
sediment in response to a constraint exerted by a trackmaker; the original morphol-
ogy of tracks therefore depends on both the trackmaker (locomotion, biology,
behaviour, etc.) and the formation sediment (physical and mechanical properties,
topography, etc.). Over time, this original morphology will be inuenced by various
taphonomic phenomena (erosion, bioturbation, lling, etc.). The interpretation of
tracks must therefore be based rst and foremost on a knowledge and understanding
of their setting.
However, a third parameter can inuence this morphology: the geomorphology of
the karst. The movements and behaviour of trackmakers will then be highly depen-
dent on the topography of the ground and the morphology of the walls, the height of
the ceilings and the width of the network. Consequently, the resulting tracks will
have a particular morphology whose interpretation will also depend on how well the
trackmakers perception of the cave is understood.
While tracks found in open-air settings generally belong to trackways, those
found in caves are much more varied. Testimonies of intentional or
non-intentional actions, these tracks are characterized by complete (foot, hand,
knee, etc.) or partial (ngers, toes, heels, etc.) body segments. Often associated
with wall traces (torch, colour or clay marks), they are the result of a variety of
behaviours that are inuenced either by the geometry of the cavity or by the activities
; Arias
et al.
that took place inside them (Bégouën et al. 2009; Pastoors and Weniger 2011
2011; Ledoux et al. 2017; Medina-Alcaide et al. 2018; Romano et al. 2019).
70 L. Ledoux et al.
The surface soils of a cave may have different characteristics depending on the
area (various clastic sediment deposits with different sedimentary properties, calcite
deposits, etc.). It may therefore be difcult to attribute several tracks to a single
trackmaker if they are not produced in the same area, especially where isolated tracks
are concerned.
Preservation Context
The stability of caves makes them ideal environments for the preservation of the
most fragile archaeological remains, as is very well reected in rock art. However,
despite their exceptional conservative properties, caves may be subjected to
taphonomical processes during their lifetime. These phenomena are varied and are
generally classied into two categories (Fig. 4.1):
Natural phenomena including sediment llings (various sedimentary deposits),
ooding, calcite deposits, erosion, desiccation, etc.
Non-natural phenomena including trampling, track superimpositions,
excavation, etc.
The same track may have been altered by one or more of these taphonomic
processes. Therefore, the karst setting must be understood before the tracks can be
interpreted.
Cussac Cave
Contextual Setting
Discovered in 2000 by the speleologist Marc Delluc, Cussac Cave is located south of
Périgord in Dordogne (southwestern France). It opens onto a Campanian limestone
cliff on the right bank of the Belingou, a tributary of the Dordogne River. It extends
along some 1.6 km in a single sub-horizontal gallery divided into two parts: the
Downstream Branch and the Upstream Branch (Fig. 4.2). This particularly well-
preserved cave is characterized by parietal engravings and human remains deposited
in bear hibernation nests, both associated with varied traces of human and
non-human activity (Aujoulat et al. 2001,2002,2013; Fourment et al. 2012; Jaubert
et al. 2012; Henry-Gambier et al. 2013; Jaubert 2015; Ledoux et al. 2017). All
archaeological remains (art, charcoal, human bones) are attributed to human occu-
pation in the Middle Gravettian period (2028 ka calBP) (Jaubert et al. 2017). Since
2008, a multidisciplinary team has been studying the cave in order to gain a global
understanding of the site. During the rst few years of research, a pathway was
4 Reproduce to Understand: Experimental Approach Based on Footprints in Cussac... 71
Fig. 4.1 Examples of taphonomical processes occurring in karst settings; (a) human footprint
covered by clay deposits in Cussac Cave (Dordogne, France); (b) human footprint covered by
concretion in Fontanet Cave (Ariège, France) (Ledoux et al. 2017); (c) bear manus track
transformed into rimstone in Bruniquel Cave (Tarn-et-Garonne, France)
marked throughout the cave, following the one rst taken by the discoverer. The aim
of maintaining this single pathway is to ensure optimum preservation of the cave
oors and walls.
Tracks at Cussac and Taphonomy
Although the cave is very well preserved, the current oors are not exactly the same
as they were in the Palaeolithic. Consequently, few tracks have been clearly
72 L. Ledoux et al.
Fig. 4.2 (a) Geographic location of Cussac Cave; (b) general topography of Cussac Cave and close
up of the part of the cave under study
identied as human tracks. Several factors may explain this poor preservation of the
Palaeolithic cave oors (Ledoux et al. 2017; Ledoux 2019):
4 Reproduce to Understand: Experimental Approach Based on Footprints in Cussac... 73
Geological factors: after human occupation, the cave underwent various sedi-
mentary events which signicantly damaged the oors (sedimentary deposits,
erosion, ooding, desiccation cracks, etc.).
The omnipresence of bears in the cave. Bear tracks and human tracks are
superimposed in several areas.
The restricted accessibility of some areas due to the conservation policy.
The current pathway which is, in some areas, probably the same as the
Palaeolithic pathway.
As a consequence of these various taphonomical processes, most of the complete
footprints are isolated and often altered (Fig. 4.2). Below we present the experiment
carried out on the basis of one of these taphonomical phenomena, frequently
observed in caves: the overow of the subterranean river. In some areas of Cussac
Cave, several ooding episodes occurred after human frequentation, covering tracks
with clay (Fig. 4.1a). Through a controlled experiment, we intend to test the impact
of clay deposits on the morphology and topography of footprints after ooding
episodes. Assuming that water and sediment affect the contours and the general
surface of the footprints, our purpose is therefore to follow the evolution of a
footprint from immediately after its formation to its covering by clay deposits.
Materiel and Methods
Experimental Protocol
Experimental footprints were made in a cohesive, rm and moist sediment that we
selected for its high clay content, similar to that of Cussac. It allowed for the
impression of an entire foot. This sediment was sampled from a cave in the
Dordogne region of southwestern France without any archaeological remains
(Table 4.1).
From this sediment, two types of formation surface were used: one of raw clay
with a moisture content of about 50% and one of raw clay covered with a second
level of clay that had settled after ooding (called the rst decantation) (Fig. 4.3) and
Table 4.1 Grain size analyses of the sediment sampled in Cussac Cave and in the
experimental cave
Samples
Fine sand (%)
(50063 μ)
Coarse silt (%)
(6316 )μ
Fine silt (%)
(167μ)
Clay (%)
(<7 μ)
Cussac 2.56 18.99 22.59 55.85
Experimental
cave
2.8 10.78 30.42 56
× ×
74 L. Ledoux et al.
Fig. 4.3 Experimental steps
with a moisture content varying between 60% and 70% (Fig. 4.3). Before creating
the second substrate, we tested the impact of different sediment loads (60 g/l, 80 g/l
and 100 g/l) from the rst decantation on the morphology of the tracks.
Experimental footprints were made by two people: a female individual with a
height of 1.69 m, weighing 55 kg and with a foot length of 24 cm, and a male
individual with a height of 1.80 m, weighing 75 kg and with a foot length of 24 cm.
The footprints were made in boxes of identical dimensions: 50 40 25 cm.
The second step consisted in covering the footprints with water (1.5 l) that
contained a dened sediment load (called the second decantation). Based on the
scenario that the cave suffered several low-power oods, the rst substrates of three
sediment loads were arbitrarily selected and tested (20 g/l, 40 g/l and 60 g/l) to see
whether there were any noticeable differences after the last ooding episode. As the
second substrate was less cohesive and less stable, it was more difcult to control for
its properties. We therefore chose to keep the three sediment loads in order to
understand more broadly the variability of the footprints in this type of sediment.
4 Reproduce to Understand: Experimental Approach Based on Footprints in Cussac... 75
A maximum of three ooding episodes were carried out for each footprint.
Finally, out of a total of 19 footprints, 8 were selected for the comparative
analysis. The remaining 11 are the tracks made when our experimental protocol
was established.
During the decantation process, the footprints were kept in a relatively stable
environment (21C and 5085% humidity according to the weather conditions
outside). The aim was to avoid excessively rapid drying and potential desiccation
cracks.
Descriptions, Metrics and 3D Models
After each step, the footprints were described in detail, distinguishing two aspects:
the general morphology, which concerns the shape and the outline of the footprint,
and its topography, related to its elevation and the state of its surface. In addition,
seven measurements considered as most indicative of the print morphology were
recorded: length 1 (distance between the most distal point of the hallux and the
most inferior point of the pternion), length 2 (distance between the most distal
point of the second toe and the most inferior point of the pternion), length
3 (distance between the most distal point of the forefoot and the most inferior
point of the pternion), digits width (distance between the most medial point of the
hallux and the most lateral point of the last toe), distal width (distance between the
most medial point and the most lateral point of the forefoot), middle width
(distance between the most medial point and the most lateral point of the longitu-
dinal arch) and proximal width (distance between the most medial point and the
most lateral point of the heel). They were photographed using a Nikon D7100 with
a 60 mm focal length lens. Then each footprint was 3D digitized using an Artec
EVA 3D light scanner 2013 (Artec Group, Luxembourg). This scanner uses the
structured light triangulation technique to reconstruct a 3D model of the footprint.
The accuracy achieved by this scanner is 0.5 mm at a working distance of 40 cm to
1 m, and the 3D resolution goes up to 0.1 mm. The scanner takes up to 16 frames
per second and transfers them to the Artec Studio software (Modabber et al. 2016)
which aligns the frames in real time.
Post-processing was performed on the Artec Studio 9 software, which recreated a
colour texturized 3D mesh.
The 3D models of the footprints at different moments of the experiment were
visualized and compared with CloudCompare (2.8.1.). We used part of the standard
protocol proposed by Falkingham et al. (2018) to record, present and archive our 3D
data. The true colour image, depth map and contour map (range of 0.5 mm) were
therefore created for each footprint.
76 L. Ledoux et al.
For the comparative analysis, clouds of each footprint were aligned using the
CloudCompare Align tool. The multiscale Model to Model Cloud Comparison
algorithm (M3C2) (Lague et al. 2013) was then used. It computes the local distance
directly between two point clouds along the normal surface direction. For each
distance measurement, it calculates a condence interval based on the point cloud
roughness and coregistration error. This computation serves to evaluate morpholog-
ical 3D changes in surface orientation. These changes are expressed in colourized
texture from the reference point cloud.
Results
Formation Sediment and Flooding Sediment Load
Formation Sediment
The tests carried out in order to verify the impact of different sediment loads (60 g/l,
80 g/l and 100 g/l) from the rst decantation on the morphology of the tracks do not
show any obvious differences between the footprints made on these three sediment
loads. Since the general morphology and topography did not seem to vary, we used
the average load of 60 g/l for subsequent experiments (Fig. 4.4a).
Flooding Sediment Load
No obvious differences were identied between the three sediment loads (20 g/l,
40 g/l and 60 g/l) tested on the footprints made in the rst substrate, particularly as
regard the loads of 40 g/l and 60 g/l. The average load of 60% was therefore used for
subsequent experiments (Fig. 4.4b).
General Morphology
The original experimental footprints are well dened and complete, regardless of the
formation sediment and the trackmaker. The distal and proximal parts are the deeper
ones. Although the middle part is shallower, the medial longitudinal arch is generally
well dened. Digit prints are also easily distinguishable throughout our sample
(Figs. 4.5a, b, c,4.6a, b, c).
Experimental ooding affects the footprintsmorphology in several ways. The
medial part of the footprint is the rst to disappear after ooding episodes,
irrespective of the formation sediment and the sediment load. After the third
4 Reproduce to Understand: Experimental Approach Based on Footprints in Cussac... 77
Fig. 4.4 (a) Footprints
made on the second surface,
no obvious difference
according to the sediment
load used; (b) footprints
made in raw clay and
covered with clay deposit
after three ooding
episodes, no obvious
difference according to the
sediment load used
ooding, this part is no longer visible on any print. The proximal part of the footprint
is the second part to disappear after ooding episodes. After the last ooding, this
part remains on two prints only. The distal part is the one that persists the longest
throughout the ooding episodes. Five footprints retain their distal part after the last
ooding. Within this part, the forefoot tends to be less visible more frequently than
the digits. The hallux is the most persistent of the digits (Figs. 4.5d, e, f,4.6d, e, f).
Generally, the ooding causes aloss of denition of the contours of the prints,
which could have distorted the way they were perceived when measurements were
taken. However, the many ooding episodes only modify the dimensions of the
remaining areas by a few millimetres. For some footprints, the measurements of
certain areas were sometimes over- or underestimated (Table 4.2). The footprint
made by individual 2 in the second surface, ooded with water loaded with 80 g/l of
sediment, is very representative, with a length that varies by almost 4 cm between the
original experimental footprint and the remaining part of it after the rst ooding
(Table 4.2).
78 L. Ledoux et al.
Fig. 4.5 Footprint made in the rst surface by individual 1 and ooded with water containing a
sediment load of 60 g/l; (a)rst step, true colour image; (b)rst step, contour map; (c)rst step,
depth map; (d) third ooding, true colour image; (e) third ooding, contour map; (f) third ooding,
depth map; (g) M3C2 distance between the rst step (original footprint) and the last step (after the
third ooding)
4 Reproduce to Understand: Experimental Approach Based on Footprints in Cussac... 79
Fig. 4.6 Footprint made in the second surface by individual 2 and ooded with water containing a
sediment load of 60 g/l; (a)rst step, true colour image; (b)rst step, contour map; (c)rst step,
depth map; (d) third ooding, true colour image; (e) third ooding, contour map; (f) third ooding,
depth map; (g) M3C2 distance between the rst step (original footprint) and the last step (after the
third ooding)
37 9
37 9
80 L. Ledoux et al.
Table 4.2 Biometry of the experimental footprints (cm)
Surface SL I Step
FL
1
FL
2
FL
3
FW
distal
FW
middle
FW
proximal
FW
digits
First
surface
60 g/l 1 1 23.8 23 19.9 8.5 3.2 5.3 9.5
2 23.3 22.4 19.4 8.3 5.3 9.6
3 23 22.3 19.8 8.5 5.5 9.5
4 8.5 9.5
2 1 24.7 24.3 20.8 10 4.7 9.5
2 9.7 9.5
3 9 8.5
Second
surface
40 g/l 1 1 23.7 23.5 20.4 8.5 4.2 5.3 9.0
2 23.4 22.4 20.7 8 4.5 5.2 9.3
4 7 8.4
2 1 25.2 23.3 21 11.2 3.3 6.4 9.7
2 25.5 24.4 20.7 10.7 5.3 10.6
4 7 8.4
60 g/l 1 1 24 23.7 20.4 10.2 6.2 6.2 10.5
2 24.4 4.5
2 1 24 23.4 20 10.9 3.5 6 10.3
2 23 22.5 19.2 10.4 4.5 10.2
3 10.6 10.2
4 9.2 9
80 g/l 1 1 23.6 22.9 20 8.9 5.1 5.7 9.6
2 23.5 22 19.4 8.2 5 5.6 9.6
3 23.1 22 20.1 8.1 4.6 6.5 9.4
2 1 24.9 24 21.4 10.9 5.2 5.7 10.5
2 21.4 20.7 20 10.5 5.3 4.9 10.4
FL foot length, FW foot width, Iindividual, SL sediment load
Topography
Original Experimental Footprints
The original experimental footprints made in the second surface are deeper than
those made in the raw clay.
For the footprints made in raw clay, raised rims are observed around the margins
of the digits, between the digits and the forefoot and sometime in the proximal part of
the heel (Fig. 4.5a, b, c).
For the footprints made in the second surface, prominent raised rims associated
with sediment displacement around their margins are more common (Fig. 4.6a, b, c).
These pronounced raised rims and the lack of cohesion of this substrate sometimes
led to the detachment of sediment plates (Fig. 4.6a, b, c).
4 Reproduce to Understand: Experimental Approach Based on Footprints in Cussac... 81
Footprint Evolution
Following ooding, the topography of the footprints was affected. The M3C2
algorithm allowed us to compare the surface changes of a single track between
two steps. This analysis reveals that footprints are lled up by clay deposits whose
thickness depends on the prints and the surface area of the prints (Figs. 4.5g and
4.6g).
For the footprints made in the rst surface, the deposits that formed on the surface
never exceed 10 mm after the last ooding (Fig. 4.5g).
For the footprints made in the second surface, the inlling is more complex
(Fig. 4.6g). In addition to clay deposits, footprints are often lled up by detached
sediment plates; inllings can then reach 20 mm.
In both surfaces, the majority of the areas affected by the inllings are most often
the deepest, such as the forefoot, the digits and the heel. The relief of the margins of
the prints tends to decrease. Raised rims were attened out and sediment plates
eroded (Figs. 4.5g and 4.6g).
All footprints lose denition after ooding episodes, and their margins are less
easily identiable. In general, impressions made in the second surface appear to be
more markedly altered than those made in the raw clay (Fig. 4.6d, e, f, g).
Discussion
Our experiments demonstrate that low-power oods do not modify the general
morphology of the prints, regardless of the formation sediment and the sediment
load used. However, as highlighted by the M3C2 algorithm, their topography is
altered by the clay deposits and a reduction in the relief of their margins. Some
detached plates resulting from the erosion of raised rims may also ll up the
footprints, particularly those made in the second surface. This detachment and
displacement of sediment is likely caused by the lack of cohesion of the substrate
due to its high moisture content: the higher the moisture content of the sediment, the
less cohesive it is. It may also be due to the lack of cohesion between the two levels
of the second surface. These characteristics make the surface more fragile, and the
track may be modied during ooding. Therefore, ooding episodes contribute to
the loss of track denition, and the forefoot and digits are generally the most
persistent areas.
These experiments also highlight that the use of biometric data on footprints to
infer biological characteristics such as sex, age, stature or body mass remains an
uncertain exercise. The lack of track denition and the taphonomical processes can
lead to measurement errors of several centimetres. These results are consistent with
previous taphonomic studies carried out on tracks found in open-air settings (Wise-
man and De Groote 2018; Zimmer et al. 2018). Based on the Holocene site of
Formby Point (North West England) for the former and the Pleistocene site Engare
Sero (Tanzania) for the latter, they perfectly illustrate and quantify the erosional
processes that occur immediately after track exposure. Both conclude that erosion-
related changes to tracks inuence biological inferences. Previous studies have also
demonstrated the uniqueness of tracks and the crucial role of the substrate in which
they were formed (Pales 1976; Marty et al. 2009; Morse et al. 2013; Bennett and
Morse 2014). Furthermore, it is known that a single trackmaker could produce a
range of footprints with various morphologies according to the sediment on which
they were formed (Morse et al. 2013; Bennett and Morse 2014). It has also been
demonstrated that footprints are almost systematically larger than the feet that made
them (Pales 1976; Hatala et al. 2018). Additionally, most inferences are based on
modern reference populations. Regarding fossil tracks, there is no guarantee that the
reference population used is representative of past variability (Bennett and Morse
2014). Although inferences made on tracks should be used with caution, they
provide some insights for interpretation purposes. Experiments are therefore a useful
tool to approximate the original shape of a track as closely as possible and/or to
understand its alterations (Bennett and Morse 2014; Falkingham et al. 2018).
82 L. Ledoux et al.
This work, based on observations made on the footprints found in Cussac Cave,
provides some insights into the taphonomical effects of ooding events on the
morphology and topography of footprints. Experiments based on taphonomical
phenomena are still limited and mainly concern tracks found in open-air settings
(Marty et al. 2009; Scott et al. 2010; Morse et al. 2013; Bennett and Morse 2014;
Roach et al. 2016; Wiseman and De Groote 2018). The major difference between
tracks found in open-air settings and those found in caves is probably the speed of
taphonomical processes affecting them. Studies of tracks found in open-air settings
have shown that a multitude of taphonomical processes (weather condition, biotur-
bation, properties of the sediment, etc.) preceded the burial and the diagenesis of the
tracks. Consequently, their morphology was rapidly altered (Marty et al. 2009; Scott
et al. 2010; Bennett et al. 2013; Wiseman and De Groote 2018; Zimmer et al. 2018).
Additionally, their exposure led to degradations (Wiseman and De Groote 2018;
Zimmer et al. 2018). Conversely, caves are stable environments allowing a high
degree of track preservation. While tracks found in caves may be altered, it is
assumed that they are disturbed less than those found in open-air settings. However,
our work has demonstrated that although the damage to the footprint does not
substantially alter its general morphology, its loss of denition or the destruction
of certain parts can lead to unreliable interpretation.
Our analysis was based on 3D data. These techniques have become crucial in the
study of ornated caves and are now replacing casts and other recording methods.
They are most often used for conservation purposes and to encourage ex situ studies.
They are also essential as they provide a precise picture of human and animal use of
caves (Ortega Martinez et al. 2014; Pastoors et al. 2017; Citton et al. 2017; Ledoux
2019; Romano et al. 2019). Here we used the M3C2 algorithm (Lague et al. 2013)in
order to quantify the surface changes to single footprints between each step. The
same algorithm was used to quantify the ongoing erosion of the Engare Sero tracks
(Zimmer et al. 2018). So far, most of the tools developed in ichnological studies have
been based on the biomechanics of hominin locomotion. Consequently, these studies
are more focused on the nature of the formation substrate and its interaction with the
foot (Crompton et al. 2012; Morse et al. 2013; Hatala et al. 2018). While some tools
such as pedobarographic statistical parametric mapping (pSPM), based on the
comparison of pressure at the substrate-foot interface and footprint depth (Pataky
et al. 2008a; Crompton et al. 2012; Morse et al. 2013) or biplanar X-rays studying
the 3D dynamics at the foot-substrate interface, have focused on the formation of
tracks to infer foot anatomy or biomechanics data (Hatala et al. 2018), the M3C2
algorithm focused on the evolution of these tracks over time. The application of such
methods is then useful to complement qualitative observations and can help to
understand certain taphonomical processes such as erosion or sedimentation.
4 Reproduce to Understand: Experimental Approach Based on Footprints in Cussac... 83
While the experiment presented here provides promising data on the impact of
clay deposits on the morphometry of a footprint, our sample was limited, and we
only explored and controlled a few parameters. Additionally, these parameters do
not necessarily extend to all caves and all tracks. The future integration of a larger
sample of tracks produced by a larger number of trackmakers, in a variety of
sediments combined with varied sediment loads contained in the ooding water,
will undoubtedly further substantiate our results. This would also allow researchers
to create reference tracks for each possible setting that could be used to study the
tracks of different sites. Many phenomena and their inuence on the morphology
and biometry of tracks found in karst settings have yet to be documented: these
experiments are the rst step in the development of more experimental work. The
creation of articial ooding on the very limited surface of the box does not
accurately reect the reality of the overow of a subterranean river. It would
therefore be appropriate to carry out experiments directly in karst settings. One of
the advantages of laboratory experiments is that they make it possible to recreate
taphonomical phenomena in a very short time. However, the more complex the
phenomenon, the more difcult it will be to control. Although this requires much
more time, it would therefore be better to follow the evolution of tracks in real
conditions and also taking the geometry of the cave into consideration.
Our results demonstrate that ooding and subsequent clay deposits in some areas
of Cussac contributed to the lack of visibility of tracks. However, they do not explain
the lack of details in Cussacs tracks. Only one complete footprint could be
interpreted as undoubtedly human in the submerged areas (Fig. 4.1a). Although its
outline is clearly visible with all the foot areas represented (forefoot, longitudinal
arch and heel), no detail is apparent. These experiments show that clay deposits did
not radically modify the morphology of the footprints and allowed the preservation
of certain details such as digits, regardless of the formation sediment and the
sediment load used. Additionally, the existence of low-power oods has been proven
at Cussac. Apart from the clay deposits and some desiccation cracks, the oor does
not seem to have suffered any other alteration. In our experiments, what seems to
have had the most signicant impact is the lack of cohesion of the second surface due
to its high moisture content and its level of clay resulting from settling, causing the
raising and displacement of sediment plates on the surface during water inltration.
However, this phenomenon did not occur at Cussac. The next step will therefore be
to understand this lack of detail. Is it due to ooding or another taphonomical
phenomenon? In addition to ooding, the areas involved were intensively trampled
by bears, so discrimination between the two species is a challenge. By continuing
our experiments, we hope to improve the determination of the prints present in these
problematic areas.
84 L. Ledoux et al.
Conclusion
As very few experimental works have been carried out in caves, this study will
emerge as original in this type of setting. It brings new data on the taphonomy of
tracks when they are subjected to ooding. Although ooding does not modify the
general morphology of the tracks, their topography is altered by successive episodes
and clay deposits. However, the loss of track denition and the taphonomical
processes can lead to unreliable interpretation and measurement errors of several
centimetres. Inferences on fossil tracks should therefore be made with caution. Our
experiments were based on taphonomical phenomena observed in Cussac Cave.
Although we do not yet have all the means to reliably interpret the tracks of Cussac, a
larger sample involving more parameters and in situ experiments will undoubtedly
allow us to rene our results and apply them to other caves.
Tracks are a signicant testimony of the frequentation of caves by Palaeolithic
people and their ability to adapt to an unsuitable or even dangerous environment. It is
therefore essential to understand their history if we want to reconstruct past human
behaviour and activities in caves and in Palaeolithic societies.
Acknowledgements We are grateful to the editors for allowing us to contribute to this chapter of
this book. Our work is supported by the Projet Collectif de Recherche Cussac, and we thank all the
members of its team. We are also grateful to the French Ministry of Culture and to the LaScArBx, a
research programme supported by the ANR (ANR-10-LABX-52). We are also grateful to the
University of Bordeaux, the PACEA laboratory and the Pôle mixte de recherche archéologique
de Campagne. We also thank Mathieu Bosq for his contribution to the experiments and Christophe
Mallet for his help and advice.
References
Altamura, F., Bennett, M. R., DAoût, K., Gaudzinski-Windheuser, S., Melis, R. T., Reynolds,
S. C., & Mussi, M. (2018). Archaeology and ichnology at Gombore II-2, Melka Kunture,
Ethiopia: Everyday life of a mixed-age hominin group 700,000 years ago. Scientic Reports, 8,
2815. https://doi.org/10.1038/s41598-018-21158-7.
Ambert, P., Colomer, A., & Galant, P. (2000). Datations mésolithiques des empreintes humaines de
létage Cathala de la grotte dAldène (Cesseras, Hérault). Comptes Rendus de lAcadémie des
Sciences Series IIA Earth and Planetary Science, 331,6774. https://doi.org/10.1016/
S1251-8050(00)01380-X.
4 Reproduce to Understand: Experimental Approach Based on Footprints in Cussac... 85
Ambert, P., Galant, P., & Colomer, A. (2001). Incursions spéléologiques mésolithiques dans la
grotte dAldène (Cesseras, Hérault). Bulletin de la Société préhistorique française, 98, 497503.
https://doi.org/10.3406/bspf.2001.12535.
Aramayo, S. A. (2009). A brief sketch of the Monte Hermoso human footprint site, South Coast of
Buenos Aires Province, Argentina. Ichnos, 16,4954. https://doi.org/10.1080/
10420940802470748.
Arias, P., Laval., E., Menu, M., Sainz, C.G., Ontañon, R. (2011). Les colorants dans lart pariétal et
mobilier paléolithique de la Garma (Cantanbrie, Espagne). LAnthropologie, 115(34),
425445.
Ashton, N., Lewis, S. G., De Groote, I., Duffy, S. M., Bates, M., Bates, R., Hoare, P., Lewis, M.,
Partt, S. A., Peglar, S., Williams, C., & Stringer, C. (2014). Hominin footprints from early
Pleistocene deposits at Happisburgh, UK. PLoS ONE, 9.https://doi.org/10.1371/journal.pone.
0088329.
Aujoulat, N., Geneste, J.-M., Archambeau, C., Delluc, M., Duday, H., & Gambier, D. (2001). La
grotte ornée de Cussac (Dordogne). Observations préliminaires. PALEO Revue darchéologie
préhistorique, 13,918.
Aujoulat, N., Geneste, J.-M., Archambeau, C., Delluc, M., Duday, H., & Henry-Gambier,
D. (2002). La grotte ornée de Cussac Le Buisson-de-Cadouin (Dordogne) : premières
observations. Bulletin de la Société préhistorique française, 99, 129137. https://doi.org/10.
3406/bspf.2002.12612.
Aujoulat, N., Feruglio, V., Fourment, N., Henry-Gambier, D., & Jaubert, J. (2013). Le sanctuaire
gravettien de Cussac (Le Buisson-de-Cadouin, Dordogne, France) : premiers résultats dun
projet collectif de recherche. International Newsletter on Rock Art (INORA), 65,719.
Bégouën, R., Fritz, C., Tosello, G., Clottes, J., Pastoors, A., & Faist, F. (2009). Le Sanctuaire secret
des bisons. Il y a 14000 ans, dans la caverne du Tuc dAudoubert. Paris: Editons dart Somogy
Association Louis Bégouën.
Bennett, M. R., & Morse, S. A. (2014). Human footprints: Fossilised locomotion? Cham: Springer
International Publishing.
Bennett, M. R., Harris, J. W. K., Richmond, B. G., Braun, D. R., Mbua, E., Kiura, P., Olago, D.,
Kibunjia, M., Omuombo, C., Behrensmeyer, A. K., Huddart, D., & Gonzalez, S. (2009). Early
hominin foot morphology based on 1.5-million-year-old footprints from Ileret, Kenya. Science,
323, 11971201. https://doi.org/10.1126/science.1168132.
Bennett, M. R., Falkingham, P., Morse, S. A., Bates, K., & Crompton, R. H. (2013). Preserving the
impossible: Conservation of soft-sediment hominin footprint sites and strategies for three-
dimensional digital data capture. PLoS One, 8, e60755. https://doi.org/10.1371/journal.pone.
0060755.
Bennett, M. R., Reynolds, S. C., Morse, S. A., & Budka, M. (2016). Laetolis lost tracks: 3D
generated mean shape and missing footprints. Scientic Reports, 6, 21916.
Berge, C., Penin, X., & Pellé, É. (2006). New interpretation of Laetoli footprints using an
experimental approach and Procrustes analysis: Preliminary results. Comptes Rendus Palevol,
5, 561569. https://doi.org/10.1016/j.crpv.2005.09.001.
Brand, L. R. (1996). Variations in salamander trackways resulting from substrate differences.
Journal of Paleontology, 70, 10041010. https://doi.org/10.1017/S0022336000038701.
Burns, A. (2014). The prehistoric footprints at Formby: Discover the footprints on the Sefton coast
and take a glimpse into prehistoric Britain. Ainsdale: Sefton Coast Landscape Partnership
Scheme.
Citton, P., Romano, M., Salvador, I., & Avanzini, M. (2017). Reviewing the upper Pleistocene
human footprints from the Sala dei Misteriin the Grotta della Bàsura (Toirano, northern Italy)
cave: An integrated morphometric and morpho-classicatory approach. Quaternary Science
Reviews, 169,5064. https://doi.org/10.1016/j.quascirev.2017.05.016.
86 L. Ledoux et al.
Clottes, J., & Simonnet, R. (1972). Le réseau René Clastres de la caverne de Niaux (Ariège).
Bulletin de la Société préhistorique française Études et travaux, 69(1), 293323.
Crompton, R. H., Pataky, T. C., Savage, R., DAoût, K., Bennett, M. R., Day, M. H., Bates, K.,
Morse, S., & Sellers, W. I. (2012). Human-like external function of the foot, and fully upright
gait, conrmed in the 3.66 million year old Laetoli hominin footprints by topographic statistics,
experimental footprint-formation and computer simulation. Journal of the Royal Society Inter-
face, 9, 707719. https://doi.org/10.1098/rsif.2011.0258.
DAoût, K., Meert, L., Van Gheluwe, B., De Clercq, D., & Aerts, P. (2010). Experimentally
generated footprints in sand: Analysis and consequences for the interpretation of fossil and
forensic footprints. American Journal of Physical Anthropology, 141, 515525. https://doi.org/
10.1002/ajpa.21169.
Duday, H., & Garcia, M. A. (1983). Les empreintes de lHomme préhistorique. La grotte du Pech-
Merle à Cabrerets (Lot): une relecture signicative des traces de pieds humains. Bulletin de la
Société préhistorique française, 80(7), 208215.
Duday, H., & Garcia, M. A. (1985). Lhomme et la caverne. Dossiers dArchéologie, 90,3539.
Duday, H., & Garcia, M. A. (1986). La Paléo Ichnologie humaine (étude des empreintes fossiles).
Bulletin de la Société dAnthropologie du Sud-Ouest Bordeaux, 21,1.
Falkingham, P. L., Bates, K. T., Avanzini, M., Bennett, M., Bordy, E. M., Breithaupt, B. H.,
Castanera, D., Citton, P., Díaz-Martínez, I., Farlow, J. O., Fiorillo, A. R., Gatesy, S. M., Getty,
P., Hatala, K. G., Hornung, J. J., Hyatt, J. A., Klein, H., Lallensack, J. N., Martin, A. J., Marty,
D., Matthews, N. A., Meyer, C. A., Milàn, J., Minter, N. J., Razzolini, N. L., Romilio, A.,
Salisbury, S. W., Sciscio, L., Tanaka, I., Wiseman, A. L. A., Xing, L. D., & Belvedere,
M. (2018). A standard protocol for documenting modern and fossil ichnological data.
Palaeontology, 61, 469480. https://doi.org/10.1111/pala.12373.
Felstead, N. J., Gonzalez, S., Huddart, D., Noble, S. R., Hoffmann, D. L., Metcalfe, S. E., Leng,
M. J., Albert, B. M., Pike, A. W. G., Gonzalez-Gonzalez, A., & Jiménez-López, J. C. (2014).
Holocene-aged human footprints from the Cuatrociénegas Basin, NE Mexico. Journal of
Archaeological Science, 42, 250259. https://doi.org/10.1016/j.jas.2013.11.010.
Fourment, N., Baraud, D., Kazmierczak, M., & Rieu, A. (2012). La grotte de Cussac (Le Buisson-
de-Cadouin, Dordogne, France): Applications des principes de conservation préventive au cas
dune découverte récente. In J. Clottes (Ed.), Lart pléistocène dans le monde/Pleistocene artof
the worls/Arte pleistoceno en el mundo. Actes du Congrès IFRAO, Tarascon-sur-Ariège,
septembre 2010 (pp. 343354). Tarascon-sur-Ariège: Société Préhistorique Ariège-Pyrénées.
Garcia, M. A. (1986). Pour une bibliographie des empreintes humaines et animales. Paris:
Publications scientiques du Muséum national dHistoire naturelle.
Garcia, M. A. (2001). Les empreintes et les traces humaines et animales. In J. Clottes (Ed.), La
grotte Chauvet. Lart des origines (pp. 3443). Paris: Éditions du Seuil.
Garcia, M. A. Ichnologie générale de la grotte Chauvet. Bulletin de la Société préhistorique
française, 102, 103108. https://doi.org/10.3406/bspf.2005.13341.
Garcia, M. A., & Duday, H. (1983). Grotte de Foissac (Aveyron) A propos dune découverte
récente ou de lichnologie comme mode dapproche des structures préhistoriques en grotte.
Bulletin de la Société préhistorique française, 80, 184187. https://doi.org/10.3406/bspf.1983.
5441.
Garcia, M. A., Duday, H., & Courtaud, P. (1990). Les empreintes du Réseau Clastres. Préhistoire
Ariégeoise, 45, 167174.
Gatesy, S. (2003). Direct and indirect track features: What sediment did a dinosaur touch? Ichnos,
10,9198. https://doi.org/10.1080/10420940390255484.
Hatala, K. G., Perry, D. A., & Gatesy, S. M. (2018). A biplanar X-ray approach for studying the 3D
dynamics of human track formation. Journal of Human Evolution, 121, 104118. https://doi.
org/10.1016/j.jhevol.2018.03.006.
Hatala, K. G., Dingwall, H. L., & Wunderlich, R. E. Richmond, B.G. (2013) the relationship
between plantar pressure and footprint shape. Journal of Human Evolution, 65,2128. https://
doi.org/10.1016/j.jhevol.2013.03.009.
4 Reproduce to Understand: Experimental Approach Based on Footprints in Cussac... 87
Henry-Gambier, D., Courtaud, P., Dutailly, B., Villotte, S., Deguilloux, M. F., Pemonge, M. H.,
Aujoulat, N., Delluc, M., Fourment, N., & Jaubert, J. (2013). Grotte de Cussac (Le Buisson-de-
Cadouin, Dordogne): un exemple de comportement original pour le Gravettien. In J. Jaubert,
N. Fourment, & P. Depaepe (Eds.), Transitions, ruptures et continuité en Préhistoire, Actes du
XXVIIe congrès préhistorique de France, Bordeaux-Les Eyzies 31 mai-5 juin 2010
(pp. 169182). Paris: Société Préhistorique Française.
Jaubert, J. (2015). Une invitation à Jiří Svoboda pour jumeler le complexe pavlovien de Moravie et
la grotte ornée et funéraire dâge Gravettien de Cussac. In S. Sázelová, M. Novák, &
A. Mizerová(Eds.), Forgotten times and spaces: New perspectives in paleoanthropological,
paleoetnological and archeological studies (pp. 214228). Brno: Masaryk university.
Jaubert, J., Aujoulat, N., Courtaud, P., Deguilloux, M. F., Delluc, M., Denis, A., Duday, H.,
Dutailly, B., Ferrier, C., Feruglio, V., Fourment, N., Geneste, J.-M., Genty, D., Goutas, N.,
Henry-Gambier, D., Kervazo, B., Klaric, L., Lastennet, R., Levêque, F., Malaurent, P., Mallye,
J.-B., Mora, P., Pemonge, M. H., Peyraube, N., Peyroux, M., Plisson, H., Portais, J.-C.,
Valladas, H., Vergnieux, R., & Villotte, S. (2012). Le projet collectif de recherche Grotte de
Cussac(Dordogne, France): étude dune cavité ornée à vestiges humains du Gravettien. In
J. Clottes (Ed.), Lart pléistocène dans le monde/Pleistocene artof the worls/Arte pleistoceno en
el mundo. Actes du Congrés IFRAO, septembre 2010 (pp. 325342). Tarascon-sur Ariège:
Société Préhistorique Ariège-Pyrénées.
Jaubert, J., Genty, D., Valladas, H., Camus, H., Courtaud, P., Ferrier, C., Feruglio, V., Fourment,
N., Konik, S., Villotte, S., Bourdier, C., Costamagno, S., Delluc, M., Goutas, N., Katnecker, É.,
Klaric, L., Langlais, M., Ledoux, L., Maksud, F., OFarrell, M., Mallye, J.-B., Pierre, M., Pons-
Branchu, E., Régnier, É., & Théry-Parisot, I. (2017). The chronology of human and animal
presence in the decorated and sepulchral cave of Cussac (France). Quaternary International,
432, 5424. https://doi.org/10.1016/j.quaint.2016.01.052.
Lague, D., Brodu, N., & Leroux, J. (2013). Accurate 3D comparison of complex topography with
terrestrial laser scanner: Application to the Rangitikei canyon (N-Z). ISPRS Journal of Photo-
grammetry and Remote Sensing, 82,1026. https://doi.org/10.1016/j.isprsjprs.2013.04.009.
Ledoux, L. (2019). Lichnologie préhistorique et les traces dactivités au sein des cavités ornées.
Les grottes de Fontanet (Ariège) et de Cussac (Dordogne). PhD thesis, Université de Bordeaux,
France.
Ledoux, L., Fourment, N., Maksud, F., Delluc, M., Costamagno, S., Goutas, N., Klaric, L.,
Laroulandie, V., Salomon, H., & Jaubert, J. (2017). Traces of human and animal activity
(TrAcs) in Cussac Cave (Le Buisson-de-Cadouin, Dordogne, France): Preliminary results and
perspectives. Quaternary International, 430, 141154. https://doi.org/10.1016/j.quaint.2016.
06.002.
Marty, D., Strasser, A., & Meyer, C. A. (2009). Formation and taphonomy of human footprints in
microbial mats of present-day tidal-at environments: Implications for the study of fossil
footprints. Ichnos, 16, 127142. https://doi.org/10.1080/10420940802471027.
Masao, F. T., Ichumbaki, E. B., Cherin, M., Barili, A., Boschian, G., Iurino, D. A., Menconero, S.,
Moggi-Cecchi, J., & Manzi, G. (2016). New footprints from Laetoli (Tanzania) provide
evidence for marked body size variation in early hominins. eLife, 5.https://doi.org/10.7554/
eLife.19568.
McLaren, D., Fedje, D., Dyck, A., Mackie, Q., Gauvreau, A., & Cohen, J. (2018). Terminal
Pleistocene epoch human footprints from the Pacic coast of Canada. PLoS One, 13,
e0193522. https://doi.org/10.1371/journal.pone.0193522.
Medina-Alcaide, M. Á., Garate-Maidagan, D., Ruiz-Redondo, A., & Sanchidrián-Torti, J. L.
(2018). Beyond art: The internal archaeological context in Paleolithic decorated caves. Journal
of Anthropological Archaeology, 49, 114128. https://doi.org/10.1016/j.jaa.2017.12.005.
Mietto, P., Avanzini, M., & Rolandi, G. (2003). Palaeontology: Human footprints in Pleistocene
volcanic ash. Nature, 422, 133133.
88 L. Ledoux et al.
Milàn, J., & Bromley, R. G. (2007). The impact of sediment consistency on track and undertrack
morphology: Experiments with emu tracks in layered cement. Ichnos, 15,1927. https://doi.org/
10.1080/10420940600864712.
Modabber, A., Peters, F., Kniha, K., Goloborodko, E., Ghassemi, A., Lethaus, B., Hölzle, F., &
Möhlhenrich, S. C. (2016). Evaluation of the accuracy of a mobile and a stationary system for
three-dimensional facial scanning. Journal of Cranio-Maxillofacial Surgery, 44, 17191724.
Moreno, K., Bostelmann, J. E., Macías, C., Navarro-Harris, X., Pol-Holz, R. D., & Pino, M. (2019).
A late Pleistocene human footprint from the Pilauco archaeological site, northern Patagonia,
Chile. PLoS One, 14, e0213572. https://doi.org/10.1371/journal.pone.0213572.
Morse, S. A., Bennett, M. R., Liutkus-Pierce, C., Thackeray, F., McClymont, J., Savage, R., &
Crompton, R. H. (2013). Holocene footprints in Namibia: The inuence of substrate on footprint
variability: Holocene footprints in Namibia. American Journal of Physical Anthropology, 151,
265279. https://doi.org/10.1002/ajpa.22276.
Ortega Martinez, A. I., Ruiz García, F., Benito Calvo, A., Martín Merino, M. A., Karampaglidis, T.,
& Campaña Lozano, I. (2014). Escaneado en 3D de las Galerías de las Huellas (Ojo Guareña,
Merindad de Sotoscueva, Burgos). Cubía Boletin, 18,3847.
Pales, L. (1976). Les Empreintes de pieds humains dans les cavernes. In Les empreintes du réseau
nord de la caverne de Niaux (Ariège). Paris et al.: Masson.
Panerello, A., Santello, L., Farinaro, G., Bennett, M. R., & Mietto, P. (2017). Walking along the
oldest human fossil pathway (Roccamonna volcano, Central Italy)? Journal of Archaeological
Science: Reports, 13, 476490. https://doi.org/10.1016/j.jasrep.2017.04.020.
Pastoors, A., & Weniger, G.-C. (2011). Cave art in context: Methods for the analysis of the spatial
organization of Cave Sites. Journal of Archaeological Research, 19, 377400.
Pastoors, A., Lenssen-Erz, T., Ciqae, T., Kxunta, U., Thao, T., Bégouën, R., Biesele, M., & Clottes,
J. (2015). Tracking in caves: Experience based reading of pleistocene human footprints in
French caves. Cambridge Archaeological Journal, 25, 551564. https://doi.org/10.1017/
S0959774315000050.
Pastoors, A., Lenssen-Erz, T., Breuckmann, B., Ciqae, T., Kxunta, U., Rieke-Zapp, D., & Thao,
T. (2017). Experience based reading of Pleistocene human footprints in Pech-Merle. Quater-
nary International, 430, 155162. https://doi.org/10.1016/j.quaint.2016.02.056.
Pataky, T. C., & Goulermas, J. Y. (2008). Pedobarographic statistical parametric mapping (pSPM):
A pixel-level approach to foot pressure image analysis. Journal of Biomechanics, 41,
21362143. https://doi.org/10.1016/j.jbiomech.2008.04.034.
Pataky, T. C., Caravaggi, P., Savage, R., Parker, D., Goulermas, J. Y., Sellers, W. I., & Crompton,
R. H. (2008a). New insights into the plantar pressure correlates of walking speed using
pedobarographic statistical parametric mapping (pSPM). Journal of Biomechanics, 41,
19871994. https://doi.org/10.1016/j.jbiomech.2008.03.034.
Pataky, T. C., Goulermas, J. Y., & Crompton, R. H. (2008b). A comparison of seven methods of
within-subjects rigid-body pedobarographic image registration. Journal of Biomechanics, 41,
30853089. https://doi.org/10.1016/j.jbiomech.2008.08.001.
Raichlen, D. A., Gordon, A. D., Harcourt-Smith, W. E. H., Foster, A. D., & Haas, W. M. R. (2010).
Laetoli footprints preserve earliest direct evidence of human-like bipedal biomechanics. PLoS
One, 5, e9769. https://doi.org/10.1371/journal.pone.0009769.
Roach, N. T., Hatala, K. G., Ostrofsky, K. R., Villmoare, B., Reeves, J. S., Du, A., Braun, D. R.,
Harris, J. W. K., Behrensmeyer, A. K., & Richmond, B. G. (2016). Pleistocene footprints show
intensive use of lake margin habitats by Homo erectus groups. Scientic Reports, 6, 26374.
https://doi.org/10.1038/srep26374.
Romano, M., Citton, P., Salvador, I., Arobba, D., Rellini, I., Firpo, M., Negrino, F., Zunino, M.,
Starnini, E., & Avanzini, M. (2019). A multidisciplinary approach to a unique Palaeolithic
human ichnological record from Italy (Bàsura Cave). bioRxiv, 529404.https://doi.org/10.1101/
529404.
4 Reproduce to Understand: Experimental Approach Based on Footprints in Cussac... 89
Scott, J. J., Renaut, R. W., & Owen, R. B. (2010). Taphonomic controls on animal tracks at saline,
alkaline Lake Bogoria, Kenya Rift Valley: Impact of salt eforescence and clay mineralogy.
Journal of Sedimentary Research, 80, 639665. https://doi.org/10.2110/jsr.2010.057.
Smith, G. M., Parsons, T., Harrod, R. P., Holmes, C. E., Reuther, J. D., & Potter, B. A. (2019). A
track in the Tanana: Forensic analysis of a late Holocene footprint from Central Alaska. Journal
of Archaeological Science: Reports, 24, 900912. https://doi.org/10.1016/j.jasrep.2019.03.016.
Sollas, W. J. (1879). On some three-toed footprints from the Triassic conglomerate of South Wales.
Quarterly Journal of the Geological Society, 35, 511515. https://doi.org/10.1144/GSL.JGS.
1879.035.01-04.33.
Webb, S., Cupper, M. L., & Robins, R. (2006). Pleistocene human footprints from the Willandra
Lakes, southeastern Australia. Journal of Human Evolution, 50, 405413. https://doi.org/10.
1016/j.jhevol.2005.10.002.
Wiseman, A. L. A., & De Groote, I. (2018). A three-dimensional geometric morphometric study of
the effects of erosion on the morphologies of modern and prehistoric footprints. Journal of
Archaeological Science: Reports, 17,93102. https://doi.org/10.1016/j.jasrep.2017.10.044.
Zimmer, B., Liutkus-Pierce, C., Marshall, S. T., Hatala, K. G., Metallo, A., & Rossi, V. (2018).
Using differential structure-from-motion photogrammetry to quantify erosion at the Engare Sero
footprint site, Tanzania. Quaternary Science Reviews, 198, 226241. https://doi.org/10.1016/j.
quascirev.2018.07.006.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Chapter 5
Experimental Re-creation
of the Depositional Context in Which Late
Pleistocene Tracks Were Found
on the Pacic Coast of Canada
Duncan McLaren, Quentin Mackie, and Daryl Fedje
Abstract To better understand the depositional context of Late Pleistocene human
tracks found at archaeology site EjTa-4 on Calvert Island, on the Pacic Coast of
Canada, we present here the results of an experiment designed to recreate the
conditions by which these tracks were formed, preserved and then revealed through
excavation. Based on radiocarbon ages on small twigs and the analysis of sediments
and microfossils, the interpretation of the site formation processes relate that the
tracks were impressed into a clayey soil substrate just above the high tide line
between 13,317 and 12,633 calBP. The features were subsequently encapsulated
by black sand, which washed over the tracks from the nearby intertidal zone during a
storm event. To test this interpretation, we enlisted the aid of high school student
volunteers to recreate the conditions by which the tracks were formed. A clayey
substrate was prepared in a laboratory setting at the University of Victoria and a few
plant macrofossils were placed on top it. This was followed by having the students
create tracks in the clay, which were then covered with a layer of sand. Upon
excavation of these experimental tracks, we found that they had a very similar
character to those found in the eld, including the pressing of macrofossils into
the clay by the weight of the track maker. These results support the interpretation and
chronological assessment of the depositional events that occurred during late Pleis-
tocene times at archaeology site EjTa-4.
Keywords Footprint · Experimentation · Open air · Monitoring
D. McLaren (*) · D. Fedje
Hakai Institute, Campbell River, BC, Canada
University of Victoria, Victoria, BC, Canada
e-mail: dsmclaren@gmail.com
Q. Mackie
University of Victoria, Victoria, BC, Canada
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_5
91
Introduction
This paper follows up on archaeological ndings of human tracks dating to the Late
Pleistocene occupation of the Pacic coast of Canada (McLaren et al. 2018)
(Fig. 5.1). These tracks were found in soft sediments beneath intertidal beach
deposits on Calvert Island, British Columbia. A total of 29 tracks were identied
and isolated during the excavation of a 4 ×2 m area at archaeological site EjTa-4
(Meay Channel I). Lodgepole pine twigs found on the track surface and pressed into
the base of the footprint (referred to here as the true track) provide us with radio-
carbon age estimates of 13,31712,633 calBP (see Marty et al. 2009; Bennett and
Morse 2014 for denition of track terms).
92 D. McLaren et al.
Archaeological site EjTa-4 is located on the western shore of Meay Channel, an
inner and protected waterway situated between Calvert and Hecate Islands. Calvert
Island features as a location in events related in the oral histories of the Heiltsuk
(Olson 1955) and Wuikinuxv First Nations (Walkus et al. 1982). Some of the
recorded oral histories from the region relate events, such as large-scale glaciation,
that have not occurred since Late Pleistocene times (Gauvreau and McLaren 2016).
Ancient human track sites have not been widely reported in North America north
of Mexico. There are some exceptions. For example, a recent publication describes a
Late Pleistocene trackway in New Mexico which has associated human and giant
ground sloth footprints dating between 15,500- and 10,000-year-old (Bustos et al.
2018). Willey et al. (2009) provide a summary of Holocene human footprints
reported from North America, with a more recent discovery reported from Swan
Fig. 5.1 Location of archaeological site EjTa-4 on the Pacic coast of Canada
Point Alaska dated to 1840 calBP (Smith et al. 2019). Overall, however, human
tracks appear to be an extremely rare site type across the continent.
5 Re-creation of Depositional Context of Footprints 93
Nearshore Late Pleistocene archaeological research on west coast of Canada
requires a good knowledge of local relative sea level history (Clague et al. 1982;
Fedje and Mathewes 2005; Fedje et al. 2018; Shugar et al. 2014). Due primarily to
the dynamic interplay of isostatic and eustatic factors, sea level on Calvert Island was
23 m lower than today between 14,000 and 11,000 years ago (McLaren et al. 2014).
The tracks at this site were discovered during subsurface testing below beach
deposits. This subsurface testing was specically targeting this lower shoreline and
time period. However, we were not expecting to nd human tracks.
Of the 29 individual tracks found, 18 were complete enough to take measure-
ments of length and width (McLaren et al. 2018). These measurements fall into three
broad categories of size (15.5 ×7; 20 ×9; and 25.5 ×11.5 cm), suggesting that a
minimum of three individuals of different foot sizes left the tracks. The majority of
tracks were found to be oriented towards the northwest or landward and away from
the ocean. A few rough grained stone tools were found in the same stratigraphic
layer.
The track surface is a light brown clayey paleosol that was located above the high
tide line at the time of deposition (referred to as Stratum X). It is overlain by black
pebbly sand which was washed up from the beach lling the tracks (Stratum IX).
Based on our analyses of these strata, we interpret that the formation of these features
involved a minimum of three people leaving footprints in a clayey area above the
high tide line between 13,317 and 12,633 calBP. Later, by at least 12,64012,576
calBP, a change in sea level, storm surge or tsunami event resulted in the dumping of
sand and pea gravel onto the track surface thereby lling and capping the features.
Overlying all of this are sandy gravels with late Holocene artefacts and bone (Strata
VII through II), capped at the top by active sands just below the beach surface
(Stratum I).
The contrasting colours between the track surface and the over track deposits
enabled us to identify the true tracks. As our eld crew excavated down through the
black sand deposits into the more clayey deposits below, sediment displacement
rims were the rst indicators found (Fig. 5.2). Through careful and delicate excava-
tion, the sediment rims were isolated revealing the tracks. In some cases, toe marks
were clearly visible. Photographs were taken of each of the individual tracks, and
contrast enhancement software was used to further reveal the features for publication
(Fig. 5.3). Multiple photos of all the tracks found are included in the supplemental
data that is associated with the original publication (McLaren et al. 2018).
Beyond providing information on the inhabitants that made these tracks, the
ndings are of signicance as they have bearing on the Late Pleistocene occupation
of North America. This is one of the earliest human occupational records on the
Pacic coast of Canada and provides evidence of the early postglacial use of this part
of the coast. Those who left the footprints at EjTa-4 could only have reached Calvert
Island by means of watercraft. These inhabitants most likely had an economy that
was heavily focussed on the marine environment, as did later populations in the
region (Dufeld 2017).
94 D. McLaren et al.
Fig. 5.2 Plan view of 4 ×2 m excavation unit and the locations and orientation of tracks and
radiocarbon dates
As a part of the ongoing research associated with the discovery of these tracks, we
undertook a lab-based experiment to see if we could recreate the sedimentary
conditions in which the tracks were created and then buried. As a part of this process,
the experimental footprints were excavated to compare our eld ndings with those
in the lab. This primary goal of the exercise was to help understand if our site
formation process interpretation was supported by experimental approach.
As a part of informing the public about aspects of our research, we have been
working with school groups from local communities including Bella Bella,
Oweekeno Village, Bella Coola as well as in and around Victoria, British Columbia.
The experiments discussed here were conducted as part of the Lets Talk Science
Program undertaken with high school students at the University of Victoria.
Experimental track re-creations have been used by a number of researchers to
help understand the processes by which ancient tracks were created from different
perspective. For example, Ruiz and Torices (2013) used experiments to help deter-
mine speed estimations for human trackways. Marty et al. (2009) created tracks in a
number of different contexts to help determine differences in morphology and
taphonomy.
5 Re-creation of Depositional Context of Footprints 95
Fig. 5.3 Examples of
images taken of tracks that
were excavated
96 D. McLaren et al.
Methods
The experimental re-creation of the tracks from EjTa-4 was conducted during
educational outreach sessions at the University of Victoria led by Duncan McLaren
and Quentin Mackie. The experiment was run a total of four times with the help of
high school students whose feet were used to make the footprints and University
student volunteers who helped with setting up and monitoring the experiments
(Fig. 5.4).
A clay matrix was specially prepared for this experiment. This clay included
some ne potters sand (less than 5%) to give it some stiffness. The clay was
pounded until consistent and then was placed in plastic totes to help keep it wet.
Small twigs and leaf fragments were then placed on top of the clay substrate. High
school students were then asked to volunteer to step into the clay to make a track.
The following task involved adding dark grey coarse sand to cover the track and clay
surface completely. The top of the sand was tapped gently to pack it down. The
subsequent excavation of these features was undertaken by trowel and spoon.
Fig. 5.4 Experimental
tracks created in clay during
workshop on footprints at
the University of Victoria
5 Re-creation of Depositional Context of Footprints 97
Results
The experiment was run a total of four times with four different high school student
groups. Before covering the tracks with sand, we noted that the experimental clay
matrices held the track impression well, being supported and shaped with the aid of
the elasticity of the clay matrix. Sediment displacement rims were clearly visible, but
were not necessarily created around the entire track. The feet of all of students were
covered in the clay matrices after having completed making the track suggesting that
at least some of the true track surface stuck to the bottom and sides of the foot.
Toe prints were visible in all cases, and some had sediment displacement rims
between the individual toes. Sediment displacement rims were most prominent
between the rst and second toes. In one case, toe drag marks were left, and in
another the heel had notably slipped towards the anterior.
With the addition of a layer of sand, the experimental tracks were rendered
buried. The subsequent excavation of these tracks revealed that the sediment dis-
placement rims were the rst part of each track encountered (Fig. 5.5). Through
further excavation, the remains of the true track could be revealed, providing a
feature that could be measured. The twigs and leaves that had been left on top of the
clay prior to impressing the track were impressed into the true track and in all
instances needed to be removed from the clayey substrate below. All plant macro-
fossils recovered were covered in the clay.
Fig. 5.5 Experiment track after excavation (left) compared with ancient track after excavation
(centre) and image contrast adjustment (right)
98 D. McLaren et al.
Discussion and Conclusions
The results of this experimental re-creation help inform us about the site formation
processes of the tracks at archaeological site EjTa-4. The clayey matrix in which the
experimental tracks were created was found to be an excellent substrate to create and
hold the impressions that were made. This appears to have been somewhat depen-
dent on the amount of moisture. By extension, with too much moisture, the tracks
would have been soupy and would have quickly disappeared, and with too little
moisture, a track impression would not have been as clearly made. From this we
have learned that the conditions at EjTa-4 would have been fairly damp, but not too
wet to create conditions whereby the tracks found were initially made. This is
consistent with our interpretation of the original depositional context being imme-
diately supratidal.
During the excavation of the tracks at EjTa-4, we found that the presence and
identication of the sediment displacement rims was key to the initial identication
of the tracks found. In the experimental footprints, sediment displacement rims
featured prominently in the tracks that were created and were prominent enough to
be the rst attribute encountered during excavation.
In the experimental re-creations, toe impressions were clearly visible. These
remained fairly distinct even after being covered with sand and then excavated.
Similarly, toe impressions were found during excavations at EjTa-4 suggesting
similarities in site formation process. Details such as these toe marks are important
as they reveal that we are dealing with true track impressions as opposed to
undertrack deposits which lack this type of detail (Marty et al. 2009).
Of particular importance to the chronological interpretation of the tracks at EjTa-
4 are the twigs that were found pressed into the true track surface. On the basis of the
stratigraphic position of these twigs and the associated radiocarbon dates, we were
able to assess that they were created between 13,317 and 12,633 calBP. The
replication of this situation in the lab with the experimental true track surfaces
having plant macrofossils pressed into them lends credence to our interpretation of
the track formation process and chronology.
Based on our ndings, we think that it is most likely that the tracks at EjTa-4 were
buried relatively quickly after they had been created. We are not certain how long the
track impressions would have lasted had they not been lled with sand. With an
increase of the amount of time that the tracks were exposed it is likely that they
would have washed to mush in a rain event, dried and desiccated beyond recognition
during a dry period, or eventually would have become over trampled by other
humans or animals. However, as the tracks were pressed into clay, it is possible
that they would have retained their shape longer than tracks pressed into sand. As
with most ancient track sites, it seems to be a fortuitous set of circumstances that
resulted in the preservation of the features at all.
Overall, the experiment re-creation of tracks lends support to our interpretation of
the site formation processes at EjTa-4. A minimum of three people left tracks just
above the high tide line 13,000 years ago. These tracks were then covered by sand
deposited in a high sea level event and remained capped since this time. While most
track sites are revealed to archaeologists through erosion, those discovered at EjTa-4
were found through careful excavation. A key to the successful excavation of these
tracks was the identication of sediment displacement rims which alerted the
excavators to the likelihood that a full track lay beneath.
5 Re-creation of Depositional Context of Footprints 99
Acknowledgements The high school participants in the Lets Talk Science workshops where the
experimental tracks were made are thanked. Katie Brynjolfson and Isabelle Rutherford, University
of Victoria undergraduate volunteers, are thanked for their help in helping with the experiment
reported on here. Stephanie Calce helped to organize the Lets Talk Science event for the
Anthropology Department at the University of Victoria. George Mackie is thanked for preparing
and pounding the clay and advice on sand and moisture content.
References
Bennett, M. R., & Morse, S. A. (2014). Human footprints: Fossilised locomotion? Cham: Springer
International Publishing. https://doi.org/10.1007/978-3-319-08572-2.
Bustos, D., Jakeway, J., Urban, T. M., Holliday, V. T., Fenerty, B., Raichlen, D. A., Budka, M.,
et al. (2018). Footprints preserve terminal Pleistocene hunt? Human-sloth interactions in North
America. Science Advances, 4(4), eaar7621eaar7621.
Clague, J., Harper, J., Hebda, R., & Howes, D. (1982). Late quaternary sea levels and crustal
movements, Coastal British Columbia. Canadian Journal of Earth Sciences, 19(3), 597618.
https://doi.org/10.1139/e82-048.
Dufeld, S. E. E. (2017). Long-term use of sh and shellsh resources revealed through vibracore
Sampling at EjTa-13, Hecate Island, Central Coast, BC. Unpubl. MA Thesis, University of
Victoria. https://dspace.library.uvic.ca//handle/1828/8936. Accessed 8 Jan 2020.
Fedje, D. W., & Mathewes, R. W. (Eds.). (2005). Haida Gwaii: Human history and environment
from the time of loon to the time of the iron people. Vancouver: UBC Press.
Fedje, D., McLaren, D., James, T. S., Mackie, Q., Mackie, A. P., Smith, N. F., & Southon, J. R.
(2018). A revised sea level history for the Northern Strait of Georgia, British Columbia, Canada.
Quaternary Science Reviews, 192, 300316.
Gauvreau, A., & McLaren, D. (2016). Stratigraphy and storytelling: Imbricating indigenous oral
narratives and archaeology on the Northwest Coast of North America. Hunter Gatherer
Research, 2, 303325.
Marty, D., Strasser, A., & Meyer, C. A. (2009). Formation and taphonomy of human footprints in
microbial mats of present-day tidal-at environments: Implications for the study of fossil
footprints. Ichnos, 16(12), 127142.
McLaren, D., Fedje, D., Hay, M. B., Mackie, Q., Walker, I. J., Shugar, D. H., Eamer, J. B. R., Lian,
O. B., & Neudorf, C. (2014). A post-glacial sea level hinge on the Central Pacic Coast of
Canada. Quaternary Science Reviews, 97, 148169.
McLaren, D., Fedje, D., Dyck, A., Mackie, Q., Gauvreau, A., & Cohen, J. (2018). Terminal
Pleistocene epoch human footprints from the Pacic Coast of Canada. PLoS One, 13(3),
e0193522e0193522.
Olson, R. (1955). Notes on the Bella Bella Kwakiutl. University of California Anthropological
Records, 14(5), 319348.
Ruiz, J., & Torices, A. (2013). Humans running at stadiums and beaches and the accuracy of speed
estimations from fossil trackways. Ichnos, 20(1), 3135.
100 D. McLaren et al.
Shugar, D. H., Walker, I. J., Lian, O. B. J., Eamer, B. R., Neudorf, C., McLaren, D., & Fedje,
D. (2014). Post-glacial sea-level change along the Pacic Coast of North America. Quaternary
Science Reviews, 97, 170192.
Smith, G. M., Parsons, T., Harrod, R. P., Holmes, C. E., Reuther, J. D., & Potter, B. A. (2019). A
track in the Tanana: Forensic analysis of a Late Holocene footprint from Central Alaska. Journal
of Archaeological Science: Reports, 24, 900912.
Walkus, S., Hilton, S. S., Walkus Windsor, E., & Rath, J. C. (1982). Oowekeeno Oral Traditions as
told by the Late Chief Simon Walkus, Sr. (Mercury Series No. 84). Ottawa: National Museums
of Canada.
Willey, P., Watson, P., Crothers, G., & Stolen, J. (2009). Holocene human footprints in North
America. Ichnos, 16,7075.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Chapter 6
Reading Spoor
Epistemic Aspects of Indigenous Knowledge and its
Implications for the Archaeology of Prehistoric Human
Tracks
Tilman Lenssen-Erz and Andreas Pastoors
Abstract The spoor of animals and humans alike contain rich information about an
individual and about a momentary activity this individual performed. If the
arguably hard-wired human ability to read spoor and tracks is sufciently trained,
a footprint allows to glean from it various physical, kinetic, medical, social and
psychologic data about an individual, as has been observed among various
populations across the globe. The Ju|hoansi San from northern Namibia still
today practice traditional hunting so that tracking is a skill that is required and
trained on a daily base. For a good tracker, the information she or he gets from spoor
is equally rich on animal and human footprints, and it is not necessary that the tracker
has been exposed before to the individual whose spoor she/he reads. In order to
allow an assessment of how tenable are the interpretations by contemporary hunter-
gatherers of prehistoric human footprints, this chapter elucidates methodological
aspects of tracking and situates this ability in an epistemological framework.
Keywords Hunter-gatherers · Tracking · Induction · Deduction · Abduction ·
Hypothetico-deductive reasoning · Tacit knowledge
T. Lenssen-Erz (*)
African Archaeology, University of Cologne, Cologne, Germany
e-mail: lenssen.erz@uni-koeln.de
A. Pastoors
Institut für Ur- und Frühgeschichte Friedrich-Alexander-Universität Erlangen-Nürnberg,
Erlangen, Germany
e-mail: andreas.pastoors@fau.de
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_6
101
Introduction
Human footprints are most prominent among the long-time under-researched fea-
tures of the context of cave art. In order to compensate for this neglect, a special
research programme has focused on the merging of indigenous knowledge and
Western archaeological science for the benet of both sides. With the expert
assistance of indigenous San hunters from the Namibian Kalahari, the Tracking in
Caves project endeavoured to better understand aspects of the Upper Palaeolithic
human behaviour, traces of which are preserved in footprints in painted caves in
southern France. The three professional indigenous trackers, Thui Thao, /Ui Kxunta
and Tsamgao Ciqae (e.g. Pastoors et al. 2015. Lenssen-Erz et al. 2018), were invited
to Europe and conducted in-depth research in the caves Niaux, Pech Merle,
Fontanet, Tuc dAudoubert and Aldène (Fig. 6.1 see Pastoors et al. Chap. 13).
102 T. Lenssen-Erz and A. Pastoors
The extent of preservation of footprints from the Pleistocene depends on advan-
tageous taphonomic circumstances and on careful treatment of the caves after
modern rediscovery. Accordingly Pech Merle has less than 20 footprints, but the
other caves each have at least several dozens and some several hundred. In Aldène
they were left behind by visitors during the Mesolithic, Niaux is insufciently dated,
all others are of Upper Palaeolithic origin.
The documentation of the interpretations of the indigenous ichnologists is indis-
pensable if they should be analysed archaeologically, but it is no less important to
cross-check the results obtained with the results of the studies of Western sciences.
The circumstance that the results of indigenous ichnologists are difcult to evaluate
does not imply that they are worthless assumptions. They have to be veried or
falsied with quantitative analyses and integrated into the discussion of prehistoric
human footprints.
Even though the art of tracking has already been described comprehensively
(Liebenberg 1990), there is a certain neglect of it in the current discourse of
archaeology on prehistoric human tracks. In order to allow due appreciation of the
Fig. 6.1 The San
ichnologists Thui Thao,
Tsamgao Ciqae and /Ui
Kxunta during their spoor
investigations in the cave of
Tuc dAudoubert
...
6 Reading Spoor 103
methodological foundations of indigenous ichnologies, this chapter focuses on the
art of tracking and its implications for the archaeology of prehistoric human tracks.
The Art of Tracking
With his ground-breaking book The Art of Tracking, the Origin of Science, Louis
Liebenberg (1990) opened up new perspectives on the profoundness and epistemo-
logical complexity of the indigenous knowledge of tracking. Having immersed into
the tracking culture of southern African San hunters, he understood that tracking is
an intricate edice of thought that stands a comparison to established sciences in
Western cultures:
In the narrowest sense of the word spoorsimply means footprint, but in tracking it has a
much wider meaning, including all signs found on the ground or indicated by disturbed
vegetation. Tracking also involves signs such as scent, urine and faeces, saliva, pellets,
feeding signs, vocal and other auditory signs, visual signs, incidental signs, circumstantial
signs, blood spoor, skeletal signs, paths, homes and shelters. Spoors are not conned to
living creatures. Leaves and twigs rolling in the wind, long grass sweeping the ground or
dislodged stones rolling down a steep slope leave their distinctive spoor. Markings left by
implements, weapons or objects may indicate the activities of the persons who used them,
and vehicles also leave tracks. [ ]
Spoor includes a wide range of signs, from obvious footprints, which provide detailed
information on the identity and activities of an animal, to very subtle signs which may
indicate no more than that some disturbance has occurred. [...] Signs of spoor may vary
considerably with terrain, weather conditions, season, time of day and age. (Liebenberg
1990: 111113)
Summing up the elds of knowledge that need to be mastered for successful
tracking shows that reading spoor goes far beyond pattern recognition (cf. Gagnol
2013: 175). Tracking requires detailed zoological knowledge (behaviour, seasonal
changes, reproduction, feeding habits, etc.) of the prey but also of animals in context
including small mammals, reptiles, insects, etc. They may provide additional infor-
mation if, e.g. a nocturnal animal walks through the spoor of a tracked animal, thus
indicating how old a seemingly fresh track may be. Also all topics of ecological
knowledge are part of the tracking skills with deep insights into biosphere and
geosphere as well as pedology regarding the inuence of different soil qualities on
the ageing of spoor. The same applies for meteorological knowledge and weather
observations that have to be memorized, e.g. in knowing which were the prevailing
wind directions in the past 24 h. On top of this, each tracker needs to have exact
knowledge of the place/area regarding vegetation, water points, game trails, salt
licks, etc., all of which may be points of orientation for movements of animals. But
also in his or her own interest, a tracker needs to have an absolute sense orientation
(e.g. Brenzinger 2008), rst to nd the way home and second to being able to
communicate spots in the landscape to others (e.g. the place where the carcass of a
hunted animal is lying).
104 T. Lenssen-Erz and A. Pastoors
Also the potential of tracking to identify individuals is explained by Liebenberg:
While species can be identied by characteristic features, there also exist individual varia-
tions within a species. These variations make it possible for an experienced tracker to
determine the sex as well as an approximate estimation of the animals age, size and mass.
A tracker may also be able to identify a specic individual animal by its spoor. [...]
The age of an animal may be indicated by the size of the feet. The hoofs of young
antelope will also have sharper edges, while old individuals may have blunted hoofs with
chipped edges. With animals with padded feet, younger individuals may have more rounded
pads. Some animals have specic breeding periods. If it is known at what time of year an
animal is born, a reasonably accurate estimate of its age can be made. [...]
Apart from features characteristic to the species, there also exist random variations within
the species which may vary from individual to individual.
The exact shape of every individual is unique so that it is, in principle, possible to
identify an individual animal. In practise this requires considerable experience, and is
usually only possible with large animals. With elephant and rhinoceros it is easy to identify
an individual by the random pattern of cracks underneath the feet.
The shape of feet may also be altered by environmental factors. In hard terrain, hoofs of
ungulates may be blunted by excessive wear, or in soft, sandy terrain, they may grow
elongated hoofs due to lack of natural wear. (Liebenberg 1990: 122124)
All which is said here on animal tracks is analogically found in human spoor
(e.g. Biesele and Barclay 2001; Lowe 2002; Gagnol 2013; see Gagnol Chap. 19)
since once the subtle reading is trained, it makes no difference to which type of trace
the skill is applied. Therefore trackers are able to interpret many other signs of
animals, e.g. where and how they were lying on the ground or if there were two
animals ghting and rolling over the ground skilled trackers will be able to recon-
struct complex sequences of movements and interaction. It also means that trackers
are able to follow the tracks of an individual be that person, game or herd animal
under changing soil conditions and even if mixed with imprints of other individuals
of the same species.
Apart from these elds of knowing that are implied in tracking, it was also
Liebenberg who pointed out that reading spoor means methodologically building
hypotheses based on empirical evidence and that these hypotheses are constantly
tested against ever new data (observations, perception; Liebenberg 1990: 153157).
Methodological Aspects of Tracking
As Liebenberg has described in detail tracking, i.e. reading tracks is a special skill
that is a precondition for human hunting and therefore may be considered the
beginning of science (Liebenberg 1990). As such, it is related to ichnology, the
science of tracks and traces, which originally was mainly occupied with fossil tracks
(such as of dinosaurs), but since the discovery of the earliest hominid footprints in
Laetoli (Tanzania) has also turned to humans (Lockley 1999). In current research of
prehistoric human footprints, Western science reveals essentially two approaches:
rst, footprint outline and landmark-based geometric-morphometric analyses
(e.g. Bennett et al. 2009,2016) and, second, pixel-based quantitative analysis of
the whole foot pressure (e.g. Crompton et al. 2011).
In pre-industrial societies of hunter-gatherers and herders, mastery of track
reading is an existential necessity. It is being learned from early childhood onwards,
requiring lifelong learning and constant practice. The reference to personal experi-
ences and personal exposure to the object of description is at the same time a
6 Reading Spoor 105
reference to the fact that tacit knowledge (Polanyi 1966) is required in tracking to
a considerable extent comprising knowledge and cognitive possibilities that cannot
be made explicit but rather are specically available to each individual through an
embodiment of experience.
Presently indigenous ichnology has a much ner resolution of tracks than modern
morphometric methods since trackers are normally able to determine from a foot-
print the sex and the approximate age class of a person, where the latter is not a factor
of body height but based on overall foot proportions and traces of ageing.
Indigenous ichnology is not based in rationality, logic or causalities that differ
drastically from Western views, as may be the case with traditional ecological
knowledge (TEK) (Berkes 2008: 8). Nevertheless, taking a series of scientic
measurements (e.g. Pales 1976; Webb 2007; Kinahan 2013; Ashton et al. 2014)is
an unsatisfactory substitute and cannot produce understanding, as opposed to read-
ing the ground for tracks (Chamberlin 2002). Expert tracking produces a narrative
that is based on in-depth knowledge of the entire ecosystem and its agents, acquired
through experience (Liebenberg 1990; Blurton Jones and Konner 1976; Lowe
2002). The capabilities of hunter-gatherers in reading tracks are legendary through-
out various types of literature (e.g. Marshall Thomas 1988; Liebenberg 1990;
Biesele and Barclay 2001; Lowe 2002), and no knowledgeable author leaves a
doubt regarding the reliability of the trackersskills. And also among traditional
herders, equally deep analysis of tracks is found (Gagnol 2013). But despite the
presence of prehistoric tracks on all continents (Lockley et al. 2008; Pasda 2013),
only very little, rather anecdotal use has been made of indigenous tracking knowl-
edge in archaeological contexts (Webb et al. 2006; Franklin and Habgood 2009).
As regards scientic scepticism about the reliability of spoor analyses by indig-
enous ichnologists, there have been empirical tests under controlled conditions with
very high rates of accurateness of 98% (Stander et al. 1997) or 74% inter-rater
reliability (Wong et al. 2011). The rst study tested a group of San trackers of which
Thao was part and the task was to determine for animal spoor the species, sex and
age class of the animal and how old the spoor was. The second study aimed at
determining whether spoor reading by Inuit hunters would be reliable enough for
collecting census data on polar bears which indeed was conrmed by the study.
Furthermore, the two main ichnologists of the present study (Kxunta and Thao) have
both passed the CyberTracker tracking certication (http://www.cybertracker.org/
downloads/tracking/CyberTracker-Tracker-Certication-2018.pdf) with accuracy
results of >90% (pers. comm. Liebenberg 2018).
If the method of tracking is analysed epistemologically, it is linked to Western
scienticthought by the intellectual procedures of inductive, deductive and
abductive reasoning (after C. S. Pierce 1955, cf. Liebenberg 1990;Eco and Seboek
1988) as three options to build hypotheses in the interpretation of observations
(Fig. 6.2). Abduction, for that matter, can be described as a process that:
¼
106 T. Lenssen-Erz and A. Pastoors
Fig. 6.2 Representation of
the tracking process after
Liebenberg (1990) with
three types of reasoning and
two principal paradigms of
nding out the whereabouts
of an animal or person
whereabouts
spoor
action
behaviour
animal
Systematic tracking Speculative tracking
empirical evidence / positivist paradigm
anticipation / initial interpretation of signs
abduction
induction
deduction
begins with observations and then proceeds in a back-and-forth process of developing
hypotheses and comparing the observations with information known and led in memory.
[...] Abductive reasoning then assembles the observations and attributes a variety of
characteristics or conditions to a subject until a match is made and an hypothesis or
conclusion can be stated. (Moriarty 1996: 181)
In following a spoor, the three methods of deriving conclusions, according to
Liebenberg, are realized in an inductive-deductive practice which he labels system-
atic tracking, and a hypothetico-deductive (or abductive) one, termed speculative
tracking (Fig. 6.2; Liebenberg 1990: 106108). The former method rests quite
narrowly with the observations the trackers make on the spoor they follow, thus
forcing them to walk the same way as the pursued animal walked. The latter method
makes an educated guess about what a pursued animal is going to do next on the
basis of information gathered from the spoor up to a given moment and on general
knowledge of the animals behaviour. Thus the trackers may leave the spoor and take
a shortcut to the place where the spoor is expected to be retrieved again. Gagnol,
building on his own tracking research among Saharo-Sahelian camel herders, also
found that successful tracking is importantly based on abductive method and what
Liebenberg calls speculative tracking resounds in what Gagnol terms stratégie
hodologique (Gagnol 2013: 172; Gagnol et al. 2018: 21; hodology study of
pathways; see Gagnol Chap. 19). The essence of both methods lies in the principle of
liberating the search process of tracking from the dependence of visible spoor but
6 Reading Spoor 107
instead making educated assumptions about what the subject may have chosen to
do. In Gagnols words:
One imagines the general path, the assumed goal of the animal or person and the means or
stratagems he or she will apply to achieve it (therefore one has to adopt his or her point of
view, putting oneself in the perspective of the other). (Gagnol 2013: 173; translation from
French TLE
1
)
In view of such intellectual processes, deriving conclusions from observations
Liebenberg emphasizes their complexity which is no less than that of modern
scientists, e.g. in physics or mathematics (Liebenberg 1990:4546). Accordingly,
upon thorough study of the character of tracking, authors have no doubt of the status
of tracking as analogous to science or as its forerunner (Liebenberg 1990; Jones and
Konner 1976; Chamberlin 2002). Ciqae, Kxunta and Thao assert that the decisions
of trackers who hunt together and their interpretation of spoor are both based on
constant exchange of opinion as well as on shared expert statements amongst the
trackers (see also Liebenberg 1990; Blurton Jones and Konner 1976; Biesele and
Barclay 2001). Therefore tracking can be accepted as a serious methodology in an
epistemological sense, and trackers are justiably labelled ichnologists since their
professional practice largely is the interpretation of positivist data through reason
and logic, based upon clearly determinable, repeatable methods. Further corrobora-
tion for this epistemological assessment of tracking is provided in the fact that for the
differentiation of, e.g. male and female footprints, trackers assess the same markers
and proportions on a foot as in orthopaedics or forensics (e.g. Robbins 1985; Reel
et al. 2010) (Fig. 6.3).
At this juncture it must not go unmentioned that in hunter-gatherer societies,
skills in tracking are not the exclusive knowledge of adult male hunters, and as
mentioned before these skills are not restricted to animal tracks but also include
human spoor (cf. Marshall Thomas 1988: 26; Biesele and Barclay 2001: 79; Lowe
2002: 18, 68; see Gagnol Chap. 19 and Gagnol 2013 for tracking skills of herders).
Fig. 6.3 Ways of taking
data from a footprint: grey
lines are usually measured in
orthopaedics and forensics;
the red line represent the
measures San ichnologists
assess. The circles
circumscribe areas that are
assessed in their totality, in
particular for age
estimations
1
In original text: On imagine le parcours général, le but supposé de lanimal ou de la personne et
les moyens ou stratagèmes quil ou elle mettra en œuvre pour y parvenir (il fout donc adopter son
point de vue, se placer dans la perspective dautrui).
And as is the case with all human abilities, not every hunter or every herder in a
given group is equally good in that skill (Liebenberg 1990; Gagnol 2013). There are
always some trackers who through talent, persistence and ambition reach levels of
108 T. Lenssen-Erz and A. Pastoors
mastery so that they can read spoor which would leave other members of their
groups helpless.
Gagnol (2013; see Gagnol Chap. 19) in his research among Touareg (Niger) and
Toubou (Chad) camel herders regarding their capability of reading spoor conrms
all ndings that have been reported by other authors on hunter-gatherers even
though he seems to be totally unaware of this literature. For these herders it is normal
to have extraordinarily ne resolution of reading tracks regarding human spoor and
regarding their domestic animals like camel, horse, cattle, donkeys, goats and sheep
(Gagnol 2013: 171). Gagnol also points out that this expertise goes together with a
rich vocabulary for the description of spoor details and of ways of walking (Gagnol
2013: 170). For a camel herder, it is important to know all his animals by their spoor,
even if they may mix with another herd (which does not happen infrequently), and if
they get astray, the herder will occasionally track it for several days (Gagnol
2013: 170).
As regards human spoor, Gagnol asserts that every individual of a community can
be identied by her or his footprint, and also strangers are recognized due to their
unknown imprints. Such identication is not only based on morphological features
but also on details of the habitual gait of a person and wearing sandals is not an
impediment for such identication (Gagnol 2013:171). In following thieves the best
trackers are even able to track the culprit if he changes his shoes several times during
his escape (one of the strategies of camel thieves to complicate pursuit; Gagnol 2013:
179; see Gagnol Chap. 19). Information imprinted through a human footprint is so
rich that even social status or ethnic afliation can be gleaned from spoor (Gagnol
2013: 176; see Gagnol Chap. 19).
The described capability of extracting information from footprints was the basis
for analyses of human footprints in the Palaeolithic caves. The results of the
indigenous ichnologists compiled in the course of the various studies are as detailed
and precise as the tracking by masters would promise, and they go beyond the results
produced by Western science. This fact is perceived and reacted to in different ways
by the public. If exposed to the results, one part of the public shows scientic
curiosity, wishing to learn more about the capabilities of indigenous ichnologists
and to verify or falsify the results through their own investigations. Others, however,
show great scepticism to the extent of rejection. Such rejection does not seem
appropriate without empirical falsications, because indigenous ichnologists have
veried skills in reading tracks. Strictly speaking, their methodological approach is
not so alien to Western scientic approaches. Even though Erwin Panofskys
iconographic interpretation method refers to images, there are parallels between
reading tracks and interpreting images. According to Panofsky, in the case of a
natural subject as the object of interpretation, a pre-iconographic description of the
motifs takes rst place (Panofsky 1962). Practical experience (familiarity with
objects and phenomena) is an absolute prerequisite for a successful description,
from which a positive correlation between experience and descriptive accuracy can
be derived. In the event that the spectrum of personal experience is not sufcient, this
spectrum must be extended by consulting publications or experts. Practical experi-
ence, in turn, helps to determine which publication or professional is to be consulted
6 Reading Spoor 109
(Panofsky 1962: 9). This practical experience results not least from personal expe-
riences in the world in which we live and which:
provides the ground for all cognition and for all scientic purpose. (Husserl 1939: 38)
The concept of the lifeworld that is evoked here describes the realm of reality in
which every human being inevitably participates (Schütz and Luckmann 1975). This
concept asserts that there is a world of common, everyday experiences and interpre-
tations on which all more theoretical knowledge is dependent (Schütz and
Luckmann 1975: 23). A basic characteristic of the everyday lifeworld is its inter-
subjectivity, by which it forms a social world in which practically all members of a
social body take part with roughly the same interpretations of daily phenomena
(Schütz and Luckmann 1975: 33). The everyday lifeworld, seen as the most common
and widest accepted kind of reality, comprises physical objects, nature and the
everyday social world (Schütz and Luckmann 1975:41).
Accordingly we contend that many processes and phenomena in the empirical
world out there are understandable irrespective of the cultural imprinting an observer
has. Based on this lifeworld concept, we regard phenomena like spoor as providing
information on an implicit and an explicit level, the understanding of which is
informed by tacit and by explicit knowledge (after Polanyi, e.g. 1966). While the
implicit information is entirely embedded in the respective culture, or, as tacit
knowledge, even within an individual (and therefore largely inaccessible to us,
Polanyi 1966, see also Schütz and Luckmann 1975
Luckmann 1975:
Implications for the Archaeology of Prehistoric Human
101). As has been pointed out before, everything that can be said
on animal spoor also pertains to human footprints.
:99102), the explicit informa-
tion is based in intersubjective experiences in the empirical world. Reading animal
tracks cannot be separated from the actual behaviour of that species which the
animals perform irrespective of any cultural representation and symbolization of
this behaviour. Observing animal behaviour and the tracks it produces is a general
human experience and is based on positivist, empirical data while the sense that is
interpreted into such experience is subjective and culture-bound (Schütz and
Tracks
In archaeology, already at the beginning of the twentieth century, certain perplexity
and a lack of experience with regard to the reading of tracks in the interpretations of
prehistoric footprints in caves by Western scientists became apparent (e.g. Bégouën
1928;Lemozi 1929). Thus, not the recognition of a specic sequence of footprints
gave reason to interpretations as ritual dance (de Contenson 1949) but the transfer of
the generally perceived ritual status of the surrounding cave to the footprints. At this
point of archaeological analysis, lacking the capability to read tracks was masked by
the professional practice of interpreting cultural-historical processes. Even today not
only the lack of practical experience in reading tracks proves to be problematic but
also the existence of methodological limits of modern analytical procedures. This is
particularly evident in the difculty of making the sequence of steps of one and the
110 T. Lenssen-Erz and A. Pastoors
same person morphometrically visible, even though it is obviously from one person
(Bennett and Morse 2014).
Against the background of interpretations favouring ritual activities, the results of
the indigenous ichnologists appear unspectacular. But they ll a vacuum of descrip-
tion with content. An accumulation of footprints on a spatially limited area seems
like a chaotic mixture. This confusion dissolves when the indigenous ichnologists
combine footprints into sequences of steps of single individuals. For this purpose it is
indispensable that age, sex and individual characteristics of a person can be gleaned
from the footprint even if this is not possible with all of the extant Pleistocene
footprints. The demographic data is ultimately established by using morpho-
classicatory factors, which are essentially based on the same features that Western
science also uses.
Contextual information is also included in the track data acquisition: the nature of
the ground, room height, inclination, gradient, curvature, possible obstacles and
much more. The identied footprints are mapped and recorded in a data sheet. As
mentioned above, the result of the work of the indigenous ichnologists is not an
inventory of all footprints but of footprints about which they can give dependable
information. In this way, the indigenous ichnologistsapproach differs from that of
Western science, in which each individual footprint is recorded by using specic
attribute systems thus favouring description over interpretation. Ultimately, the
discussions about these two approaches are comparable to the dichotomy in archae-
ology between two well established methods of object analysis: the static attribute
analysis and the dynamic analysis of the chaîne opératoire. Each footprint is the
result of a unique interplay of bones, muscles and various other external factors and
represents therefore a non-repeatable event. But footprints, or human tracks in
general, are not alone with this situation in archaeological research. Every archae-
ological object is the result of certain constellations of internal and external factors
that cannot be reproduced accurately. Archaeological research has responded to this
dilemma by developing dynamic methods of investigation, including the chaîne
opératoire.
Another dynamic method is the indigenous knowledge of tracking; therefore its
application is not a matter of romanticism, and it is not aimed at providing an exotic
view of tracks from another world-view. Rather, it provides alternative interpreta-
tions of data, using the same empirical base that is accessible to any method
(Liebenberg 1990; Lockley 1999; Lowe 2002). To the present knowledge, there is
no other method for a deep understanding of spoor as remains of dynamic actions
that is equally successfully applicable to tracks in all kinds of substrate and in all
stages of taphonomic degradation. Western science responds to this situation by
applying experimental archaeology in order to develop a dynamic method (e.g. see
Ledoux et al. Chap. 4).
In fact, the interpretations of human footprints by Ciqae, Thao and Kxunta
achieve levels of precision described by Liebenberg and Gagnol. This prompts
questions as to which aspects of the footprint are signicant for such detailed
information. Liebenberg compiles different aspects of the spoors, which serve as a
base for the determination of age and sex, size, depth, way of movement, body
structure and association with other footprints, all of which is supported by Gagnols
observations. Ciqae, Thao and Kxunta corroborate that a male foot looks stronger
and wider than a female foot, indicating that, of course, an intuitive assessment of
proportions is the foundation of sex determination besides gait and step length.
According to Liebenberg (1990), wear, foot tension and again size are signicant for
age determination, which paraphrases the criteria mentioned and judged by the
trackers. Furthermore, Liebenberg noted that the exact shape of every individual is
unique, and, therefore, it is possible to identify individual animals and also humans
(the same is maintained for Australian aboriginal people by Lowe 2002 and for
Saharan nomads by Gagnol 2013). This, too, is substantiated by Ciqae, Thao and
Kxunta, who assert that, in particular, the shape of the toes and the way a foot is set
on the ground help them to identify their family, neighbours and friends by their
footprints. Also, age determination of a known or unknown person, so Ciqae, Thao
and Kxunta afrm, is largely based on judgments of the features of heels and toes,
plus a persons way of walking, since steps become shorter as a person grows old
(see also Gagnol 2013 for corroboration). Through this ne-grained differentiation,
they are able to distinguish different age classes among adults, even though mature
feet do not continue to grow. According to these trackers, the heels become harder
and more cracked the older a person gets, and also the toes become harder. Through
this, so the indigenous ichnologists maintain, the soil is being thrown up by the toes
in a different way by an old person than by a younger adult. Gagnol (2013: 171172)
6 Reading Spoor 111
describes analogous changes from young adults to mature adults among the animals
of the Toubou and Tuareg.
In a critical appreciation of the implementation of indigenous ichnology in
archaeology, it has to be conceded that there may be some inuential factors that
could generate possible biases. After all the original context of the spoor that the
indigenous ichnologists were asked to read stems from a period, environmental
conditions and population that were all entirely alien to the trackers previous
experiences. When addressing this problem with the ichnologists, they maintained
that people are people and reading the tracks of complete strangers is not uncommon
for them. Nevertheless, the following questions are some of those that may arise:
Which data are collected?
How do participants communicate?
Can technical terms be translated?
Do means of control apply (verication/falsication)?
Are there repeatable results?
Is there a second opinion?
How indigenous is indigenous knowledge?
First, it has to be emphasized that there is a general counter balance to these biases
by the practice of the San trackers which, again, shows that their approach to data
(tracks in this case) has much in common with a Western view of scientic
investigations (this list is based upon observations and interviews during common
eld work):
Empirical approach
Meticulous exactness
Best-practice ethics
Constant testing of hypotheses
Shared expert opinion
112 T. Lenssen-Erz and A. Pastoors
Prepared for constant learning
Immediate transfer and implementation of new experiences
Secondly and m
ore specically, data acquisition always takes place with intense
communication among the indigenous ichnologists and with the archaeologists. In
particular the internal exchange between the indigenous ichnologists is a guarantee
that all results that are stated are based on the inclusion of at least a second expert
opinion. Repeated visits to some spoor elds in French caves showed in a random
test that interpretations of imprints were the same after 3 years so that the results
indeed are repeatable even if conceding that the same persons interpreted the spoor
on both occasions. It is also important to emphasize that the data that are collected for
the Tracking in Caves project are the same as those which interest a tracker also
outside the research scheme: what are the characteristics of that person who left a
spoor and do I know her or him, where did she/he go, and what was her/his state of
mind and maybe her/his intentions (cf. Gagnol et al. 2018: 20). Therefore the
questions arising from the research are fully understandable to trackers, while they
do not mind the ultimate consequences of their spoor identications. Notwithstand-
ing this initial focus on every footprint in isolation, the analyses in the caves never
produced contradictory actions or behaviour. For example, in cases of footprint
superimpositions, the younger spoor would always be the one leading out of
the cave.
The indigenous ichnologists admit, however, that normally they would only be
interested in fresh spoor because tracking is a behaviour that generates information
for immediate action which is futile regarding old spoor.
With regard to due scienticdoubt about the initial results of spoor reading, there
is no independent, more reliable scientic method available, and every new inter-
pretation by other trackers of once interpreted spoor would constitute just another
opinion but not a verication or falsication.
Whether the analyses of the San trackers depend on specic terminology in
Ju|hoansi language for which there may be no equivalent in English is still a
desideratum of research, but rst investigations in other San languages clearly
point into this direction (e.g. Sands et al. 2017).
Finally the question of how indigenous the indigenous knowledge really is has no
relevance for the research questions. The reading of Pleistocene human footprints
aims at getting the maximum possible information from footprints, no matter in
which way the interpreting experts acquired their knowledge. But it is only in the
context of a life in rural areas where free roaming animals have outstanding
6 Reading Spoor 113
economic signicance where tracking is required and trained to such an extent that
the most skilled individuals attain world class knowledge.
The Wider Potential of Tracking
The extraordinary abilities of indigenous ichnologists can also be applied in other
elds of archaeological research. Particularly the ancient rock art of prehistoric
hunter-gatherers may become a promising study object since in many regions around
the globe, there are traditions of prehistoric art where animal tracks are an integral
part of the rock art motif spectrum (Lenssen-Erz et al. forthcoming for an overview;
Fig. 6.4). Considering that these depictions were produced by artists with the
mindset of hunter-gatherers, it is obvious that reading and interpreting them is best
attempted by hunter-gatherers themselves. Two pioneering eld studies in Namibia
indeed showed that in hundreds of engravings of animal spoor, indigenous
ichnologists were not only able to identify in almost all cases the exact animal
species (Nankela 2017) but also to determine sex and age class of an animal as well
as which of the four legs was depicted (Lenssen-Erz et al. forthcoming). The latter
study established three main ndings: rst the prehistoric hunter artists did not think
in generic categories by producing an exemplary track of, e.g. giraffe as is found in
eld guide books for spoor identication. Rather for each depiction a hunter artist
would conceive of a specic animal, e.g. a young male, and of this animal it would
be a particular leg that was represented by the engraved spoor.
Secondly the spectrum of species represented by spoor is much richer than the
spectrum of animals being depicted as gures, with a fair number of rather small
animals. Also the frequency of predators, especially various species of felines, is
much higher. Apparently the spoors of animals cover a different eld of symboliza-
tion than the depictions of animal silhouettes.
Fig. 6.4 San ichnologists
Ciqae, Kxunta and Thao
reading spoor in prehistoric
engravings in central
Namibias Doro !nawas
region
And thirdly it emerged that instead of producing a random distribution of all
possible features across all species that form part of the art canon, each species
shows a clear bias towards a sex, an age class and a particular leg that is preferably
depicted. The patterns that emerge cannot yet be interpreted because they produce
alliances that are not self-explicatory: for example, the features of being predomi-
nantly female is shared by leopard and guinea fowl, while zebra and duiker are
predominantly male; walking direction of almost all animals is up the wall, but
114 T. Lenssen-Erz and A. Pastoors
duiker and springbok predominantly walk down the wall. Whether such associations
have a common cognitive or symbolic base cannot be determined yet.
However, there are several reasons why it is necessary for a tracker to being able
to determine sex, age and which of the four legs of an animal are indicated by a
spoor. In the rst place, a hunter must be able to identify an individual animal within
a herd in order to be sure which animal he is tracking and hunting. Secondly the
hunter must be able to identify each leg of an animal separately, on the one hand to
understand the individual gait of every animal and on the other hand to see on which
leg it may be lame, e.g. through the impact of the arrow he launched. Thirdly, if
coming upon the spoor of a herd or pack of animals, it is important to quickly get an
overview of how many animals are in this herd or pack not only for hunting but
also for security reasons: it requires different levels of alertness and caution if one
comes upon the fresh spoor of lion whether these were left behind by two lions
walking up and down in a place or whether these spoor stem from a pack of nine
lions (pers. comm. O. Vogels 2019).
There is another eld of indigenous knowledge that is connected to tracking
which may be applicable in archaeology, notably rock art research. Indigenous
ichnologists are capable to reconstruct from tracks the behavioural body postures
of animals, be that game or domesticated animals, and therefore they are also well
trained ethologists. Various rock art traditions, be they created by hunter-gatherers or
herders, depict animals of signicance for that culture in rich variations of body
postures and gregarious congurations. Since archaeologists lack the training of
reading spoor as much as the training of reading behaviour, their interpretation of
animal behaviour depicted in rock art can only be rather supercial (e.g. Thackeray
1983; Lenssen-Erz 1994; Hollmann 2005). Involving the ethological knowledge of
hunter-gatherers or pastoralists respectively to the specicart traditions made by
their forerunners will certainly open new elds of meaning to these art corpora.
Conclusion
While giving tracking as indigenous knowledge a centre-stage position, this chapter
does not aim at providing a critical review of the concept of indigenous knowledge
since there is a broad literature on this subject (for an overview Odora Hoppers
2002). Indigenous knowledge shares at least a semantic eld if it is not identical with
terms such as traditional knowledge, local knowledge, civic science, traditional
ecological knowledge, community archaeology, etc. Flaws of the indigenous
knowledge concept are sometimes discussed due to its association with notions such
as nativism, essentialism, ethnicity of knowledge, ahistorical knowledge, (romantic)
notion of being static and bound, belief vs. analysis, tradition vs. modernity or
knowledge as a result of power relations. While being aware of this discourse, we
need not make a contribution to it. We do not imply that the indigenous knowledge
we work with necessarily has to be a pristine knowledge, but instead we involve San
ichnology because it is the highest standard we can get today and it promises to get
the maximum possible information that can be retrieved from an imprint for which
there are not yet any equally yielding machine-based analytical devices.
6 Reading Spoor 115
Further research is necessary to determine the smallest analytical steps of the
methodology applied by Ciqae, Thao and Kxunta to each single imprint. We do have
some indications about the procedure of extracting data from a footprint, manifested
in the sections of a foot that are part of the rating process and which are also used in
orthopaedics or forensics (Fig. 6.3). But this can only be the start and in order to
collect rst data on this topic, the entire determination process in each cave was
recorded as audio protocols. The transcription and translation of the discourses of the
trackers will serve as an important resource in the future for the deeper understanding
of indigenous ichnology.
In the course of the two eld studies 2013 and 2018, especially the last, there were
no inconsistencies or contradictions in the interpretation of the approximately 1000
prehistoric human footprints examined. This expresses the professionalism and
quality of the work of the indigenous ichnologists.
Currently preparations are under way to analyse most of the footprints examined
by indigenous ichnologists, using quantitative methods in order to verify or falsify
their results in the form of a cross-test with new methods. It will be a central task to
combine Western science with indigenous ichnology and to discuss the results of
both approaches. A main problem is certainly the difculty of evaluating the results
obtained by the indigenous ichnologists. But their integration into the interpretation
of prehistoric footprints is still in its infancy. Following the principle of Aristotle,
where the whole is more than the sum of its parts, indigenous ichnologists include
the behaviour of the trackmaker in their interpretations from the beginning. For
Western scientists, footprints are individual morphological features that are quanti-
ed as such, disconnected from the other footprints. The question of the behaviour of
the trackmaker is handled separately and comes at the end of the statistical analyses.
This methodical contrast offers great potential and promises benets for both sides.
For the future, these experiences mean that by applying indigenous knowledge,
selected source genres of archaeology can be explored in greater depth than would
be the case with conventional methods alone. While the case described here was
about the knowledge of hunter-gatherers, it is to be expected that the addition of, for
example, pastoral nomads in other elds of research and for other epochs will also
lead to new and deeper insights.
116 T. Lenssen-Erz and A. Pastoors
References
Ashton, N., Lewis, S. G., Groote, I. D., Duffy, S. M., Bates, M., Bates, R., Hoare, P., Lewis, M.,
Partt, S. A., Peglar, S., Williams, C., & Stringer, C. B. (2014). Hominin footprints from early
Pleistocene deposits at Happisburgh, UK. PLoS One, 9(2). https://doi.org/10.1371/journal.pone.
0088329.
Bégouën, H. (1928). Les empreintes de pieds préhistoriques (pp. 323328). Amsterdam: Congrès
International dAnthropologie et dArchéologie Préhistorique III.
Bennett, M. R., Harris, J. W. K., Richmond, B. G., Braun, D. R., Mbua, E., Kiura, P., et al. (2009).
Early hominin foot morphology based on 1.5 million year old footprints from Ileret, Kenya.
Science, 323, 11971201.
Bennett, M. R., & Morse, S. A. (2014). Human footprints: Fossilised locomotion? Cham: Springer
International Publishing. https://doi.org/10.1007/978-3-319-08572-2.
Bennett, M. R., Reynolds, S. C., Morse, S. A., & Budka, M. (2016). Laetolis lost tracks. 3D
generated mean shape and missing footprints. Scientic Reports, 6, 21916. https://doi.org/10.
1038/srep21916.
Berkes, F. (2008). Sacred ecology: Traditional ecological knowledge and resource management.
New York: Routledge.
Biesele, M., & Barclay, S. (2001). Ju|hoan womens tracking knowledge and its contribution to
their husbandshunting success. African Study Monographs, (Suppl. 26), 6784.
Blurton Jones, N., & Konner, M. J. (1976). !Kung knowledge of animal behavior: Or: The proper
study of mankind is animals. In R. B. Lee & I. DeVore (Eds.), Kalahari Hunter-Gatherers
(pp. 325348). Cambridge, MA: Harvard University Press.
Brenzinger, M. (2008). Conceptual strategies of orientation among Khwe: From sunrise/sunset
bisections to a left/right opposition. In S. Ermisch (Ed.), Proceedings of the 2nd international
symposium on Khoisan languages and linguistics, January 812, 2006, Riezlern/Kleinwalsertal
(pp. 1547). Köln: Rüdiger Köppe.
Chamberlin, J. E. (2002). Hunting, tracking and reading. In J. Brockmeier, M. Wang, & D. R. Olson
(Eds.), Literacy, narrative and culture (pp. 6785). Richmond: Curzon.
de Contenson, H. (1949). Empreintes de pieds humains préhistoriques dans la grotte de Niaux
(Ariège). Bulletin de la Société Préhistorique Française, 51(6), 214.
Crompton, R. H., Pataky, T. C., Savage, R., DAoût, K., Bennett, M. R., Day, M. H., et al. (2011).
Human-like external function of the foot, and fully upright gait, conrmed in the 3.66 million
year old Laetoli hominin footprints by topographic statistics, experimental footprint-formation
and computer simulation. Journal of the Royal Society, Interface, 9(69), 707719. https://doi.
org/10.1098/rsif.2011.0258.
Eco, U., & Seboek, T. A. (Eds.). (1988). The sign of three: Dupin, Holmes, Peirce. Bloomington:
Indiana University Press.
Franklin, N. R., & Habgood, P. J. (2009). Finger markings and the Willandra Lakes footprint site,
South-Eastern Australia. Rock Art Research, 26(2), 199203.
Gagnol, L. (2013). Identier, rechercher et surveiller par les traces: Éléments danalyse dun savoir-
faire pastoral dans les sociétés saharo-sahéliennes. Technique & Culture, 61, 166187.
Gagnol, L., Mounet, C., & Arpin, I. (2018). De la piste animal aux lignes de désir urbaines. Une
approche géoichnologique de la trace. LInformation géographique, 82(2), 1138.
Hollmann, J. (2005). Using behavioural postures and morphology to identify hunter-gatherer rock
paintings of therianthropes in the Western and Eastern Cape provinces, South Africa.
South African Archaeological Bulletin, 60(182), 8495.
Husserl, E. (1939). Erfahrung und Urteil. Untersuchungen zur Genealogie der Logik. Prag:
Academia.
Kinahan, J. (2013). The use of skeletal and complementary evidence to estimate human stature and
identify the presence of women in the recent archaeological record of the Namib desert.
South African Archaeological Bulletin, 68(197), 7278.
6 Reading Spoor 117
Lemozi, A. (1929). La grotte-temple du Pech-Merle. Un nouveau sanctuarie préhistorique. Paris:
Picard.
Lenssen-Erz, T. (1994). Jumping about - Springbok in the Brandberg rock paintings and in the
Bleek/Lloyd collection. An attempt at a correlation. In T. Dowson & J. D. Lewis-Williams
(Eds.), Contested images Diversity in Southern African rock art research (pp. 275291).
Johannesburg: Witwatersrand University Press.
Lenssen-Erz, T., Pastoors, A., Ciqae, T., Kxunta, U., Thao, T., Bégouën, R., Biesele, M., & Clottes,
J. (2018). Tracking in caves: Reading human spoor in ice age caves with San hunters. In R. F.
Puckett & K. Ikeya (Eds.), Research and activism among the Kalahari San Today: Ideals,
challenges, and debates (Senri ethnological studies 99) (pp. 103127). Osaka: National
Museum of Ethnology.
Lenssen-Erz, T., Pastoors, A., Roth, G., Ciqae, T., Kxunta U., Thao T. (forthcoming) Animal spoor
in prehistoric hunter-gatherer rock art analysed by contemporary hunter-gatherers.
Liebenberg, L. W. (1990). The art of tracking: The origin of science. Claremont: David Philip.
Lockley, M. (1999). The eternal trail: A tracker looks at evolution. Reading: Perseus Books.
Lockley, M., Roberts, G., & Kim, J. Y. (2008). In the footprints of our ancestors: An overview of
the hominid track record. Ichnos, 15, 106125.
Lowe, P. (2002). Hunters and trackers of the Australian desert. Surry Hills: Accessible Publishing
Systems.
Marshall Thomas, E. (1988). The harmless people. Cape Town: David Philip.
Moriarty, S. E. (1996). Abduction: A theory of visual interpretation. Communication Theory, 6,
167187.
Nankela, A.M. (2017). Rock art and landscape: An empirical analysis in the content, context and
distribution of the rock art sites in Omandumba East and West, Erongo Region Namibia.
Unpublished PhD dissertation, Universidade de Tras-Os-Montes e Alto Douro.
Odora Hoppers, C. A. (Ed.). (2002). Indigenous knowledge and the integration of knowledge
systems. Claremont: New Africa Education.
Pales, L. (1976). Les empreintes de pieds humains dans les cavernes: Les empreintes du resau nord
de la caverne de Niaux (Ariège) (Archives de lInstitut de Paleontologie Humaine, 36). Paris:
Masson.
Panofsky, E. (1962). Studien zur Ikonologie. Humanistische Themen in der Kunst der
Reanaissance. Köln: DuMont.
Pasda, C. (2013). Zu Fuss Spurenfossilien des eiszeitlichen Menschen. In A. Pastoors &
B. Auffermann (Eds.), Pleistocene Foragers: Their culture and environment: Festschrift in
honour of Gerd-Christian Weniger for his Sixtieth Birthday (pp. 5170). Mettmann: Neander-
thal Museum.
Pastoors, A., Lenssen-Erz, T., Ciqae, T., Kxunta, U., Thao, T., Bégouën, R., Biesele, M., & Clottes,
J. (2015). Tracking in caves: Experience based reading of Pleistocene human footprints in
French caves. Cambridge Archaeological Journal, 25(3), 551564.
Peirce, C. S. (1955). Philosophical writings of Peirce (J. Buchler, ed.). New York: Dover.
Polanyi, M. (1966). The logic of tacit inference. Philosophy, 61(155), 118.
Reel, S., Rouse, S., Vernon, W., & Doherty, P. (2010). Reliability of a two-dimensional footprint
measurement approach. Science and Justice, 50, 113118.
Robbins, L. M. (1985). Footprints: Collection, analysis, and interpretation. Springeld: Charles
C. Thomas.
Sands, B., Chebanne, A., & Shah, S. (2017). Hunting terminology in ǂHoan. In A. M. Fehn (Ed.),
Khoisan languages and linguistics proceedings of the 4th international symposium July 1113,
2011, Riezlern/Kleinwalsertal (pp. 185214). Köln: Rüdiger Köppe.
Schütz, A., & Luckmann. T. (1975). Strukturen der Lebenswelt. Darmstadt: Luchterhand. English
edition: Schutz, A., & Luckmann, T. (1973). The structures of the life-world. Evanston:
Evanston Northwestern University Press.
Stander, P. E., Ghau, I. I., Tsisaba, D., OMA, I. I., & | | UI. (1997). Tracking and the interpretation of
spoor: A scientically sound method in ecology. Journal of Zoology, 242, 329341.
118 T. Lenssen-Erz and A. Pastoors
Thackeray, J. F. (1983). Disguises, animal behaviour and concepts of control in relation to rock art
of Southern Africa. South African Archaeological Society Goodwin Series, 4,3843.
Webb, S. (2007). Further research of the Willandra Lake fossil footprint site, southeastern Australia.
Journal of Human Evolution, 52, 711715.
Webb, S., Cupper, M. L., & Robins, R. (2006). Pleistocene human footprints from the Willandra
Lakes, southeastern Australia. Journal of Human Evolution, 50(4), 405413.
Wong, P. B. Y., Groot, V. C., & P.de, Fekken, C., Smith, H., Pages, M., Boag, P.T. (2011).
Interpretation of polar bear (Ursus maritimus) tracks by Inuit hunters: Inter-rater reliability and
inferences concerning accuracy. The Canadian Field-Naturalist, 125, 140153.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Part II
Case Studies from Around the Globe
Chapter 7
Perspectives on Pliocene and Pleistocene
Pedal Patterns and Protection
Implications for Footprints
Erik Trinkaus, Tea Jashashvili, and Biren A. Patel
Abstract As a framework for interpreting Pliocene and Pleistocene hominin foot-
prints, the functional implications of australopith and Homo pedal remains are
reviewed. Despite minor variations in pedal proportions and articular morphology,
all of these remains exhibit tarsometatarsal skeletons fully commensurate with an
efcient (human) striding bipedal gait. The Middle and Late Pleistocene Homo pedal
phalanges exhibit robust and distally attened metatarsal 1 heads, hallux valgus,
relatively short lateral digits with largely straight proximal phalanges with dorsally
oriented metatarsal facets, all similar to those of recent humans. The Pliocene and
Early Pleistocene halluces lack hallux valgus and have bulbous metatarsal 1 heads.
The australopith pedal remains have lateral proximal phalanges that are relatively
long and dorsally curved and have more proximally oriented metatarsal facets. In
addition, pre-Upper Paleolithic Homo lateral phalanges have robust diaphysis imply-
ing the habitual absence of protective footwear, whereas the Upper Paleolithic ones
are variably gracile, especially at higher latitudes, indicating more consistent use of
footwear. These paleontological considerations provide a framework for interpreting
the distal portions of earlier hominin footprints (especially with respect to hallucal
orientation and digital length) and suggest that many of the Late Pleistocene
footprints may be unrecognized given the use of footwear.
E. Trinkaus (*)
Department of Anthropology, Washington University, Saint Louis, MO, USA
e-mail: trinkaus@wustl.edu
T. Jashashvili
Department of Integrative Anatomical Sciences, Keck School of Medicine,
University of Southern California, Los Angeles, CA, USA
Department of Geology and Paleontology, Georgian National Museum, Tbilisi, Georgia
e-mail: tea.jashashvili@usc.edu
B. A. Patel
Department of Integrative Anatomical Sciences, Keck School of Medicine,
University of Southern California, Los Angeles, CA, USA
e-mail: birenpat@usc.edu
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_7
121
Keywords Paleoanthropology · Feet · Toes · Shoes · Phalanges · Hallux ·
Australopithecus ·Homo
Introduction
122 E. Trinkaus et al.
The human lineage has evolved a pedal anatomy that facilitates an efcient striding
bipedal gait. As the interface between the body and the substrate during gait, it is also
the portion of the anatomy which is primarily responsible for the form and variation
in footprints. Given that hominins have had a basically bipedal pelvic and leg
anatomy since at least the early members of Australopithecus (Ward 2013), it is
likely that variation in footprints would reect a complex combination of pedal
anatomy and the behaviours imposed on the foot. This short review is therefore
intended to provide an overview of Pliocene and Pleistocene human pedal anatomy
and variation, with respect to their implications for assessing footprints from the
past. Particular focus is placed on the pedal digits, given the stability of the human
tarsometatarsal skeleton once it became basically humanlike (or bipedal) in the
earlier Pliocene (DeSilva et al. 2019).
The paleontological record for human foot evolution consists of isolated remains
and a dozen partial pedal skeletons for the earliest phases, several of uncertain
taxonomic afliation. Middle Pleistocene associated feet derive from Dinaledi and
Atapuerca-SH, there are half a dozen largely complete Middle Paleolithic pedal
skeletons and then a relative abundance of them in the Upper Paleolithic. Only in the
Middle and Upper Paleolithic, plus one Australopithecus specimen, are the pedal
remains from associated skeletons. Therefore, for the Pliocene and Early Pleisto-
cene, overall pedal anatomy is based on composites, often from diverse sites,
whereas the later periods permit assessments from single individuals (DeSilva
et al. 2019; Fig. 7.1). Isolated remains nonetheless ll out the record. The pedal
remains from Aramis, Burtele and Liang Bua are not considered here, given their
divergent congurations and their lack of association with footprints.
Individual points are not referenced in the discussion. For overall assessments,
some of the key or more complete specimens and key aspects of the discussion, see
Latimer et al. (1982), Susman (1983), Trinkaus (1983,2005), Lordkipanidze et al.
(2007), Zipfel et al. (2011), Ward (2013), Trinkaus et al. (2014,2017), Harcourt-
Smith et al. (2015), Trinkaus and Patel (2016), Pablos et al. (2017), Fernández et al.
(2018), McNutt et al. (2018) and DeSilva et al. (2019). For the earlier phases,
DeSilva et al. (2019) provide an extensive review; for the later phases, see especially
Trinkaus (1983), Trinkaus et al. (2014,2017) and Pablos et al. (2017).
The Tarsometatarsal Skeletons
The tarsometatarsal (TMT) skeletons of all of these hominins indicate pedal struc-
tures that are similar to those of habitually unshod recent humans. They have
compact and mediolaterally compressed posterior tarsals, with the calcaneal
7 Perspectives on Pliocene and Pleistocene Pedal Patterns and Protection 123
Fig. 7.1 Dorsal views of articulated pedal skeletons (above) and dorsal or plantar views of rst
metatarsals (below). The articulated pedal skeletons include an australopith composite (OH-8,
A.L. 333-115, StW 617), Dinaledi Foot 1, Kiik-Koba 1 and Sunghir 1. The more bubous heads
of the Sterkfontein, Swartkrans and Dmanisi MT-1 s are circled
tuberosity largely in line with the talar trochlea. They have low talar neck angles. The
australopiths have large naviculocuboid facets, possibly reecting modestly greater
midtarsal mobility, but they are reduced to absent in Homo tarsals. All of them have
fully adducted hallucal metatarsals despite some variation in angulation when the
skeletal elements are articulated. The adduction is reected in tarsometatarsal 1 artic-
ular orientations and the occasional metatarsal (MT) 1-2 facets; the mediolaterally
curved and distally convex TMT-1 facets of some remains enhanced joint stability
and were not abduction. They had fully formed longitudinal and transverse pedal
arches, indicated by MT torsion (especially for rays 3 and 4) and oblique and
horizontally oriented TMT articulations (especially for rays 35). In combination
with the pedal arches, the metatarsophalangeal (MTP) articulations have
mediolaterally oriented axes of rotation; for the MT-1, this resulted in perpendicular
proximal and distal articular axes of rotation, permitting effective dorsiexion at
heel-off.
In this context, there was a variation in the degree of MT-1 medial divergence,
overall pedal proportions, the relative sizes of articulations and other details of
articular facets. It remains unclear to what extent these variations reect body size
(especially between australopiths and Homo), body proportions (especially
ecogeographically in Middle and Late Pleistocene Homo), musculoskeletal hyper-
trophy and the effects of the presence/absence of habitual footwear use. None of
these variations would have affected the basic kinesiology of the foot during a
striding gait, beyond the considerable individual variation evident among recent
humans.
124 E. Trinkaus et al.
The Hallux
In the context of adducted halluces, there are contrasts between earlier and later
hominins in two aspects, the shape of the MTP articulation and the presence/degree
of distal phalangeal lateral deviation (hallux valgus). Both have functional
implications.
The MT-1 heads of most Middle and Late Pleistocene MT-1s are indistinguish-
able from those of recent humans in being relatively large and modestly distally
convex, with varying degrees of distal angulation caused by different degrees of
dorsal extension of the intersesamoid crest (Fig. 7.1). The articulation is evidently
adapted for transmitting elevated axial joint reaction forces with only modest degrees
of abduction-adduction and dorsiexion-plantarexion. The Dinaledi MT-1s are
similar to the other later Pleistocene ones in shape, but they have relatively smaller
articulations. In contrast, the australopith and initial Pleistocene Homo MT-1s
exhibit mediolaterally and dorsoplantarly bulbous heads (Fig. 7.1). Although fully
compatible with predominantly axial joint reaction forces, their marked convexities
imply increased mobility of the MTP-1 joint and/or increased joint stability relative
to mediolateral forces on the distal hallux.
As a result of normal toeing-out during walking, most recent humans exhibit a
lateral deviation of the distal hallucal phalanx (DP-1), or hallux valgus. All of the
known Late Pleistocene and the Middle Pleistocene Atapuerca-SH DP-1s exhibit a
similar lateral deviation (Fig. 7.2). In contrast, the few known DP-1s from
australopiths, Early Pleistocene Homo and the Middle Pleistocene Dinaledi sample
exhibit minimal lateral deviation of the DP-1. This is particularly evident in the
complete OH-10 phalanx. Given that DP-1 lateral deviation is produced by differ-
ential medial versus lateral metaphyseal growth during development, from habitual
forces on the hallux, the absence of this angulation in the earlier DP-1s implies little
to no toeing-out among these hominins. Yet, at least OH-10 and the Dinaledi DP-1
exhibit axial torsion, which implies a humanlike toe-off.
The Lateral Metatarsophalangeal Articulations
During heel-off and the propulsive phase of a human stance, the ball of the foot and
the toes are on the substrate, the pedal arch is raised and consequently the MTP
articulations are substantially dorsiexed. This distinctively human pedal posture
has resulted, most prominently in recent humans, in a dorsal extension (or doming)
of the metatarsal heads. The dorsal doming of the lateral metatarsal heads is present
in all of the Middle and Late Pleistocene humans (including the Dinaledi remains).
Additionally, the few Early Pleistocene Homo specimens appear to follow the recent
human pattern. However, since this feature is variably present in the australopith
MTs, it is unclear to what extent the australopith MTP articulations were habitually
hyperdorsiexed, as in a fully human heel-off.
7 Perspectives on Pliocene and Pleistocene Pedal Patterns and Protection 125
Fig. 7.2 Dorsal views of distal hallucal phalanges (below) and the distributions of DP-1 lateral
deviation angles (A)(above)
126 E. Trinkaus et al.
MTP dorsiexion at heel-off also produces a proximodorsal orientation
(or canting) in recent human lateral proximal phalangeal metatarsal facets, especially
of digits 24 (PP-2 to PP-4), such that the articular surfaces are oriented largely
perpendicular to the resultant joint reaction forces. Given the more proximal position
of the MT-5 head and the relative shortness of the fth proximal phalanx (PP-5) in
recent humans, this feature is less pronounced in the fth MTP articulations.
Because of the difculty in assigning isolated PP-2s to PP-4s to digit and
australopith PP-5s to digit, it is necessary to pool these phalanges for comparaisons.
All of the Middle and Late Pleistocene PPs follow the recent human pattern, with the
lower articular angles deriving from PP-5s (Fig. 7.3). The same applies to the Middle
Pleistocene Atapuerca-SH sample. The one Early Pleistocene phalanx, likely of
Homo (SKX-16699), is among the more recent humans. The australopiths, although
they have mostly dorsally oriented facets (in contrast to the plantar orientations of
ape facets), exhibit angles that are substantially below those of Homo PPs,
overlapping only the low values of some Late Pleistocene PP-5s.
Lateral Proximal Phalanx Lengths and Shafts
The lateral proximal pedal phalanges have generally uniform articular lengths
through the Late Pleistocene and including the Atapuerca-SH Middle Pleistocene
sample, with sample median lengths of 2325 mm. The SKX-16699 Early Pleisto-
cene Homo phalanx and the Dinaledi ones are shorter, averaging 1820 mm in
length, given smaller body sizes. However, the australopith ones, although variable,
are considerably longer, with a median length of 2728 mm, despite their small
bodies. When compared to estimated body mass from femoral head diameters
(Fig. 7.4) (by individual for the Late Pleistocene and A.L. 288-1 and by each phalanx
to every body mass estimate for the other samples), the Pleistocene Homo samples
are very similar. The australopith PPs, with their generally smaller body sizes, are
substantially relatively longer (even ignoring the few, probably inappropriate, high
ratios). A few of the australopith ratios overlap the Homo ones, and two of the
Dinaledi ones are also relatively high. But there is nonetheless a substantial dichot-
omy between the australopith and Homo relative phalangeal lengths.
The longer australopith proximal phalanges are associated with a suite of related
diaphyseal features that contrast with those of later Homo (Figs. 7.3 and 7.4). The
Homo phalanges exhibit largely dorsally straight diaphyses, ovoid-shaped midshafts
and small exor sheath ridges located on the medial and lateral midshafts (although
the Dinaledi and SKX-16699 PPs have slight dorsal convexities). The australopith
phalanges are distinctly curved on their dorsal margins. They have prominent exor
sheath crests that are on the medioplantar and lateroplantar diaphysis along the distal
halves of the shafts. The sizes and positions of these exor sheaths are a product of
having more curved diaphysis and hence greater plantarly directed forces on the
sheaths, likely arising from the stronger contraction of the long pedal digital exor
7 Perspectives on Pliocene and Pleistocene Pedal Patterns and Protection 127
Fig. 7.3 Lateral views of
lateral proximal pedal
phalanges (below) and
distributions of proximal
articular angles. The angle is
relative to the mid-articular
axis and is generally lower
than the canting angle
(which inappropriately uses
the plantar surface as the
plane of reference). The
Middle Pleistocene
Atapuerca-SH sample has
angles similar to the Late
Pleistocene samples
(Fig. 7.4)
muscles. These specic features cause the midshafts to appear semicircular in cross-
sectional shape. The phalanges are also mediolaterally expanded more distally,
giving the diaphysis a proximally waisted appearance in dorsal view. However
assessed, the australopith lateral proximal pedal phalanges imply some degree of
prehension, albeit markedly less than the much longer ones of the great apes.
128 E. Trinkaus et al.
Fig. 7.4 Dorsal views of lateral proximal pedal phalanges (below) and articular length/estimated
body mass (as a percentage) (above). For the Middle and Upper Paleolithic samples and A.L. 288-1,
the comparisons are within individuals. For the remainder of the earlier samples, given the absence
of associated phalanx lengths and body mass estimates, each phalangeal length is divided by each
femoral head-based body mass estimate available for the appropriate sample. For the two Middle
Pleistocene samples (Dinaledi and Atapuerca-SH), the comparisons are within site. For the StW,
DNH and A.L. phalanges, the body mass estimates are from the femora attributed to Au. africanus,
P. robustus and Au. afarensis respectively. For SKX-16699, the body mass estimates are for those
attributed to early Homo (sensu stricto). For these reasons, and the pooling of lengths from rays 2 to
5, the box plots for the earlier samples (especially the australopiths) exhibit considerably greater
variation than is indicated by the phalanges themselves
7 Perspectives on Pliocene and Pleistocene Pedal Patterns and Protection 129
Proximal Pedal Phalanx Diaphyseal Hypertrophy
It is also possible to assess the relative degrees of hypertrophy, or robustness, of the
lateral pedal phalanges, again pooling those from digits 2 to 5 and comparing those
phalanges without an individually associated body mass to the range of body masses
available for the appropriate sample (Fig. 7.5). To maximize sample sizes, midshaft
polar moments of area are estimated from the diaphyseal diameters using ellipse
formulae, and they are scaled using articular length times estimated body mass. The
resultant values (Fig. 7.5) provide a large range for the australopiths, low values for
the Early Pleistocene SKX-16699, higher values for the Dinaledi sample and
intermediate and similar ranges for the Middle Pleistocene and Middle Paleolithic
samples. Given the extensive overlap of these samples and the necessity to associate
almost all of the pre-Middle Paleolithic ones by sample rather than by individual,
there is probably little signicance in the variations across these fossil samples.
Fig. 7.5 Comparisons of midshaft relative strength across the paleontological samples and three
recent Native American samples. Polar moments of area (J/Ip) are computed using standard ellipse
formulae from the diaphyseal diameters, modelling the diaphysis as solid, and each is relative to the
estimated body mass times phalangeal articular length (see Trinkaus and Patel 2016). As in Fig. 7.4,
Late Pleistocene and recent phalanges, plus that of A.L. 288-1, are scaled by individual. The others
are scaled to each of the body mass estimates for the appropriate sample. The Upper Paleolithic
sample is subdivided regionally, and the recent human samples represent prehistoric Native
Americans who were habitually unshod (Pecos Pueblo), shod (Point Hope) and variably shod
(Libben)
130 E. Trinkaus et al.
In the assessment of the Upper Paleolithic phalanges, however, there is substan-
tial interregional variation. The western Eurasian sample has relatively gracile
phalanges, whereas the North African and especially the Southeast Asian one have
more robust phalanges. In contrast, there is little difference across these samples in
overall lower limb robustness. If these three samples are compared to
ecogeographically separate Native American prehistoric samples, however, a pattern
emerges. Across the Native American samples, the habitually unshod Pecos Pueblo
(New Mexico) sample has robust phalanges, similar to Middle Pleistocene and
Middle Paleolithic ones. The habitually shod Inuit Point Hope (Alaska) sample
has relatively gracile ones, similar to the western Eurasian Upper Paleolithic sample.
And the geographically intermediate Libben (northern Ohio) sample is modestly less
gracile. Given that these three samples of Native Americans were similarly robust in
their lower limbs, the variation in lateral pedal phalangeal hypertrophy reects the
degrees to which their lateral toes were protected by differences in the habitual use of
footwear. If the framework from these Native Americans is then applied to the Upper
Paleolithic samples, the inference is that the western Eurasian sample was habitually
shod, the Southeast Asian one was often barefoot and the North African sample was
intermediate but closer to the Southeast Asian one in the use of protective foot wear.
Implications for Pliocene and Pleistocene Footprints
The tarsometatarsal congurations of all of these Pliocene and Pleistocene pedal
remains are therefore basically similar to those of recent humans, despite minor
variations in size, proportions, articular details and musculoligamentous hypertro-
phy. They therefore imply that the primary forms of the footprints attributed to
australopiths or members of the genus Homo should be similar. Given the high
degree of variation in unshod footprint form within and across individuals among
recent humans, due to normal ranges of pedal size and proportions, the variation in
digital separation, degrees of toeing out during walking and variable pedal arch
height, overlain by idiosyncratic variation in walking patterns, terrain and (of course)
substrate characteristics, all of these hominins should have made footprints which
were generally similar.
The areas of functional contrast in the pedal remains involve the digits. The
australopiths and (to varying degrees) initial Homo digital remains indicate greater
hallucal mobility and/or lateral forces on the hallux, a lack of hallux valgus (hence
little toeing out), longer lateral phalanges and lateral phalanges which were less
dorsiexed in the later stages of the stance phase. The expectation would therefore
be that australopith footprints, relative to those of later Homo, would exhibit normal
human heel, arch and ball imprints, but that they would contrast in having less toeing
out of the print (or more anteroposterior orientation of the footprint) and especially
distally extended and deeper impressions from the lateral toes.
7 Perspectives on Pliocene and Pleistocene Pedal Patterns and Protection 131
The one axis of variation among later Pleistocene humans is the reduction in
lateral phalangeal robustness among the western Eurasian Upper Paleolithic humans
and to a lesser extent among the North African ones. Especially compared to the
Middle Pleistocene and the Middle Paleolithic samples, the implication is that there
was a marked increase in the use of protective foot wear among these Upper
Paleolithic human populations. Paleolithic footwear is not known, although at least
one sample (the early Upper Paleolithic Sunghir one from northern Russia) exhibits
both body decoration implying leggings/boots and extremely gracile lateral phalan-
ges, indicating their habitual use of protective boots. Interestingly, almost all of the
footprints known from Upper Paleolithic Eurasia are of unshod people, whether of
children or adults. Were these people more often barefoot than their pedal phalanges
and their cold temperate to glacial environments imply? Were they removing
footwear to walk more securely in the karstic systems in which the footprints are
primarily found? Or is there a bias in our footprint sample, such that the distinctively
human barefoot ones are readily recognized, but the more amorphous ones that
would be created by soft boots remain unrecorded?
Acknowledgements S. Potze and B. Zipfel provided support and access to the SK, SKX and StW
collections at the Ditsong National Museum of Natural History and at the Evolutionary Studies
Institute, University of the Witwatersrand, South Africa, respectively. We thank Dmanisi research
team for access to Dmanisi collection at the Dmanisi Museum-Reserve, Georgian National
Museum. We thank the Evolutionary Studies Institute, the University of the Witwatersrands
Microfocus X-ray Computed Tomography Facility for facilitating access to the high-resolution
micro-CT data used for gure presentations. The Atapuerca-Sima de los Huesos images and
comparative data have been made available by A. Pablos, and the Dinaledi images were provided
by W. Harcourt-Smith. Access to Late Pleistocene human remains has been provided by a large
number of individuals and curating institutions, and L.L. Shackelford and G. Apfeld assisted with
the collection of the recent human data. To all of these individuals and institutions, we are
immensely grateful.
References
DeSilva, J., McNutt, E., Benoit, J., & Zipfel, B. (2019). One small step: A review of Plio-
Pleistocene hominin foot evolution. American Journal of Physical Anthropology, 168(S67),
63140. https://doi.org/10.1002/ajpa.23750.
Fernández, P. J., Mongle, C. S., Leakey, L., Proctor, D. J., Orr, C. M., Patel, B. A., et al. (2018).
Evolution and function of the hominin forefoot. Proceedings of the National Academy of
Sciences USA, 115, 87468751. https://doi.org/10.1073/pnas.1800818115.
Harcourt-Smith, W. E. H., Throckmorton, Z., Congdon, K. A., Zipfel, B., Deane, A. S., Drapeau,
M. S. M., et al. (2015). The foot of Homo naledi.Nature Communications, 6, 8432. https://doi.
org/10.1038/ncomms9432.
Latimer, B. L., Lovejoy, C. O., Johanson, D. C., & Coppens, Y. (1982). Hominid tarsal, metatarsal,
and phalangeal bones recovered from the Hadar Formation: 19741977 collections. American
Journal of Physical Anthropology, 57, 701719.
Lordkipanidze, D., Jashashvili, T., Vekua, A., Ponce de León, M. S., Zollikofer, C. P. E.,
Rightmire, G. P., et al. (2007). Postcranial evidence from early Homo from Dmanisi, Georgia.
Nature, 449, 305310.
132 E. Trinkaus et al.
McNutt, E. J., Zipfel, B., & DeSilva, J. M. (2018). The evolution of the human foot. Evolutionary
Anthropology, 27, 197217. https://doi.org/10.1002/evan.21713.
Pablos, A., Pantoja, A., Martínez, I., Lorenzo, C., & Arsuaga, J. L. (2017). Metric and morpho-
logical analysis of the foot in the Middle Pleistocene sample of Sima de los Huesos (Sierra de
Atapuerca, Burgos, Spain). Quaternary International, 433A, 103113. https://doi.org/10.1016/
j.quaint.2015.08.044.
Susman, R. L. (1983). Evolution of the human foot: Evidence from Plio-Pleistocene hominids. Foot
and Ankle, 3, 365376. https://doi.org/10.1177/107110078300300605.
Trinkaus, E. (1983). Functional aspects of Neandertal pedal remains. Foot and Ankle, 3, 377390.
Trinkaus, E. (2005). Anatomical evidence for the antiquity of human footwear use. Journal of
Archaeological Science, 32, 15151526. https://doi.org/10.1016/j.jas.2005.04.006.
Trinkaus, E., & Patel, B. A. (2016). An Early Pleistocene human pedal phalanx from Swartkrans,
SKX 16699, and the antiquity of the human lateral forefoot. Comptes Rendus Palevol, 15,
978987. https://doi.org/10.1016/j.crpv.2016.07.003.
Trinkaus, E., Buzhilova, A. P., Mednikova, M. B., & Dobrovolskaya, M. V. (2014). The people of
Sunghir: Burials, bodies and behavior in the earlier Upper Paleolithic. New York: Oxford
University Press.
Trinkaus, E., Wojtal, P., Wilczyński, J., Sázelová, S., & Svoboda, J. (2017). Palmar, patellar and
pedal human remains from Pavlov. PaleoAnthropology, 2017,73101. https://doi.org/10.4207/
pa.2017.art106.
Ward, C. V. (2013). Postural and locomotor adaptations of Australopithecus species. In K. E. Reed,
J. G. Fleagle, & R. E. Leakey (Eds.), The paleobiology of Australopithecus (pp. 235245).
New York: Springer. https://doi.org/10.1007/978-94-007-5919-0_16.
Zipfel, B., DeSilva, J. M., Kidd, R. S., Carlson, K. J., Churchill, S. E., & Berger, L. R. (2011). The
foot and ankle of Australopithecus sediba.Science, 333, 14171420. https://doi.org/10.1126/
science.1202703.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Chapter 8
Frozen in the Ashes
The 3.66-Million-Year-Old Hominin Footprints
from Laetoli, Tanzania
Marco Cherin, Angelo Barili, Giovanni Boschian, Elgidius B. Ichumbaki,
Dawid A. Iurino, Fidelis T. Masao, Soa Menconero, Jacopo Moggi Cecchi,
Susanna Sarmati, Nicola Santopuoli, and Giorgio Manzi
Abstract Fossil footprints are very useful palaeontological tools. Their features can
help to identify their makers and also to infer biological as well as behavioural
information. Nearly all the hominin tracks discovered so far are attributed to species
of the genus Homo. The only exception is represented by the trackways found in the
late 1970s at Laetoli, which are thought to have been made by three Australopithecus
afarensis individuals about 3.66 million years ago. We have unearthed and described
the footprints of two more individuals at Laetoli, who were moving on the same
surface, in the same direction, and probably in the same timespan as the three found
in the 1970s, apparently all belonging to a single herd of bipedal hominins walking
from south to north. The estimated stature of one of the new individuals (about
1.65 m) exceeds those previously published for Au. afarensis. This evidence sup-
ports the existence of marked morphological variation within the species. Consid-
ering the bipedal footprints found at Laetoli as a whole, we can hypothesize that the
tallest individual may have been the dominant male, the others smaller females and
juveniles. Thus, considerable differences may have existed between sexes in these
human ancestors, similar to modern gorillas.
M. Cherin (*) · A. Barili
University of Perugia, Perugia, Italy
e-mail: marco.cherin@unipg.it
G. Boschian
University of Pisa, Pisa, Italy
E. B. Ichumbaki · F. T. Masao
University of Dar es Salaam, Dar es Salaam, Tanzania
D. A. Iurino · S. Menconero · S. Sarmati · N. Santopuoli · G. Manzi
Sapienza University of Rome, Rome, Italy
J. Moggi Cecchi
University of Florence, Florence, Italy
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_8
133
Keywords Australopithecus afarensis · Bipedalism · Body size · Conservation
134 M. Cherin et al.
Forty Years of Research at Laetoli
Fossil skeletal elements can offer plenty of data on different aspects of human
evolution and palaeobiology. However, the amount of information we can get
about our ancestors can increase signicantly through the study of fossil footprints.
In fact, these ephemeral traces of life in the geological past can provide key
palaeobiological insights on anatomy, locomotion biomechanics, body size, social
behaviours, palaeoenvironments, and even reproductive strategies of extinct
hominins (Falkingham et al. 2018). Unfortunately, due to the extremely peculiar
taphonomic conditions that can lead to their preservation, fossil footprints are very
rare. The hominin ichnological record is particularly poor (Bennett and Morse 2014),
especially when compared to that of other vertebrate groups like dinosaurs. Nearly
all the hominin tracks discovered so far are attributed to species of the genus Homo,
with the outstanding exception of the record from Laetoli (Tanzania), dated to 3.66
million years ago (Ma).
Laetoli is one of the most important palaeoanthropological sites in the world. It is
located in northern Tanzania (Fig. 8.1) at the southern margin of the Serengeti plains,
within the Ngorongoro Conservation Area (NCA), in which several other world-
famous palaeoanthropological localities like Olduvai Gorge, Lake Ndutu, and
Nasera Rock, are found.
The Laetoli stratigraphic sequence is composed of Plio-Pleistocene volcano-
sedimentary deposits divided into ve main lithological units, the Laetolil Beds,
Ndolanya Beds, Olgol Lavas/Naibadad Beds, Olpiro Beds, and Ngaloba Beds, from
bottom to top (Hay 1987). At the base, the Laetolil Beds make up most of the
sedimentary sequence, with a thickness of more than 120 m (Ditcheld and Harrison
2011). They probably formed from tephra erupted from the extinct Sadiman vol-
cano, located about 20 km to the east of Laetoli (Hay 1987; Mollel et al. 2011),
although this hypothesis is questioned by some authors (Zaitsev et al. 2011,2015).
The Laetolil Beds are divided into two units, namely, the Lower and Upper Laetolil
Beds (4.363.85 Ma and 3.853.63 Ma, respectively) (Hay 1987; Deino 2011). The
latter consist of a series of aeolian and fall-out tuffs (Hay 1987) and are well known
for their abundant palaeontological content (Harrison and Kweka 2011).
The palaeoanthropological relevance of the Laetoli area and of the Upper Laetolil
Beds in particular is well known since the mid-1930s (Reck and Kohl-Larsen 1936;
Kohl-Larsen 1943), but the site turned the attention of the academia and general
public in the 1970s, with the discovery of the holotype and other remains of
Australopithecus afarensis (Leakey et al. 1976; Johanson et al. 1978), as well as
of the earliest bipedal hominin tracks in the world (Leakey and Hay 1979; Leakey
and Harris 1987).
Mammal, bird, and insect prints and trails were identied by Mary Leakey and
collaborators in 18 sites (labelled from A to R) out of 33 total palaeontological
8 Frozen in the Ashes 135
Fig. 8.1 Geographical location and site map; (a) location of Tanzania; (b) location of Laetoli in
northern Tanzania; (c) plan view of Laetoli Locality 8 (Sites G and S); (d) plan view of the four test-
pits excavated at Laetoli Site S. Dashed lines indicate uncertain contours. Hominin tracks in orange,
equid tracks in dark green, rhinoceros track in red, giraffe tracks in light brown, and bird tracks in
blue. (Modied from Masao et al. 2016)
localities in the Laetoli area (Leakey 1987; Harrison and Kweka 2011; Musiba et al.
2008). The so-called Footprint Tuff, which corresponds to the lower part of Tuff 7 in
the Upper Laetolil Bedsstratigraphic sequence, hosts at least ten sublevels in which
footprints are found (Hay 1987). Among these, hominin tracks were originally
discovered at Site G (Locality 8). A short trackway of humanlike footprints was
also unearthed at Site A (Locality 6) but was later attributed to a bear (Tuttle 2008).
Site G footprints were referred to three individuals (G1, G2, G3) of different body
sizes: the smaller G1 walked side by side on the left of the larger G2, while the
intermediate-sized G3 superimposed its feet over those of G2 (Leakey 1981). These
trackways are ascribed to Au. afarensis (White and Suwa 1987; Masao et al. 2016),
136 M. Cherin et al.
which is the only hominin taxon found to date in the Upper Laetoli Beds (Harrison
2011).
Immediately after the rst publication (Leakey and Hay 1979), the scientic and
public interest in the Laetoli footprints has spread extraordinarily. Since then, they
have been mentioned in hundreds, if not thousands, of scientic works(Jungers
2016), and a Google search for Laetoli footprints returns more than 66,000 results at
the date of writing this contribution. In the rst years after the discovery, the tracks
were studied in several papers dealing with interactions between trackmakers (Lea-
key and Hay 1979; Leakey 1981), foot anatomy and locomotion (Day and Wickens
1980; White 1980; Charteris et al. 1981,1982; Stern and Susman 1983), and
depositional and palaeoenvironmental setting (Leakey and Hay 1979; Leakey
1981). In more recent years, with the development of new technologies and methods,
many researchers worked on the key topic of locomotion of the Laetoli trackmakers
by means of different approaches. These analyses led to conicting views, with some
authors (e.g. Raichlen et al. 2010; Crompton et al. 2012) inferring that the gait
pattern of the Laetoli hominins was similar to that of modern humans, while others
(e.g. Meldrum 2004; Schmid 2004; Bennett et al. 2009; Hatala et al. 2016) inferring
that it was qualitatively and/or quantitatively different. However, regardless of the
methods used, all the above studies are equally negatively affected by the fact that
they are focused only on a limited number of G1 tracks. Although most of the G1
trackway is well preserved (unlike the overlapping footprints of G2 and G3), it
belongs to the smallest of G individuals, which was very likely a juvenile (see
Laetoli Site S Footprints: Results and Implications). Moreover, the original tracks
are today buried under a protective cover (Feibel et al. 1996), and most of the studies
were carried out on casts.
In light of the above, the recent discovery of new human footprints in Laetoli is of
crucial importance for the knowledge of the anatomy, palaeobiology, and behaviour
of Pliocene hominins.
The Discovery of Laetoli Site S
In 2014, two of us (F.T.M., E.B.I.) were commissioned to carry out a Cultural
Heritage Impact Assessment (CHIA) aimed at evaluating the impact of a proposed
new eld museum in the area of Locality 8, that is, the palaeontological locality in
which M. Leakey and co-authors discovered the rst human tracks in the 1970s.
According to Tanzanias Environmental Management Act (United Republic of
Tanzania 2004), the CHIA is part of the Environmental and Sociological Impact
Assessment (ESIA), which is a mandatory evaluation process expected to address
the impact of a certain development project (e.g. infrastructure construction) on the
environment, landscape, and social context (Ichumbaki and Mjema 2018). In par-
ticular, the CHIA is focused on the possible impacts on cultural heritage, both extinct
(e.g. archaeological and palaeontological record) and extant (e.g. ethno-
anthropological context).
Specic objectives of the CHIA were:
1. Salvaging as much of the threatened heritage as possible through surface collec-
tion and excavation
8 Frozen in the Ashes 137
2. Preliminary analysis of the archaeological and palaeontological material rescued
3. Packing the material and presenting it to Ngorongoro Conservation Area Author-
ity (NCAA) for curation and storage
4. Proposing mitigation measures including immediate conservation of special
features encountered in the eldwork process and proposing an appropriate
monitoring schedule
The CHIA assignment was accomplished through two main eldwork seasons.
During the rst season (June 2130, 2014), the team of archaeologists, cartogra-
phers, conservators, and skilled workers aimed at obtaining an overall picture of the
cultural heritage features in the area impacted by the project. In particular, the team
surveyed a wide area within 500 m radius from the Site G trackways, i.e. the core
area of the proposed museum project. The second season (September 13 to October
22, 2014) focused on the area of maximum impact, i.e. the surroundings of Site
G. Sixty-two 2 ×2 m test-pits (each corresponding to about 2% of the total surface)
were randomly positioned within a grid and carefully excavated down to the
Footprint Tuff and sometimes deeper. If necessary, in case of particularly signicant
nds (see below), some pits were enlarged compared to the standard of 2 2m.
×
About 150 m to the south of Site G, the team unearthed 14 hominin tracks
associated with abundant tracks of other vertebrates. Footprints were found in
three test-pits, respectively labelled L8, M9, and TP2 from north to south. The
original square shape of L8 was modied soon after the discovery of the rst bipedal
tracks in order to follow the trail, thus obtaining a quite irregular shape of this test-pit
(southern side, 2 m; western oblique side, 4 m). M9 was excavated some 14 m to the
SSE of L8 and kept the standard size of 2 ×2 m. Following the putative alignment of
the trackway, a third smaller test-pit, TP2 (1 ×1.2 m), was excavated at some 8 m to
138 M. Cherin et al.
Fig. 8.2 Test-pit L8. Photo (left) and shaded 3D photogrammetric model (right) of the southern
part of the printed surface, with close-ups (bottom) of the four footprints
the SSE of M9. Finally, test-pit M10 (2 ×3 m) was excavated about 15 m to the east
of M9 to assess the occurrence of other interesting tracks (Fig. 8.1).
Once the presence of the new tracks has been ascertained and with the aim of
characterizing the printed surface with a multidisciplinary approach, the Tanzanian-
Italian research group was established, pivoting on a collaboration already started for
years in Olduvai Gorge (e.g. Cherin et al. 2016). The new team reopened the four
test-pits in September 2015. Fourteen hominin tracks in different preservation states
always associated with tracks of other vertebrates were unearthed in test-pits L8, M9,
and TP2 (Masao et al. 2016). All these prints are clearly referable to a single trail,
with an estimated total length of 32 m and trending SSE to NNW. Following the
code used for the other footprint sites in Laetoli (Leakey 1981; Leakey 1987;
Harrison and Kweka ), the new site was identied as Site S, and the new tracks
are attributed to individual S1 (footprint numbers S1-17 in L8, S1-14 in M9, and
S1-12 in TP2) (Fig. 8.2
2011
). An additional track referable to a second individual
(S2) was found in the SW corner of TP2. Conversely, only non-hominin footprints
were recorded in M10 (Masao et al. 2016).
8 Frozen in the Ashes 139
Survey of Laetoli Site S: A Case Study for Photogrammetry
Application in Extreme Environments
Modern developments in computing power, rendering software, and hardware
availability allowed a rapid and widespread diffusion of photogrammetry techniques
in Earth Sciences and other disciplines. The majority of these techniques are based
on Structure from Motion (SfM) algorithms (e.g. Luhmann et al. 2013; Mallison and
Wings 2014; James et al. 2017). Among others, SfM techniques have been used in
recent years to study river systems (e.g. Marteau et al. 2017), landslide dynamics and
volumes (e.g. Stumpf et al. 2015), cliff morphology (e.g. Warrick et al. 2017), active
fault structure and dynamics (e.g. Johnson et al. 2014), geobody architecture of
depositional systems (e.g. Mancini et al. 2019), as well as ichnological contexts with
human tracks (e.g. Rüther et al. 2012; Bennett et al. 2016; Bustos et al. 2018; Helm
et al. 2018; Zimmer et al. 2018; Romano et al. 2019). SfM algorithms allow
obtaining three-dimensional models from a series of overlapping pictures taken
from different camera positions. The obtained high-resolution models can be easily
shared between researchers and can be used for detailed qualitative descriptions and
accurate quantitative analyses at the sub-mm-scale (Mallison and Wings 2014).
However, in order to get affordable data from SfM, eld data must be supported
by accurate in situ topographic measurements.
The photogrammetric survey of the new footprint Site S was carried out in an
extreme environmental context, characterized by unfavourable climatic conditions,
need for light equipment, and little time available. Therefore, we had to set up a
working procedure that, despite these problems, could lead to good results in terms
of accuracy and precision and could also serve as a reference for other scientic
activities in similar contexts (Menconero et al. 2019).
The Laetoli area is located over a wide plateau at about 1700 m above sea level, to
the west of the volcanic complex of Sadiman (2870 m), Lemagrut (3135 m), and
Oldeani (3200 m), and north to the Lake Eyasi basin. The plateau is characterized by
a tabular or slightly corrugated morphology. In some areas, the landscape is more
articulated due to the presence of valleys, gorges, and gullies originated by the action
of wind and small streams, whose erosional energy is very intense during the dry
season (MayOctober) and rainy season (NovemberMay), respectively. The cur-
rent vegetation cover is primarily determined by topography, soil composition, and
climate (Anderson 2008)but is also inuenced by natural and anthropic res, as well
as by the grazing activity of the extremely abundant wild herbivore mammals and
domestic livestock (cattle, sheep, and goats) bred by local tribes with nomadic/semi-
nomadic pastoral economy (Holdo et al. 2009). The vegetation mainly includes
thorny thickets and dry bushland, consisting of shrubby and arboreal deciduous
species of the genera Vachellia,Senegalia, and Commiphora, associated with
several forms of grasses (e.g. Sporoboro,Digitaria,Themeda,Aristida,Brachiaria,
Cenchrus) (Herlocker and Dirschl 1972;Andrews and Bamford 2008). The presence
of numerous and densely distributed thorny plants can cause numerous problems
during research activities. The wild animal community of the Laetoli area is still
abundant and diverse, thanks to the low human demographic density, the presence of
impenetrable thorny xerophilous shrublands, and the protection measures by the
NCAA. Among reptiles, several snakes such as the black mamba (Dendroaspis
polylepis), green mamba (D. angusticeps), Egyptian cobra (Naja haje), spitting
cobra (N. nigricollis), and puff adder (Bitis arietans)can be potentially very
dangerous to humans. The same goes for some large-sized mammals, like the
African bush elephant (Loxodonta africana), spotted hyena (Crocuta crocuta),
leopard (Panthera pardus), and African lion (Panthera leo). With regard to scientic
activities, the low demographic density makes very hard to get consumer goods like
food, water, and materials, which have to be bought in larger villages, such as
140 M. Cherin et al.
Karatu, about 4-hour drive from Laetoli. As for the hygienic and sanitary conditions,
thanks to the high average altitude and predominantly dry climate of the plateau, the
whole Laetoli area is less affected by tropical pathologies that are common in the
nearby low-altitude areas. However, especially in wet areas close to perennial small
rivers, there are small populations of hematophagous dipterans like the mosquitos
Anophele (vector of Plasmodium, responsible for malaria) and Aedes (carrier of
various viruses responsible for serious diseases), as well as some horseies
(Tabanidae) and blackies (Simuliidae), which can cause painful bites and severe
skin irritations.
The above environmental conditions (climate, vegetation, fauna) are added to
logistical complications (short time available, problems related to natural lighting,
lack of electricity, long car trips along rough trails) in making extremely difcult the
eldwork in Laetoli (and similar contexts). Under these conditions, clear goals for
the eldwork are necessary. In our case, the work at Site S was aimed at obtaining
3D models of the new tracks for documentation and morphometric analysis. We
chose the SfM photogrammetry technique, thanks to its technical advantages (rela-
tively short time of data acquisition and processing, light and handy equipment,
reduced costs) and excellent results in terms of resolution (Westoby et al. 2012).
Each test-pit (L8, M9, TP2, and M10) was entirely surveyed at lower resolution
(Fig. 8.3), and then detailed 3D models of some inner portions (single tracks or
groups of close prints) were acquired (Fig. 8.4). Targets of the control point system
were immediately positioned after excavation. We placed four perimeter targets at
the corner of each test-pit and four inner targets around each subarea surveyed in
detail (14 in L8, 10 in M9, 14 in TP2, and 14 in M10). For the measurement of the
control points, we used a measuring tape and a water level, which are lighter and
easier to handle than a total station theodolite. We selected these low-tech tools also
considering (1) measuring only four points for each test-pit to scale the general 3D
models and (2) aligning the detailed 3D models of the single footprints to the general
models using the coordinates of the inner targets. For the perimeter target measure-
ments, we placed two rods equipped with a spherical level on successive pairs of
targets, and we marked points at the same height on the rods for each pair by using
the water level device. The vertical distance between these points and the targets, as
well as their mutual distance, were recorded. Repeating this process for all pairs of
8 Frozen in the Ashes 141
targets, the relative plan position and the height of the control points were deter-
mined respectively by trilateration and levelling. A preliminary accuracy check was
carried out by means of trilateration graphic rules in plan and with the method of
successive levelling for heights. By assigning a z-coordinate to the rst control point,
all subsequent coordinates were derived from addition and subtraction of heights
between two successive points. The check was performed by computing the sum of
all height differences and by verifying that the obtained value was close to zero.
Finally, the error obtained in each test-pit was distributed to every z-coordinate of the
points, in order to minimize it.
Photographic acquisition was performed with a DSLR camera, sometimes xed
on a 3-m-long telescopic rod for photographic shots from the top downwards. With
regard to scene lighting, since we had no possibility to control light intensity and
direction, we tried to reduce shadows by shooting especially during the central hours
of the day (i.e. with subvertical sun rays). However, this was not always possible due
to the excavation schedule and little time available. Therefore, we had to address the
problem of high-contrast shadows in post-processing.
The texture resolution control of 3D models, namely, the Ground Sampling
Distance (GSD), can be performed a priori using geometric formulas. The calcula-
tion is based on the principle of similar triangles, which are found in the geometry of
the shooting. The variables are the size of the sensor (Sw) and the focal length of the
camera (Fl), the size in pixel of the images (Iw), and the distance (H). The triangle
with the base Sw and height Fl is similar to the triangle which has the base Gw
(width of the image on the ground) and height H; consequently, the two triangles
have proportional respective sides (Sw: Gw ¼Fl: H). The GSD is the ratio between
Gw and Iw multiplied by 100 (GSD ¼Gw/Iw ×100). Connecting the proportion
with the formula of GSD, the nal formula GSD ¼(Sw ×H×100): (Fl ×Iw) is
obtained. Among the variables, the one that can be easily managed is the distance H,
142 M. Cherin et al.
Fig. 8.3 Test-pit L8; (a) eidotype; (b) shaded model; (c) textured model; (d) textured and shaded
model; (e) drawing; (f) density of the point cloud by determining the number of nearest neighbours
in a sphere with 0.5 cm radius
8 Frozen in the Ashes 143
Fig. 8.4 Three footprints from test-pit L8; (a) textured model; (b) textured and shaded model; (c)
shaded model; (e) density of the point cloud by determining the number of nearest neighbours in a
sphere with 0.5 cm radius
since all the others depend on the photographic equipment available (Menconero
et al. 2019).
It is impossible to know a priori the density of the point cloud coming from a
photogrammetric process. As for the Laetoli footprints, the goal was to obtain a
texture resolution less than 0.1 cm/px. This was achieved by choosing suitable
shooting distance both for the whole test-pits and individual footprints. More than
2000 photos were taken in three working days, for a total of about 50 GB. Especially
when working in remote areas, it is important not to economize on shots and possibly
select them a posteriori.
Data processing started with checking topographic measurements in plan and
height, which is preliminary to the denition of the control point coordinates. The
trilateration method was used to obtain x, y coordinates of the control points in plan.
For each test-pit, six measurements were taken at the same height: the length of the
four sides of the perimeter and the length of the two diagonals. Redundant measure-
ments were used to compute the errors. In addition to a preliminary graphical control
by CAD software, we used an automatic calculation software to adjust a new set of x,
y coordinates and heights of the control points by least squares technique. The
residues of adjustments never exceeded 10 mm, which is fully acceptable consider-
ing the size of the test-pits. We used the adjusted x, y, z coordinates of the control
points to scale and locate in the 3D space the SfM models. A check on point cloud
density was also carried out by a software for 3D point cloud and mesh processing
and analysis. The average density found in the Laetoli point clouds is around
20 points/cm
3
for the test-pits and 1500 points/cm
3
for the detailed footprints (Masao
et al. 2016; Menconero et al. 2019).
The 3D models obtained by SfM were also used in the morphometric analysis of
the hominin tracks. We used a contouring and modelling software that transforms x,
y, z data into maps. The x, y, z-format les were imported into the software and
transformed into grid les. The software uses randomly spaced x, y, z data to create
regularly spaced grids composed of nodes with x, y, z coordinates. The Triangula-
tion with Linear Interpolation gridding method was applied, because it works better
with data that are evenly distributed over the grid area. This method creates network
of triangles with no edge intersection starting from data points and computes new
144 M. Cherin et al.
values along the edges. The grid spacing was set at 1 mm. Standard morphometric
measurements (footprint length, footprint max width, footprint heel width, angle of
gait, step length, and stride length) were taken from contour maps and compared
with those taken manually both on the original tracks during the eldwork and on 1:1
scale sketches of the test-pits, hand-drawn on transparent plastic sheets (Masao et al.
2016
Laetoli Site S Footprints: Results and Implications
).
The detailed analysis of the new bipedal footprints at Site S started trying to frame
this outstanding nding into the stratigraphic context of the Upper Laetolil Beds. A
detailed sequence analysis of the excavation proles at Site S and extended geolog-
ical observations in the whole Laetoli area were performed. In particular, we tried to
reconstruct the stratigraphic relationships between the footprint-bearing units of Site
S and Site G, using both eld observation and literature descriptions of the sequence
outcropping in the original site.
The Laetoli Footprint Tuff is part of Tuff 7 together with the overlying Augite
Biotite Tuff and can be divided into a lower and an upper unit. These can be
respectively subdivided into 14 and 4 sublevels. Tracks are found on eight sublevels
within the lower unit and two within the upper one (Leakey and Hay 1979; Hay and
Leakey 1982; Hay 1987). In particular, hominin tracks at Site G are located on the
top of horizon B, namely, on sublevel 14 of the Footprint Tuff lower unit (Hay and
Leakey 1982; White and Suwa 1987). Though with some local differences presum-
ably due to lateral variability, we found that the Site S sequence corresponded quite
well with the original description of the Footprint Tuff stratigraphy provided by Hay
(1987). In particular, we observed that the Site S tracks were printed on the top of the
lower subunit of the Footprint Tuff, corresponding to the aforementioned horizon B
(Masao et al. 2016). Consequently, our data indicate with reasonable condence that
the footprints of S1 and S2 lie on the same stratigraphic position as those at Site
G. Considering that (1) Tuff 7 includes a sequence of several sublevels originated by
distinct volcanic eruptions close in time, and that its overall deposition time is
estimated in weeks (Hay and Leakey 1982; Hay 1987), (2) trackways from Site G
and Site S show almost the same orientation, and (3) all trackmakers were moving
approximately at the same moderate speed (see below), we hypothesized that the
tracks from the two sites were left by a homogeneous group of hominins walking
together on the same palaeosurface (Masao et al. 2016).
The overall morphology of the S1 tracks ts those from Site G and is particularly
similar to the prints of G2, namely, the larger individual (Robbins 1987): the heel is
8 Frozen in the Ashes 145
oval shaped and is pressed deeply into the ground; the medial side of the arch is
higher than the lateral one; the ball region is oriented at an angle of about 75with
respect to the longitudinal axis of the foot and is delimited anteriorly by a transversal
ridge, formed when the toes gripped the wet volcanic ash and pushed it posteriorly;
the adducted hallux extends more anteriorly than the other toes in all visible
footprints; unfortunately, no clear distinction among the other toes is visible. The
only preserved track of S2 is abnormally widened in the anterior part, probably due
to a lateral slipping of the foot before the toe-off and/or to taphonomic factors related
to the fragmentation of the Footprint Tuff.
Stride length was used to estimate the walking speed of the Laetoli trackmakers.
Mean values of about 0.440.9 m/s were obtained, depending on the computing
method (Alexander 1976; Dingwall et al. 2013). The average length of the S1 tracks
is 261 mm (range 245274 mm). Lower average values were measured for the three
individuals at Site G: 180 mm for G1, 225 mm for G2, and 209 mm for G3 (Leakey
1981; Tuttle 1987), although a study of some G footprint casts based on digital
methods (Bennett et al. 2016) suggested higher values for G1 (193 mm) and G3
(228 mm). Stature was computed rst with Tuttles(1987) approach, which is based
on the ratio between foot length and stature in modern humans (foot length in Homo
sapiens is generally about 1416% of stature). We also estimated stature using the
two methods published by Dingwall et al. (2013). The rst is based on regressions of
stature by footprint length in modern Daasanach people from Lake Turkana (Kenya);
the second which we considered more reliable because it is not inuenced by
modern human data is based on the foot/stature ratio known for Au. afarensis.
Similarly, we estimated the body mass of the trackmakers by means of the regression
equation that relates footprint area to body mass in H. sapiens, as well as of the
equation based on the ratio between foot length and body mass in Au. afarensis
(Dingwall et al. 2013). All the above data were also measured/calculated for G
individuals, using a 3D model of a rst-generation cast of the southern portion of the
Site G trackways.
Our results showed that no matter which method is employed to estimate stature
and body mass, S1 and S2 were taller and had a larger body mass than the G
individuals (S1, 161168 cm/41.348.1 kg; S2, 142149 cm/36.542.4 kg; G1,
111116 cm/28.533.1 kg; G2, 139145 cm/35.641.4 kg; G3, 129135 cm/
33.138.5 kg) (Masao et al. 2016). These results extended the dimensional range
of the Laetoli trackmakers and identied S1 as a large-sized individual, probably a
male. The stature of about 165 cm for S1 is remarkable and exceeds those estimated
to date for any australopithecine. The stature of S1 falls within the maximum range
of modern Homo sapiens and also ts the available Homo erectus sensu lato
estimates based on both skeletal remains and footprints. The body mass range
Fig. 8.5 Minimum and maximum estimated statures of selected fossil hominins by species and
locality over time for the interval 41 million years
estimated for S1 falls within the range of male Au. afarensis (40.261.0 kg)
(Grabowski et al. 2015).
146 M. Cherin et al.
Our results provided independent evidence for the occurrence of large-sized
individuals among hominins as ancient as 3.66 Ma and supported a nonlinear
evolutionary trend in hominin body size (Jungers et al. 2016). Moreover, ascribing
the S1 tracks to a possible male allowed reconsidering the sex and age of the other
Laetoli individuals. According to our body-mass estimations, G1 and G3 fall within
the range of putative Au. afarensis females, whereas G2 and S2 span across the upper
female and the lower male ranges, which are estimated at 25.538.1 and
40.261.0 kg, respectively (Grabowski et al. 2015). A possible reconstruction is
that the Laetoli individuals are S1, a male; G2 and S2, females; and G1 and G3,
smaller females or juvenile individuals.
Both the new composition of the group and the impressive body size difference
point to a considerable sexual dimorphism in Au. Afarensis (Fig. 8.5), as hypothe-
sized by many scholars on the basis of skeletal remains (e.g. Johanson and White
1979; Kimbel and White 1988; McHenry 1991; Richmond and Jungers 1995;
Lockwood et al. 1996; Plavcan et al. 2005; Harmon 2006; Gordon et al. 2008). In
turn, this view supports social organization and reproductive strategies closer to the
polygynous gorillas (Harcourt and Stewart 2007) than to other moderately dimor-
phic species, like the promiscuous chimpanzees or the extant and, possibly, extinct
humans (Masao et al. 2016).
8 Frozen in the Ashes 147
Laetoli Footprints: Perspectives
The recent discovery of Laetoli Site S footprints, after about 40 years from the
pioneering works by Mary Leakey and colleagues at Site G, has achieved a remark-
able media coverage and has drawn the attention of the scientic community.
Raichlen and Gordon (2017) used proportional toe depth (i.e. a measure of the
difference between toe depth and hindfoot depth in tracks) as a proxy to get
information about the locomotor style of the Laetoli trackmakers. They conrmed
that the footprints from Site S are overall very similar to those from Site G, thus
supporting the hypothesis that bipedal locomotion in Au. afarensis was more similar
to modern-human-like extended-limb pattern than to chimpanzee-like bent-knee-
bent-hip pattern (Raichlen et al. 2010).
DeSilva et al. (2019) included data on the Laetoli footprints from Sites G and S in
their comprehensive review of Plio-Pleistocene hominin foot evolution.
In their very interesting work, Villmoare et al. (2019) inferred data on sexual
dimorphism in H. erectus s.l. through the analysis of fossil footprints from Ileret,
Kenya (about 1.5 Ma). Their results are in perfect agreement with ours in the
recognition of a gorilla-like high level of dimorphism in Au. afarensis from Laetoli.
These data are in contrast with those obtained for the Ileret sample, in which
footprints show a much lower degree of sexual dimorphism, although slightly higher
than that of modern humans. According to the authors, this would suggest that by 1.5
million years ago, at least H. erectus s.l. had transitioned away from polygyny
(Villmoare et al. 2019).
Following a completely different line of research, the original contribution by
Ichumbaki et al. (2019) addressed the topic of local communitys interpretations of
the Laetoli hominin footprints. For the rst time, the authors documented narratives
of Maasai (i.e. local people living in the Laetoli area) dealing with their perceptions
on what the footprints are and to whom they belong. The Maasai people connect
Laetoli footprints to the tale of Lakalanga, a strong hero who helped them to win a
battle against a neighbouring community. According to the story which is consol-
idated into the local community oral tradition Lakalanga was so big that wherever
he walked, he left visible tracks on the ground. Thus, the discovery of the large-sized
footprints at Site S has offered a further conrmation to the Maasai that the hero
warrior Lakalanga really existed (Ichumbaki et al. 2019).
The aforementioned papers represent a synthetic selection of those in which Site
S footprints have been studied/mentioned after their recent description (Masao et al.
2016). However, besides these research contributions, the discovery of Site S calls
on the whole international scientic community to question itself on the challenging
issue of conservation. Our eldwork at Laetoli in 20142015 showed us the rele-
vance and peculiarities of the site. Besides the good preservation of the footprints
and their outstanding scientic signicance, we could also verify the aggressiveness
of the East African environment on the ichnological record. Through qualitative
observations of the conservation status of the Footprint Tuff, we could ascertain how
the disruptive action of weathering, ora, and fauna is threatening the footprints even
before excavation. This is jeopardizing a unique piece of cultural heritage that is still
largely unknown. Similar concerns were highlighted by several authors (Feibel et al.
1996; Getty Conservation Institute 1996; Agnew and Demas 1998) after the assess-
ment of the state of conservation of the Site G footprints, which in the 1990s had
been the subject of a project of consolidation, reburial, and protection coordinated by
the Getty Conservation Institute and Tanzanian Antiquities Division (Musiba et al.
2012).Further analyses are necessary at Site S to address the crucial issue of the
conservation of this invaluable palaeontological heritage. Large portions of the
printed surfaces unearthed in the test-pits are already severely threatened by natural
agents and could quickly disappear even if unexcavated (Masao et al. 2016).
Fractures are bringing to disintegration of part of the Footprint Tuff into small
148 M. Cherin et al.
cubic blocks. Roots are displacing the stratigraphic sequence and are opening
preferential ways for water to penetrate the substrate and for arthropods to dig
burrows in the ground. Therefore, keeping the situation as it is may not be the
right way to preserve the site, because unexcavated footprints may be saved from
weathering if they are excavated and properly treated. At the same time, we are
aware that any excavation without a clear understanding of the physical, chemical,
and biological risks to which the Footprint Tuff is exposed should not be undertaken.
A modern project aimed at the excavation, conservation, and valorization of the
Laetoli tracks must be preceded by a multidisciplinary study of environmental data
such as local microclimate, temperature and humidity variations, rainfall, dominant
wind, geological and petrographic characteristics of the substrate, interstitial water
composition and ow, physical damage due to trespassing of livestock and wild
animals, and local vegetation composition and its change. Future plans must include
aprogramme of continuous monitoring of the footprints, especially with the involve-
ment of the local community (Ichumbaki et al. 2019).
Once established a comprehensive plan of conservation and valorization of the
Laetoli footprints in close collaboration with all the involved Tanzanian
(e.g. Ministry of Natural Resources and Tourism, NCAA) and international
(e.g. UNESCO) institutions, new systematic excavations can be carried out. This
would allow unearthing the entire S1 and S2 trackways and also opening up the
intriguing possibility of discovering tracks of other individuals. This new research
would noticeably improve the available dataset on the foot anatomy, locomotor
pattern, body size variation, social structure, and behaviour of the Laetoli
australopithecines.
References
Agnew, N., & Demas, M. (1998). Preserving the Laetoli footprints. Scientic American, 262,
4455.
Alexander, R. M. (1976). Estimates of speeds of dinosaurs. Nature, 261,129130.
Anderson, T. M. (2008). Plant compositional change over time increase with rainfall in Serengeti
grasslands. Oikos, 117, 675682.
8 Frozen in the Ashes 149
Andrews, P., & Bamford, M. (2008). Past and present vegetation ecology of Laetoli, Tanzania.
Journal of Human Evolution, 54,7898.
Bennett, M. R., & Morse, S. A. (2014). Human footprints: Fossilised locomotion? New York:
Springer.
Bennett, M. R., Harris, J. W., Richmond, B. G., Braun, D. R., Mbua, E., Kiura, P., et al. (2009).
Early Hominin foot morphology based on 1.5-million-year-old footprints from Ileret, Kenya.
Science, 323, 11971201.
Bennett, M. R., Reynolds, S. C., Morse, S. A., & Budka, M. (2016). Laetolis lost tracks: 3D
generated mean shape and missing footprints. Scientic Reports, 6, 21916.
Bustos, D., Jakeway, J., Urban, T. M., Holliday, V. T., Fenerty, B., Raichlen, D. A., et al. (2018).
Footprints preserve terminal Pleistocene hunt? Human-sloth interactions in North America.
Science Advances, 4, eaar7621.
Charteris, J., Wall, J. C., & Nottrodt, J. W. (1981). Functional reconstruction of gait from the
Pliocene hominid footprints at Laetoli, northern Tanzania. Nature, 290, 496498.
Charteris, J., Wall, J. C., & Nottrodt, J. W. (1982). Pliocene hominid gait: New interpretations based
on available footprint data from Laetoli. American Journal of Physical Anthropology, 58,
133144.
Cherin, M., Iurino, D. A., Njau, J. K., & Masao, F. T. (2016). New material of hyaenids (Mammalia,
Carnivora) from Olduvai Gorge, Tanzania (Early Pleistocene). Bollettino della Società
Paleontologica Italiana, 55,19.
Crompton, R. H., Pataky, T. C., Savage, R., DAoût, K., Bennett, M. R., Day, M. H., et al. (2012).
Human-like external function of the foot, and fully upright gait, conrmed in the 3.66 million
year old Laetoli hominin footprints by topographic statistics, experimental footprint-formation
and computer simulation. Journal of the Royal Society Interface, 9, 707719.
Day, M. H., & Wickens, E. H. (1980). Laetoli Pliocene hominid footprints and bipedalism. Nature,
286, 385387.
Deino, A. L. (2011).
40
Ar/
39
Ar dating of Laetoli, Tanzania. In T. Harrison (Ed.), Paleontology and
geology of Laetoli: Human evolution in context (Vol. 1, pp. 7797). New York: Springer.
DeSilva, J., McNutt, E., Benoit, J., & Zipfel, B. (2019). One small step: A review of Plio-
Pleistocene hominin foot evolution. American Journal of Physical Anthropology, 168(S67),
63140.
Dingwall, H. L., Hatala, K. G., Wunderlich, R. E., & Richmond, B. G. (2013). Hominin stature,
body mass, and walking speed estimates based on 1.5 million-year-old fossil footprints at Ileret,
Kenya. Journal of Human Evolution, 64, 556568.
Ditcheld, P., & Harrison, T. (2011). Sedimentology, lithostratigraphy and depositional history of
the Laetoli area. In T. Harrison (Ed.), Paleontology and geology of Laetoli: Human evolution in
context (Vol. 1, pp. 4776). New York: Springer.
Falkingham, P. L., Bates, K. T., Avanzini, M., Bennett, M., Bordy, E. M., Breithaupt, B. H., et al.
(2018). A standard protocol for documenting modern and fossil ichnological data.
Palaeontology, 61, 469480.
Feibel, C. S., Agnew, N., Latimer, B., Demas, M., Marshall, F., Waane, S. A. C., et al. (1996). The
Laetoli hominid footprints A preliminary report on the conservation and scientic restudy.
Evolutionary Anthropology, 5, 149154.
Getty Conservation Institute. (1996). Laetoli project: Conservation of the hominid trackway site at
Laetoli, Tanzania: Report on the 1995 eld season. Los Angeles: Getty Conservation Institute.
Gordon, A. D., Green, D. J., & Richmond, B. G. (2008). Strong postcranial size dimorphism in
Australopithecus afarensis: Results from two new resampling methods for multivariate data sets
with missing data. American Journal of Physical Anthropology, 135, 311328.
Grabowski, M., Hatala, K. G., Jungers, W. L., & Richmond, B. G. (2015). Body mass estimates of
hominin fossils and the evolution of human body size. Journal of Human Evolution, 85,7593.
Harcourt, A. H., & Stewart, K. J. (2007). Gorilla society: Conict, compromise, and cooperation
between the sexes. Chicago/London: University of Chicago Press.
150 M. Cherin et al.
Harmon, E. H. (2006). Size and shape variation in Australopithecus afarensis proximal femora.
Journal of Human Evolution, 51, 217227.
Harrison, T. (2011). Hominins from the Upper Laetolil and Upper Ndolanya Beds, Laetoli. In
T. Harrison (Ed.), Paleontology and geology of Laetoli: Human evolution in context (Vol. 2, pp.
141188). New York: Springer.
Harrison, T., & Kweka, A. (2011). Paleontological localities on the Eyasi plateau, including Laetoli.
In T. Harrison (Ed.), Paleontology and geology of Laetoli: Human evolution in context (Vol.
1, pp. 1745). New York: Springer.
Hatala, K. G., Demes, B., & Richmond, B. G. (2016). Laetoli footprints reveal bipedal gait
biomechanics different from those of modern humans and chimpanzees. Proceedings of the
Royal Society B, 283, 20160235.
Hay, R. L. (1987). Geology of the Laetoli area. In M. D. Leakey & J. M. Harris (Eds.), Laetoli: A
Pliocene site in northern Tanzania (pp. 2347). Oxford: Clarendon.
Hay, R. L., & Leakey, M. D. (1982). The fossil footprints of Laetoli. Scientic American, 246,
5057.
Helm, C. W., McCrea, R. T., Cawthra, H. C., Lockley, M. G., Cowling, R. M., Marean, C. W., et al.
(2018). A new Pleistocene hominin tracksite from the Cape south Coast, South Africa. Scientic
Reports, 8, 3772.
Herlocker, D. J., & Dirschl, H. J. (1972). Vegetation of Ngorongoro conservation area, Tanzania.
Ottawa: Canadian Wildlife Service.
Holdo, R. M., Sinclair, A. R. E., Dobson, A. P., Metzger, K. L., Bolker, B. M., Ritchie, M. E., et al.
(2009). A disease-mediated trophic cascade in the Serengeti and its implications for Ecosystem
C. PLoS Biology, 7, e1000210.
Ichumbaki, E. B., & Mjema, E. (2018). The impact of small-scale development projects on
archaeological heritage in Africa: The Tanzanian experience. Journal of Conservation and
Management of Archaeological Sites, 2,1834.
Ichumbaki, E. B., Cherin, M., Masao, F. T., & Moggi Cecchi, J. (2019). Local peoples interpre-
tations of the hominin footprints at Laetoli, Tanzania. Journal of Community Archaeology &
Heritage, 6, 122138.
James, M. R., Robson, S., & Smith, M. W. (2017). 3-D uncertainty-based topographic change
detection with structure-from-motion photogrammetry: Precision maps for ground control and
directly georeferenced surveys. Earth Surface Processes and Landforms, 42, 17691788.
Johanson, D. C., & White, T. D. (1979). A systematic assessment of early African hominids.
Science, 203, 321330.
Johanson, D. C., White, T. D., & Coppens, Y. (1978). A new species of the genus Australopithecus
(Primates: Hominidae) from the Pliocene of eastern Africa. Kirtlandia, 28,114.
Johnson, K., Nissen, E., Saripalli, S., Arrowsmith, J. R., McGarey, P., Scharer, K., et al. (2014).
Rapid mapping of ultrane fault zone topography with structure from motion. Geosphere, 10,
969986.
Jungers, W. L. (2016). These feet were made for walking. eLife, 5,e22886.
Jungers, W. L., Grabowski, M., Hatala, K. G., & Richmond, B. G. (2016). The evolution of body
size and shape in the human career. Philosophical Transactions of the Royal Society B, 371,
20150247.
Kimbel, W. H., & White, T. D. (1988). Variation, sexual dimorphism and the taxonomy of
Australopithecus. In F. E. Grine (Ed.), Evolutionary history of the RobustAustralopithecines
(pp. 175192). New York: Aldine de Gruyter.
Kohl-Larsen, L. (1943). Auf den Spuren des Vormenschen. Stuttgart: Strecker und Schröder.
Leakey, M. D. (1981). Tracks and tools. Philosophical Transactions of the Royal Society B, 292,
95102.
Leakey, M. D. (1987). Introduction. In M. D. Leakey & J. M. Harris (Eds.), Laetoli: A Pliocene site
in Northern Tanzania (pp. 122). Oxford: Clarendon.
Leakey, M. D., & Harris, J. M. (1987). Laetoli: A Pliocene site in Northern Tanzania. Oxford:
Clarendon.
Leakey, M. D., & Hay, R. L. (1979). Pliocene footprints in the Laetolil Beds at Laetoli, northern
Tanzania. Nature, 278, 317323.
8 Frozen in the Ashes 151
Leakey, M. D., Hay, R. L., Curtis, G. H., Drake, R. E., Jackes, M. K., & White, T. D. (1976). Fossil
hominids from the Laetolil Beds. Nature, 262, 460466.
Lockwood, C. A., Richmond, B. G., Jungers, W. L., & Kimbel, W. H. (1996). Randomization
procedures and sexual dimorphism in Australopithecus afarensis.Journal of Human Evolution,
31, 537548.
Luhmann, T., Robson, S., Kyle, S., & Boehm, J. (2013). Close-range photogrammetry and 3D
imaging. Berlin/Boston: Walter de Gruyter.
Mallison, H., & Wings, O. (2014). Photogrammetry in paleontology: A practical guide. Journal of
Paleontological Techniques, 12,131.
Mancini, A., Capezzuoli, E., Erthal, M., & Swennen, R. (2019). Hierarchical approach to dene
travertine depositional systems: 3D conceptual morphological model and possible applications.
Marine and Petroleum Geology, 103, 549563.
Marteau, B., Vericat, D., Gibbins, C., Batalls, R. J., & Green, D. R. (2017). Application of structure-
from-motion photogrammetry to river restoration. Earth Surface Processes and Landforms, 42,
503515.
Masao, F. T., Ichumbaki, E. B., Cherin, M., Barili, A., Boschian, G., Iurino, D. A., et al. (2016).
New footprints from Laetoli (Tanzania) provide evidence for marked body size variation in early
hominins. eLife, 5, e19568.
McHenry, H. M. (1991). Femoral lengths and stature in Plio-Pleistocene hominids. American
Journal of Physical Anthropology, 85, 149158.
Meldrum, D. J. (2004). Fossilized Hawaiian footprints compared with Laetoli hominid footprints.
In D. J. Meldrum & C. E. Hilton (Eds.), From biped to strider (pp. 6383). New York: Springer.
Menconero, S., Barili, A., Boschian, G., Cherin, M., Ichumbaki, E. B., Iurino, D. A., et al. (2019).
3D survey in extreme environment: The case study of Laetoli hominin footprints in Tanzania. In
W. Börner & S. Uhlirz (Eds.), Proceedings of the 23rd international conference on cultural
heritage and new technologies 2018 (pp. 120). Vienna: Museen der Stadt Wien,
Stadtarchäologie.
Mollel, G. F., Swisher, C. C., III, Feigenson, M. D., & Carr, M. J. (2011). Petrology, geochemistry
and age of Satiman, Lemagurut and Oldeani: Sources of the volcanic deposits of the Laetoli
Area. In T. Harrison (Ed.), Paleontology and geology of Laetoli: Human evolution in context
(Vol. 1, pp. 99119). New York: Springer.
Musiba, C. M., Mabula, A., Selvaggio, M., & Magori, C. C. (2008). Pliocene animal trackways at
Laetoli: Research and conservation potential. Ichnos, 15, 166178.
Musiba, C., Mabulla, A., Mutakyahwa, M., Masao, F. T., Runyoro, V., Yul Kim, J., et al. (2012).
Tanzania and the outstanding universal value of its paleoanthropology: Approaches at Laetoli
and lessons learned. UNESCO World Heritage Series, 33, 115125.
Plavcan, J. M., Lockwood, C. A., Kimbel, W. H., Lague, M. R., & Harmon, E. H. (2005). Sexual
dimorphism in Australopithecus afarensis revisited: How strong is the case for a human-like
pattern of dimorphism? Journal of Human Evolution, 48, 313320.
Raichlen, D. A., & Gordon, A. D. (2017). Interpretation of footprints from Site S conrms human-
like bipedal biomechanics in Laetoli hominins. Journal of Human Evolution, 107, 134138.
Raichlen, D. A., Gordon, A. D., Harcourt-Smith, W. E. H., Foster, A. D., & Haas, W. R. (2010).
Laetoli footprints preserve earliest direct evidence of human-like bipedal biomechanics. PLoS
One, 5, e9769.
Reck, H., & Kohl-Larsen, L. (1936). Erster Überblick über die jungdiluvialen Tier- und
Menschenfunde Dr. Kohl-Larsens im nordöstlichen Teil Des Njarasa-Grabens (Ostafrika)
und die geologischen Verhältnisse des Fundgebietes. Geologische Rundschau, 27, 401441.
Richmond, B., & Jungers, W. L. (1995). Size variation and sexual dimorphism in Australopithecus
afarensis and living hominoids. Journal of Human Evolution, 29, 229245.
Robbins, L. M. (1987). Hominid footprints from site G. In M. D. Leakey & J. M. Harris (Eds.),
Laetoli: A Pliocene site in Northern Tanzania (pp. 497502). Oxford: Clarendon.
Romano, M., Citton, P., Salvador, I., Arobba, D., Rellini, I., Firpo, M., et al. (2019). A
multidisciplinary approach to a unique palaeolithic human ichnological record from Italy
(Bàsura Cave). eLife, 8, e45204.
152 M. Cherin et al.
Rüther, H., Smit, J., & Kamamba, D. (2012). A comparison of close-range photogrammetry to
terrestrial laser scanning for heritage documentation. South African Journal of Geology, 1,114.
Schmid, P. (2004). Functional interpretation of the Laetoli footprints. In D. J. Meldrum & C. E.
Hilton (Eds.), From biped to strider (pp. 4962). New York: Springer.
Stern, J. T., & Susman, R. L. (1983). The locomotor anatomy of Australopithecus afarensis.
American Journal of Physical Anthropology, 60, 279317.
Stumpf, A., Malet, J. P., Allemand, P., Pierrot-Deseilligny, M., & Skupinski, G. (2015). Ground-
based multi-view photogrammetry for the monitoring of landslide deformation and erosion.
Geomorphology, 231, 130145.
Tuttle, R. H. (1987). Kinesiological inferences and evolutionary implications from Laetoli bipedal
trails G-1, G-2/3, and A. In M. D. Leakey & J. M. Harris (Eds.), Laetoli: A Pliocene site in
Northern Tanzania (pp. 503522). Oxford: Clarendon.
Tuttle, R. H. (2008). Footprint clues in hominid evolution and forensics: Lessons and limitations.
Ichnos, 15, 158165.
United Republic of Tanzania. (2004). The environmental management act. Dar es Salaam: Gov-
ernment Printing Press.
Villmoare, B., Hatala, K. G., & Jungers, W. L. (2019). Sexual dimorphism in Homo erectus inferred
from 1.5 Ma footprints near Ileret, Kenya. Scientic Reports, 9, 7687.
Warrick, J. A., Ritchie, A. C., Adelman, G., Adelman, K., & Limber, P. W. (2017). New techniques
to measure cliff change from historical oblique aerial photographs and structure-from-motion
photogrammetry. Journal of Coastal Research, 33,39
Westoby, M. J., Brasington, J., Glasser, N. F., Hambrey, M. J., & Reynolds, J. M. (2012).
Structure-from-motionphotogrammetry: A low-cost, effective tool for geoscience applica-
tions. Geomorphology, 179, 300
White, T. D. (1980). Evolutionary implications of Pliocene hominid footprints. Science, 208,
175176.
White, T. D., & Suwa, G. (1987). Hominid footprints at Laetoli: Facts and interpretations. American
Journal of Physical Anthropology, 72, 485
Zaitsev, A. N., Wenzel, T., Spratt, J., Williams, T. C., Strekopytov, S., Sharygin, V. V., et al.
(2011). Was Sadiman volcano a source for the Laetoli footprint tuff? Journal of Human
Evolution, 61, 121124.
514.
Zaitsev, A. N., Spratt, J., Sharygin, V. V., Wenzel, T., Zaitseva, O. A., & Markl, G. (2015).
Mineralogy of the Laetolil Footprint Tuff: A comparison with possible volcanic sources from
the Crater Highlands and Gregory Rift. Journal of African Earth Sciences, 111, 214221.
Zimmer, B., Liutkus-Pierce, C., Marshall, S. T., Hatala, K. G., Metallo, A., & Rossi, V. (2018).
Using differential structure-from-motion photogrammetry to quantify erosion at the Engare Sero
footprint site, Tanzania. Quaternary Science Reviews, 198, 226241.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
314.
55.
Chapter 9
Steps from History
The Happisburgh Footprints and Their Connections
with the Past
Nick Ashton
Abstract Human footprints were discovered at Happisburgh, UK, in 2013. This
paper describes their discovery and the difculties of recording such enigmatic
remains in a coastal environment. The geological and environmental context in
which they were found is given, together with the evidence of the dating of the
site to either 850,000 or 950,000 years ago. The implications of how humans coped
with long, cold winters of northern Europe is discussed; the evidence of a family
group indicates that seasonal migration is highly unlikely, leaving the possibilities of
either physiological adaptations, such as functional body hair, or the use of technol-
ogies such as shelter, clothing and re. The second part of the paper shows the
various ways in which the footprints have reached wide and diverse audiences
through media reports, exhibitions and books. They show the powerful messages
that footprints can generate through the ideas and emotions that they provoke and the
immediacy of their connection with the deep past.
Keywords Human footprints · Lower Palaeolithic · Early Pleistocene · Britain ·
Europe
Introduction
Footprints tell stories. They can provide information about bipedalism, posture, gait
and stature, as well as on occasion, the sex and age range of a group and activities
being undertaken (e.g. Leakey and Hay 1979; Day and Wickens 1980; Charteris
et al. 1981; Behrensmeyer and Laporte 1981; Bell 2007; Roberts 2008; Bennett et al.
2009; Dingwall et al. 2013). Modern-day trackers can provide new insights into past
N. Ashton (*)
Department of Britain, Europe and Prehistory, British Museum, London, UK
e-mail: nashton@britishmuseum.org
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_9
153
prints through their skills and knowledge (Pastoors et al. 2015,2017). But footprints
also connect a wider public directly to our past. They are the visible trace fossils that
everyone can recognise from life today and therefore have a resonance with a wider
audience than the bones and stones that contribute most of the evidence from our
deep past. The Happisburgh footprints were discovered in 2013. This paper explains
their discovery and the information that can be gleaned from them while reviewing
how these short-lived glimpses of our distant cousins make unexpected connections
154 N. Ashton
with the present.
Background to Happisburgh
Happisburgh is a small village on the northeast coast of Norfolk in the UK (Fig. 9.1).
The cliffs on which the village sits consist of glacial sands, silts and clays that were
deposited by the Anglian Glaciation, which is correlated with Marine Isotope Stage
(MIS) 12, c. 450,000 years ago (450 ka). Beneath the glacial succession lies the
Cromer Forest-bed Formation (CF-bF), which is composed of estuarine, uvial and
alluvial deposits that span the Early and early Middle Pleistocene, between c. 2 and
0.5 million years ago (Ma). The CF-bF outcrops extensively around a 70 km stretch
of coast between Sheringham in the north to Pakeeld in the south. The deposits
include several important interglacial sites famous for Early and early Middle
Pleistocene fossil remains (Reid 1882; West 1980; Preece et al. 2009; Stuart and
Lister 2010; Preece and Partt 2012).
In the last 20 years, there has been accelerated erosion of the coastal cliffs and
the underlying sediment, which has led to increased exposures of the CF-bF and, for
the rst time, the discovery of undisputed Lower Palaeolithic artefacts within the
sediments. Of particular note are Pakeeld, dating to c. 700 ka (Partt et al. 2005),
Happisburgh Site 1 (HSB1) dating to c. 500 ka (Ashton et al. 2008; Lewis et al.
2019) and Happisburgh Site 3 (HSB3), dating to c. 850 ka or possibly c. 950 ka
(Partt et al. 2010; Ashton et al. 2014). This evidence has extended the record of
human occupation of northern Europe by at least 350,000 years and has also
provided important insights into the environments of the early human occupation
in northern latitudes (Candy et al. 2011; Ashton and Lewis 2012).
Happisburgh Site 3
The pre-glacial Pleistocene succession at Happisburgh was rst investigated by Reid
(1882) and more recently by West (1980). West described the sediments exposed at
the base of the cliffs and in the foreshore at a number of locations and also in a
9 Steps from History 155
Fig. 9.1 (a) Map of Britain showing location of Happisburgh; (b) plan of Happisburgh Site
3, exposed and recorded foreshore sediments, location of footprint surface and of borehole HC;
(c) schematic cross-section of recorded sediments from Happisburgh Site 3 through to borehole HC
showing stratigraphic position of footprint surface. (Illustration C. Williams)
borehole near the former slipway on to the beach (Fig. 9.1). This borehole
(HC) demonstrated a sequence of laminated silts and sands beneath the Happisburgh
Till. Palynological data from the laminated silts indicated an interglacial vegetational
succession, which West attributed to a late stage of the Early Pleistocene.
Happisburgh Site 3 was discovered in 2005, some 330 m to the northeast of
Borehole HC, while undertaking a coastal survey of CF-bF deposits on a 3 km
stretch of coast to the north of Happisburgh. Seasonal excavation until 2012 revealed
a series of deposits that relate to those of West in Borehole HC (Partt et al. 2010).
At Site 3 they consist of a series of estuarine sands and silts which inll channels.
The channels have a lag gravel at their base up to 0.2 m in thickness, from which an
156 N. Ashton
artefactassemblage has been recovered, consisting of c. 80 int akes, ake tools
and cores, all in remarkably fresh condition.
The sediments also contain a rich assemblage of fauna and ora (Partt et al.
2010). Pollen, wood and other plant remains indicate a regional vegetational suc-
cession that had changed from deciduous woodland to coniferous forest. The more
localised environment can be reconstructed from study of the insect remains
suggesting a oodplain that consisted of a mosaic of grassland, stands of alder,
small pools and marsh. This is supported by grassland pollen recovered from a
hyaena coprolite. The vertebrate remains include part of the skull of European
sturgeon (Acipenser sturio), which today spawn in deep-water estuaries. Together
with other indicators of brackish water and the interpretation of the laminated silts
and sands, the evidence suggests that the site was in the upper reaches of an estuary
of a large river. Other vertebrate fauna includes giant elk (Cervalces latifrons), red
deer (Cervus elaphus) and an extinct form of horse (Equus suessenbornensis),
alongside larger herbivores such as an early form of mammoth (Mammuthus
meridionalis) and hippopotamus (Hippopotamus amphibius).
The shift in the vegetational succession towards coniferous forest suggests that
the site dates towards the end of an interglacial with a cooler climate (Partt et al.
2010). This is supported by the beetle remains, which indicate that summer temper-
atures were similar to East Anglia today with an average of about 17 C. But the
winters were between 3 and 0 C, whereas today the average is 4 C. A modern-
day analogue would be southern Scandinavia or Denmark, which would have made
winters a challenge to survive and prompts questions about the level of technology in
terms of shelter, clothing and re (Ashton and Lewis 2012).
The age of the site is constrained by the overlying glacial sediments, which
indicate that it is older than 450 ka. A reversed palaeomagnetic signal suggests
that the site predates the Brunhes-Matuyama boundary at 780 ka and is Early
Pleistocene in age. Renement of the age can be determined by the mammalian
fossils; Mammuthus meridionalis is known to have become extinct about 800 ka,
and the horse, Equus Suessenbornensis, also became extinct about this time, both of
which support the evidence from the palaeomagnetics. A maximum age can be
determined from the extinct giant elk, Cervalces latifrons, and red deer, Cervus
elaphus, which rst evolved about a million years ago. The pollen suggests a date
towards the end of an interglacial, with the two most likely stages being MIS 21 at
850 ka or MIS 25 at 950 ka (Partt et al. 2010).
9 Steps from History 157
The Footprint Surface
Fieldwork at Happisburgh continued after the excavations were completed in 2012,
with funding by English Heritage (now Historic England), through a programme of
geophysical and coring surveys to understand better the distribution of CF-bF
sediments on the foreshore and inland and whether evidence of their survival
could be found offshore through further survey and diving (Ashton et al. 2018). In
early May 2013, during the survey work, an area of laminated silts was exposed
c. 110 m northwest of borehole HC and c. 140 m of the excavations of Site
3 (Figs. 9.2 and 9.3; Ashton et al. 2014). The laminated sediments could be traced
laterally between the three locations. Although the exact stratigraphic relationships
Fig. 9.2 View of footprint surface cliff top looking south. (Photo M. Bates)
Fig. 9.3 View of footprint
surface looking north.
(Photo M. Bates)
remain uncertain, they are of a similar age and, based on Site 3, date to either 850 ka
or 950 ka.
158 N. Ashton
In the new exposure, beach sand had been removed by the sea, and the laminated
sediments were subject to wave erosion. When exposed, the bedding surfaces
provide natural planes of weakness, and the washing out of sandy laminae results
in the removal of layers of laminated silts and the exposure of new, undisturbed
bedding surfaces. In most cases, these surfaces are at or gently undulating and
display ripple structures formed during the original deposition of the sediments.
However, one horizon had very different surface characteristics where a series of
hollows ranging from circular to elongate in outline were visible over an area of
c. 12 m
2
. The elongate hollows were generally 3050 mm in depth, 140250 mm in
length and 60110 mm in width. The visual similarity to Holocene footprint surfaces
prompted more detailed investigation of this horizon. However, the surface was
located in the intertidal zone and was prone to rapid destruction by wave action or to
reburial as the beach was re-established. The situation presented particular chal-
lenges for recording and analysis of the features and prevented either lifting of the
footprint surface as sediment blocks or standard casting of moulds of the surface.
Initially it was hoped to laser scan the surface, but availability of equipment was a
problem. However, a relatively new technique of multi-image photogrammetry was
just beginning to be more widely used in archaeology, and so, with the expertise of
Sarah Duffy from the University of York, a team was mobilised, and this method
was used a few days after discovery.
Multi-image photogrammetry simply uses a series of digital images of an object
or surface with xed points, taken from different angles, which when combined with
specialist software creates a 3D model (Fig. 9.4). The principle was ne, but the
practicality was more difcult. The combination of tides, blown beach sand, weather
conditions and time constraints made recording the surface extremely difcult. Prior
to recording, water was used to wash away the beach sand that had been deposited
during previous high tides, though it was impossible to completely clear the surface
and remove all water from the hollows due to persistent rain. Field measurement of
the hollows was not possible because of the time constraints, but multi-image
photogrammetry proved to be an effective method for rapid recording of the surface
features and allowed subsequent metric analysis of footprint shape and size,
although estimates of depth were more problematic. Laser scanning was also
attempted a week later, but by this time, the features had become severely eroded
through successive tidal cycles, and by the end of May 2013, they had been
completely removed (Ashton et al. 2014).
After recording, the rst task was to determine the agency responsible for their
formation. The possibility of them being recent footprints was immediately ruled out
due to the hardness and compaction of the laminated sediments. Walking across
similar sediments has little impact even in heavy boots. Extensive searches were
made for natural erosive agencies that might be responsible, but none of the hollows
were consistent with the range of processes that are normally found in an estuarine
environment. After initial scepticism and careful scrutiny of the evidence, it was
9 Steps from History 159
Fig. 9.4 Model of footprint
surface produced from
photogrammetric survey.
(Modelling S. Duffy)
concluded that the hollows were indeed ancient human footprints (Ashton et al.
2014).
The surface was analysed using vertical images produced from the multi-image
photogrammetry. Depth measurements were not possible as water or sand was often
retained in the base of the prints. A total of 152 hollows were measured, and this
revealed that the lengths, widths and width/length ratios were consistent within the
expected range of juvenile and adult hominin footprints (Ashton et al. 2014). In
some cases, left or right and front or back of the foot were also apparent, including
two instances of toes, providing information about direction of movement. The less
elongated features were also potentially hominin footprints, where impressions from
just heels or the front of feet were preserved, or overprinting had obscured original
features. The time elapsed from initial exposure to recording also led to some erosion
of the surface, which affected the shape and clarity of the prints.
More detailed analysis by Isabelle de Groote, from Liverpool John Moores
University, was limited to 12 prints where complete outlines could be clearly
identied (Ashton et al. 2014). They were thought to indicate at least ve individuals
with foot lengths between 140 and 260 mm. Based on recent populations, stature can
be estimated from foot length using a ratio of 0.15 for foot length/stature (Dingwall
et al. 2013). Fossil skeletal evidence suggests that body proportions of Middle
Pleistocene hominins were similar to modern humans, and therefore this ratio can
also be applied to past populations (Carretero et al. 2012; Pablos et al. 2012). The
0.15 ratio suggested a height range for the Happisburgh hominins of between 0.93
and 1.73 m, indicating the presence of adults and children. For the orientation
studies, a larger dataset of 49 prints was analysed, showing a preferred south-north
orientation. In 29 cases where the arch and the front/back of the foot could be
identied, the direction of movement was also assessed, showing a preferred direc-
tion of movement to the south.
160 N. Ashton
Unfortunately, there are no human fossils from Britain that date to this period, but
the closest comparison is Gran Dolina (TD6) at Atapuerca in northern Spain where
bones and teeth dating to c. 800 ka have been named as Homo antecessor or Pioneer
Man (Carbonell et al. 2005,2008). This attribution has recently been examined using
2D morphometrics on a range of footprints from Pliocene, Pleistocene and Holocene
sites (Wiseman et al. 2020). They conclude that the dimensions of the Happisburgh
footprints were most similar to the H. erectus footprints from Ileret in Kenya. This
conforms with an attribution of the Happisburgh hominins to H. antecessor, thought
to be a European cousin of H. erectus. If this attribution is correct, then estimates of
stature from the fossil bones from Gran Dolina TD6 can be compared to
Happisburgh. The tali recovered from TD6 show a mean stature of 1.73 m for
males and 1.68 m for females (Pablos et al. 2012). This would suggest that the
tallest individual at Happisburgh was an adult male with the smaller footprints being
produced by either adult females or juveniles and by children. An obvious interpre-
tation is that the Happisburgh footprints were left by a family group.
The search for further footprints at Happisburgh has been difcult. Excavation is
not practical as the deposits extend for several hundred metres and are up to 2 m in
depth with multiple horizons. The chance of selecting the right area and encounter-
ing a footprint surface is minimal. In fact the sea is the best excavator through the
twice daily peeling off of surfaces in an impartial way. Since 2014 there have been
periodic exposures of the laminated silts between Site 3 and Borehole HC, and on
occasion there have been reports of possible footprints. Often by the time a visit to
the site is made, the prints have either eroded away or have been buried by beach
sand. In one case, a small exposure was revealed during a eld visit and a record
made (pers. comm. Simon Lewis). The most successful approach has been through
several local collectors and trained amateurs, who equipped with GPS have been
able to collect, record and report new artefacts and fossils and also alert us to any
new exposures with potential prints. On-the-spot photography with multiple images
is encouraged, and this will hopefully capture enough information of any future
footprint surfaces before they are eroded away.
9 Steps from History 161
Implications of the Happisburgh Footprints
Questions remain about how the Happisburgh hominins survived the long, cold
winters of northern Europe. One suggestion is that they seasonally migrated. How-
ever to make any appreciable difference to winter temperatures, they would have had
to have travelled to coastal areas of southern Europe. This might have been feasible
for adult hunting groups, but the evidence from Happisburgh shows the presence of
children. Such a journey would have been virtually impossible as a family group.
The implication from the footprints is that the humans were residents and surviving
the long, cold winters.
An alternative option for survival was that the Happisburgh humans had func-
tional body hair that gave them sufcient protection from the cold. The favoured
hypothesis that hominins lost their body hair over several million years in open,
equatorial areas of Africa deserves re-examination (Wheeler 1984,1991,1992). The
argument goes that with bipedalism, there was less need for protection from the sun,
leading to a reduction in body hair, other than the scalp. One of the evolutionary
advantages was better thermoregulation through more efcient sweat glands, which
also enabled longer day-time hunting. This may have been the case, but there is no
direct proof. It may have had advantages for Africa, but there were serious short-
comings for the more seasonal climates of Europe. So perhaps humans entering
Europe from Africa still had body hair, or it redeveloped as they evolved in more
northerly latitudes.
The simplest answer to how the humans coped with cooler climates is that they
had better control of re and were more capable of making clothes and shelters than
previously thought. Unfortunately there is no evidence for the use of these technol-
ogies at this time. Better evidence for ways of buffering against the cold start to be
introduced from around 500 ka. At High Lodge in Suffolk, there are scrapers that
were ideal tools for processing hides, presumably for building simple shelters or use
as clothing (Ashton et al. 1992). From 400 ka at Beeches Pit, also in Suffolk, or
Menez Dregan in Brittany, there are distinct hearths from res (Gowlett et al. 2005;
Preece et al. 2007; Ravon 2018). If earlier evidence is to be found, then Happisburgh
with its rich organic preservation is an obvious place to look.
If the footprint evidence is correct and the humans were all-year residents, then
perhaps the biggest challenge was the short growing season of northern latitudes
(Ashton 2015; Hoseld 2016). This implied a greater dependence on meat and more
effective scavenging or possibly hunting. If meat acquisition was a struggle, what
other resources were available? The big advantage for Happisburgh was its estuary
situation, providing important resources such as collectable shellsh and seaweed
over the difcult winter months. Perhaps these pioneering populations were able to
cope in northern Europe but only in coastal or estuary situations (Cohen et al. 2012).
162 N. Ashton
Impact of the Happisburgh Footprints
The evidence that the footprints provide of a family group, wandering along the edge
of an estuary, not only received academic attention as the oldest footprints outside
Africa but drew wide appreciation from public audiences around the globe. Here was
the story of a family group, somehow surviving the cold winters of northern Europe
at almost a million years ago. The footprints were published in February 2014 to
coincide with an exhibition Britain: One Million years of the Human Story at the
Natural History Museum in London, where Happisburgh began the story with video
footage of the footprint discovery (Ashton et al. 2014; Dinnis and Stringer 2014). A
British Museum press release on February 7th, a few hours before the publication,
led to widespread coverage by all UK television and radio networks, as well as many
abroad, with an astonishing 250 newspaper reports around the world.
But news is short-lived, so it is more gratifying to see how the footprints have
endured in other, sometimes unexpected, ways. The footprints prompted mention in
several books. One of the more popular accounts has been in The Road to Little
Dribbling: More Notes from a Small Island by Bill Bryson, who visited
Happisburgh and described the footprints on his return journey around Britain
(Bryson 2015). A more unusual project was undertaken by the Dutch radio broad-
caster and writer, Mathijs Deen, who in Over Oude Wegen: Een Reis door de
Geschiedenis van Europa (Down Old Roads: A Journey through Europes History,
Deen 2018) explores by car the famous ancient journeys and stories along routeways
that still connect us today. The rst journey that he describes was that potentially
taken by Homo antecessor from Atapuerca in northern Spain to Happisburgh.
This year, a beautifully written book, Time Song, was published by Julia Black-
burn (2019). It interleaves her own stories and encounters, with thoughts on
Doggerland, the now vanished prehistoric landscape that lies beneath the North
Sea. She reects on the Happisburgh footprints and the people who left them behind.
They feature in 2 of the 18 Time Song poems, with the rst about their discovery:
The weather was bad.
Rain falling,
Waves crashing.
Over the next two weeks.
The hollows were photographed.
and scanned with lasers,
Before they vanished,
Leaving no trace.
One hundred and two footprints.
Twelve of them complete,
Indicating ve individuals.
Of different ages:
A little human group.
Moving in a southerly direction.
Across the mudats.
Of a large tidal river,
9 Steps from History 163
Between eight hundred and fty.
And nine hundred and fty.
Thousand.
Years ago.
Making a further jump back.
In the history of habitation.
In this country,
Now called England.
(Blackburn 2019:6
263)
A very different topic has been covered by Antonia Malchik in her thought-
provoking book, A Walking Life: Reclaiming Our Health and Our Freedom One
Step at a Time (Malchik 2019). The book examines from an American perspective
the car-centric culture in which we live and the barriers imposed by a modern world
on the freedom to walk. Bipedalism, she argues, is one of the characteristics that
make us human, represented in part by the footprints from Happisburgh. Walking is
essential for our health, a powerful means to rehabilitation and important for societal
welfare. The interactions that it brings can encourage and bind communities, in
contrast to the often cocooned, loneliness of suburban, motor-driven life.
She devotes much of the nal chapter, Meandering, to her visit to Happisburgh,
describing her own journey and the signicance of the discovery. One section
particularly struck me, where she comments on a deep, underlying lesson from the
Happisburgh footprints: the importance of meandering, rather than efciency, for
learning.
The footprints werent in a straight line,Ashton told me when we met at his British
Museum ofce. Not being in a straight line was a criticism other researchers had levelled at
the nd. But to him, the wandering nature of the footprints made complete sense. Because
the Happisburgh footprints included children. This wasnt just a temporary hunting party, a
group moving through seasonally. These people were living there.
The Laetoli footprints are in a straight line, and its easy to imagine those hominins some
three or four million years ago walking across the savannah, heading ... where? The
Happisburgh footprints, though, give us movement and life, images of children veering
off to poke in the mud, chase some small animal or crustacean, or peer at a plant, just as my
children did at that age. Just as the infants in Karen Adolphs lab do, roaming around in the
most inefcient manner possible because that is how we grow and explore and learn.
(Malchik 2019: 204)
Is this another human characteristic, at times forgotten from our childhood that
of curiosity?
Connections to the present were also made through a small, but powerful,
exhibition at the British Museum in April 2017. Called Moving Stories, the exhibi-
tion drew together three journeys about migration. The rst was the million-year-old
journey told by footprints discovered at Happisburgh. A life-size image of the
footprint surface was projected onto the oor, and visitors were encouraged to step
into the prints of their distant relatives (Fig. 9.5). The surface was animated with
water ebbing and owing across the surface an image that would have been
apparent at the time of their creation over 850,000 years ago, as well a view that
we had on their discovery. A muted soundtrack took the visitor back to Happisburgh
through the sounds of gently owing water, with the cries and calls of estuary birds
164 N. Ashton
Fig. 9.5 Moving Stories exhibition at the British Museum, April 2017. The footprint surface is
projected onto the oor and gently animated with owing water. It has its own space within a
shipping container, representing modern migration, with a window onto the second story of the
pictorial diary, Alis Boat. (Photo British Museum)
and background chatter of children and their parents. The exhibition explained that
this was a metaphorical journey that had taken the family beyond the natural
boundaries of the known world. But it was this and similar journeys that eventually
led to adaptation to more difcult environments through better provision of basic
human needs: food, clothing, shelter and re.
The Happisburgh journey was juxtaposed against the heart-wrenching story of
exile in the contemporary work, Alis Boat, by Iraqi artist Sadik Kwaish Alfraji. Alis
Boat is a pictorial diary that tells the story of a young boy wishing to escape the
horrors of present-day Iraq. It is inspired by an encounter with his 11-year-old
nephew who, on Sadiks departure from Iraq to the Netherlands in 2009, gave him
a drawing of a boat, with the words I wish this boat takes me to you. The
exhibition showed how in the present day many migrants still have a quest for the
same basic human needs of food, warmth and shelter but with boundaries and
barriers drawn by politics rather than geography.
The third journey of Moving Stories was told through the work of Édouard
Glissant, a poet and philosopher from Martinique, about the slave trade diasporas
from Africa. Remarkably his work offers a positive outlook, suggesting that
although migrants may lose their social and cultural unity, they gain cultural
diversity and multiplicity; importantly differences can unite, rather than divide and
are the means to build global communities. The underlying lesson of all three
journeys was that migration not only brings hardship but also opportunities and it
is an inherent human trait that goes back to the deep past.
9 Steps from History 165
Conclusion
It is quite astonishing how a few simple footprints beneath a beach in Norfolk can
invoke such wide and varied reaction and generate such profound thoughts and
ideas. Many people are simply awe-inspired by their age, brief appearance and the
serendipity of being spotted and recorded. For others they are a journey by pioneers,
pushing the boundaries of the known world, or one to the unknown world through
their enigmatic appearance and disappearance as a brief glimpse into the now
drowned landscapes of the North Sea. They are both past and present. The impor-
tance of history is what it can tell us about today or tomorrow or the thoughts and
emotions that it can provoke. For most, the power of Happisburgh lies in the family
group and the everyday story of children playing at the waters edge. I admire the
eyes, knowledge and skills of the modern tracker and envy their ability to interpret a
different world. I lack their skills, but I do have my own vision of Happisburgh some
850,000 years ago.
The tide was gently rising as the family group picked their way around the shallow pools on
the mudats of the estuary edge. They paused to watch a herd of horse grazing near the
reedswamp on the far bank. A lone rhinoceros and three mammoths could be seen as
silhouettes in the distance. The parents watched warily and rather enviously as a cackle of
hyaenas greedily tore apart the esh of an elk, little more than a stones throw away. They
had been beaten by their competitors to the injured animal. Today their family would survive
on the plant roots and shellsh they had eaten earlier. Tomorrow there might be other
opportunities before the lions, wolves and hyaenas took their share. The three children
seemed oblivious to the danger, splashing about bare-foot at the waters edge. The older boy
realised the risk and encouraged them on; they had to reach the safety of the deep pine forest
before dusk. The sun was sinking fast and a chill breeze rippled across the water as a skein of
geese took ight. The family continued on their way leaving trails of footprints in the estuary
muds. (Ashton 2017:1)
Acknowledgements I would like to thank North Norfolk District Council for permission to
undertake the work and the people of Happisburgh for their continued support. I also thank Martin
Bates, Richard Bates, Sarah Duffy, Isabelle De Groote, Peter Hoare, Simon Lewis, Simon Partt,
Chris Stringer and Craig Williams for their contributions to the eldwork and subsequent analyses. I
am very grateful to Julia Blackburn, Martijn Deen and Anita Malchik for conversations and their
writings, which have taken the footprints to new levels of meaning. Finally I thank the Calleva
Foundation for support of the Pathways to Ancient Britain Project of which the research at
Happisburgh forms a part.
References
Ashton, N. M. (2015). Ecological niches, technological developments and physical adaptations of
early humans in Europe: The handaxe-heildelbergensis hypothesis. In F. Wenban-Smith,
F. Coward, & R. Hoseld (Eds.), Settlement, Society and Cognition in Human Evolution:
Landscapes in Mind (pp. 138153). Cambridge: Cambridge University Press.
Ashton, N. M. (2017). Early Humans. London: HarperCollins.
Ashton, N. M., Cook, J., Lewis, S. G., & Rose, J. (1992). High lodge: Excavations by G. de
G. Sieveking 196268 and J. Cook 1988. London: British Museum Press.
166 N. Ashton
Ashton, N. M., & Lewis, S. G. (2012). The environmental contexts of early human occupation of
north-west Europe: The British Lower Palaeolithic record. Quaternary International, 271,
5064.
Ashton, N. M., Partt, S. A., Lewis, S. G., Coope, G. R., & Larkin, N. R. (2008). Happisburgh site
1 (TG388307). In I. Candy, J. R. Lee, & A. M. Harrison (Eds.), The Quaternary of Northern
East Anglia (pp. 151156). London: Quaternary Research Association.
Ashton, N. M., Lewis, S. G., De Groote, I., Duffy, S., Bates, M. B., et al. (2014). Hominin footprints
from Early Pleistocene deposits at Happisburgh, UK. PLoS One, 9(2), e88329. https://doi.org/
10.1371/journal.pone.0088329.
Ashton, N. M., Lewis, S. G., Partt, S. A., Bates, M. R., Bates, C. R., et al. (2018). Understanding
and Monitoring the Cromer Forest-bed Formation. Historic England research. Report series
no. 622018.
Bell, M. (Ed.). (2007). Prehistoric Coastal Communities: The Mesolithic in Western Britain
(Council for British Archaeology Research Report 149). York: Council for British Archaeology.
Behrensmeyer, A. K., & Laporte, L. F. (1981). Footprints of a Pleistocene hominid in northern
Kenya. Nature, 289, 167169.
Bennett, M. R., Harris, J. W. K., Richmond, B. G., Braun, D. R., Mbua, E., et al. (2009). Early
hominin foot morphology based on 1.5-million-year-old footprints from Ileret, Kenya. Science,
323, 11971201.
Blackburn, J. (2019). Time song. London: Penguin.
Bryson, B. (2015). The Road to Little Dribbling. More Notes from a Small Island. London:
Doubleday.
Candy, I., Silva, B., & Lee, J. R. (2011). Climates of the early Middle Pleistocene in Britain:
Environments of the earliest humans in northern Europe. In N. M. Ashton, S. G. Lewis, & C. B.
Stringer (Eds.), The Ancient Human Occupation of Britain Project (pp. 1121). Amsterdam:
Elsevier.
Carbonell, E., Bermudez de Castro, J. M., Arsuaga, J. L., Allue, E., Bastir, M., et al. (2005). An
early Pleistocene hominin mandible from Atapuerca-TD6, Spain. PNAS, 102, 56745678.
Carbonell, E., Bermúdez de Castro, J. M., Parés, J. M., Pérez-González, A., Cuenca Bescós, G.,
et al. (2008). The rst hominin of Europe. Nature, 452, 465469.
Carretero, J. M., Rodríguez, L., García-González, R., Arsuaga, J. L., Gómez-Olivencia, A., et al.
(2012). Stature estimation from complete long bones in the Middle Pleistocene humans from the
Sima de los Huesos, Sierra de Atapuerca (Spain). Journal of Human Evolution, 62, 242255.
Charteris, J., Wall, J. C., & Nottrodt, J. W. (1981). Functional reconstruction of gait from the
Pliocene hominid footprints at Laetoli, northern Tanzania. Nature, 290, 496498.
Cohen, K. M., MacDonald, K., Joordens, J. C. A., Roebroeks, W., & Gibbard, P. L. (2012). The
earliest occupation of north-west Europe: A coastal perspective. Quaternary International, 271,
7083.
Day, M. H., & Wickens, E. H. (1980). Laetoli Pliocene hominid footprints and bipedalism. Nature,
286,385387.
Deen, M. (2018). Over Oude Wegen: Een Reis door de Geschiedenis van Europa. Amsterdam:
Thomas Rap.
Dingwall, H. L., Hatala, K. G., Wunderlich, R. E., & Richmond, B. G. (2013). Hominin stature,
body mass, and walking speed estimates based on 1.5 million-year-old fossil footprints at Ileret,
Kenya. Journal of Human Evolution, 64, 556568.
Dinnis, R., & Stringer, C. B. (2014). Britain: One Million Years of the Human Story. London:
Natural History Museum.
Gowlett, J. A. J., Hallos, J., Hounsell, S., Brant, V., & Debenham, N. C. (2005). Beeches Pit
Archaeology, assemblage dynamics and early re history of a Middle Pleistocene site in East
Anglia, UK. Eurasian Prehistory, 3,338.
Hoseld, R. T. (2016). Walking in a winter wonderland? Strategies for Early and Middle Pleisto-
cene survival in mid-latitude Europe. Current Anthropology, 57, 653682.
9 Steps from History 167
Leakey, M. D., & Hay, R. L. (1979). Pliocene footprints in the Laetoli Beds at Laetoli, northern
Tanzania. Nature, 278, 317323.
Lewis, S. G., Ashton, N. M., Field, M. H., Hoare, P. G., Kamermans, H., et al. (2019). Human
occupation of northern Europe in MIS 13: Happisburgh Site 1 (Norfolk, UK) and its European
context. Quaternary Science Reviews, 211,3458.
Malchik, A. (2019). A Walking Life: Reclaiming our Health and our Freedom One Step at a Time.
New York: Da Capo Press.
Pablos, A., Lorenzo, C., Martínez, I., Bermúdez de Castro, J. M., Martinón-Torres, M., et al. (2012).
New foot remains from the Gran Dolina-TD6 Early Pleistocene site (Sierra de Atapuerca,
Burgos, Spain). Journal of Human Evolution, 63, 610623.
Partt, S. A., Barendregt, R. W., Breda, M., Candy, I., Collins, M. J., et al. (2005). The earliest
record of human activity in northern Europe. Nature, 438, 10081012.
Partt, S. A., Ashton, N. M., Lewis, S. G., Abel, R. L., Coope, G. R., et al. (2010). Early Pleistocene
human occupation at the edge of the boreal zone in northwest Europe. Nature, 466, 229233.
Pastoors, A., Lenssen-Erz, T., Ciqae, T., Kxunta, U., Thao, T., et al. (2015). Tracking in caves:
Experience based reading of Pleistocene human footprints in French caves. Cambridge Archae-
ological Journal, 25, 551564.
Pastoors, A., Lenssen-Erz, T., Breuckmann, B., Ciqae, T., Kxunta, U., et al. (2017). Experience
based reading of Pleistocene human footprints in Pech-Merle. Quaternary International, 430,
155162.
Preece, R. C., & Partt, S. A. (2012). The Early and early Middle Pleistocene context of human
occupation and lowland glaciation in Britain and northern Europe. Quaternary International,
271,628.
Preece, R. C., Partt, S. A., Bridgland, D. R., Lewis, S. G., Rowe, P. J., et al. (2007). Terrestrial
environments during MIS 11: Evidence from the Palaeolithic site at West Stow, Suffolk,
UK. Quaternary Science Reviews, 26, 12361300.
Preece, R. C., Partt, S. A., Coope, G. R., Penkman, K. E. H., Ponel, P., et al. (2009). Biostrati-
graphic and aminostratigraphic constraints on the age of the Middle Pleistocene glacial succes-
sion in North Norfolk, UK. Journal of Quaternary Science, 24, 557580.
Ravon, A.-L. (2018). Land use in Brittany during the Middle Pleistocene. The example of the
persistent place of Menez-Dregan I (Plouhinec, Finistère). In M. Pope, J. McNabb, & C. S.
Gamble (Eds.), Crossing the Human Threshold (pp. 106122). London: Routledge.
Roberts, D. L. (2008). Last interglacial hominid and associated vertebrate fossil trackways in
coastal eolianites, South Africa. Ichnos, 15, 190207.
Reid, C. (1882). The Geology of the Country around Cromer. London: HMSO.
Stuart, A. J., & Lister, A. M. (Eds.). (2010). The West Runton elephant and its Cromerian
environment. Quaternary International, 228,1248.
West, R. G. (1980). The Pre-glacial Pleistocene of the Norfolk and Suffolk Coasts.Cambridge:
Cambridge University Press.
Wheeler, P. E. (1984). The evolution of bipedality and loss of functional body hair in hominids.
Journal of Human Evolution, 13,9198.
Wheeler, P. E. (1991). The thermoregulatory advantages of hominid bipedalism in open equatorial
environments: The contribution of increased convective heat loss and cutaneous evaporative
cooling. Journal of Human Evolution, 21, 107115.
Wheeler, P. E. (1992). The inuence on the loss of functional body hair on the water budgets of
early hominids. Journal of Human Evolution, 23, 379388.
Wiseman, A. L. A., Stringer, C. B., Ashton, N. M., Bennett, M. R., Hatala, K. G., et al. (2020). The
morphological afnity of the early Pleistocene footprints from Happisburgh, England with other
tracks of Pliocene, Pleistocene and Holocene age. Journal of Human Evolution, 144, 102776.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
168 N. Ashton
Chapter 10
Reconsideration of the Antiquity
of the Middle Palaeolithic Footprints from
Theopetra Cave (Thessaly, Greece)
Nina Kyparissi-Apostolika and Sotiris K. Manolis
Abstract During the 1996 eld season, four footprints were found in undisturbed
deposits at the borders of squares Θ10-I10 at a depth of 3.5 m at the Theopetra Cave
excavation site. The footprints lie adjacent to an ash horizon that has been dated to ca
~135 ka. Two footprints in the trail are complete and measure 150.4 mm and
138.96 mm in length. Based on modern European standards, these lengths would
be consistent with young children aged between 2 and 4 years old and 90100 cm in
stature. The two complete footprints, which follow each other in the trail, appear
both to have been left feet. The partial print, which immediately precedes the two
complete prints in the series, also appears to have been by a left foot. This suggests
that what initially seems to be a single trail is actually a composite of two or more
trails of prints. This hypothesis is supported by the different characteristics of the two
complete prints. One is consistent with a bare foot and clearly shows the impressions
of the toes, ball, arch and heel. The other is characterized by a simpler contour and is
more sharply dened and indicates that the individual was wearing some kind of foot
covering. An important question is what kind of hominid made the footprints? These
footprints may have been made by Neanderthals or early Homo sapiens, based on
thermoluminescence dating results.
Keywords Footprints · Early Homo sapiens · Neanderthals · Middle Pleistocene ·
Theopetra Cave · Europe · Greece
N. Kyparissi-Apostolika (*)
Ephorate of Palaeoanthropology and Speleology, Ministry of Culture, Athens, Greece
e-mail: nkyparissi@hotmail.com
S. K. Manolis
Department of Animal and Human Physiology, Faculty of Biology, National and Kapodistrian
University of Athens, Athens, Greece
e-mail: smanol@biol.uoa.gr
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_10
169
Introduction
Theopetra Cave (214004600E, 394005100N) is a unique prehistoric site in central
Greece (Fig. 10.1), as several cultural periods (Middle and Upper Palaeolithic,
Mesolithic and Neolithic) are represented (Kyparissi-Apostolika 1998,1999;
Facorellis et al. 2001; Karkanas 2001).
170 N. Kyparissi-Apostolika and S. K. Manolis
Excavations at Theopetra Cave have produced many signicant anthropological
ndings, among which the prehistoric footprints are distinctive. During the 1996
eld season, four footprints were found in undisturbed deposits in square Θ10 at a
depth of 3.5 m (Fig. 10.2) (Manolis et al. 2000). The footprints lie adjacent to an ash
horizon that has been dated to ca ~130 ka (new date after Valladas et al. 2007). These
Fig. 10.1 Map of Greece showing the site of Theopetra Cave
10 Middle Palaeolithic Footprints from Greece 171
Fig. 10.2 The footprints found in the Theopetra Cave
footprints were found among others printed in an extensive ashy and burnt area of
ve serial trenches in axis 10 of the excavation grid (Z10, H10, Θ10, I10, K10), 16 to
20 m
2
. Most of the remaining footprints rather belong to animals, but the presence of
a few more human ones among them cannot be excluded (Fig. 10.2).
In palaeoanthropology it is well understood that we cannot know in detail the
behaviour of upright body posture and bipedalism. Our inferences concerning the
shaping of the body and the way of walking are derived from comparing data after
following a logical sequence of thoughts. These are based mostly on the hypothesis
of the homology of various anatomic characteristics.
The footprints are an undeniable proof of the existence of hominids and/or
prehistoric humans in general, and we can draw from them very important informa-
tion about locomotion and composition of the group, through comparative and
experimental methods.
Such information might be (after Day 1991):
Morphological: Because in palaeoanthropology the ndings are mostly hard
fossilized bones the discovery of footprints may give us information about the
shape of the soft parts of the foot and the size and morphology of its anatomical
picture. This includes the position of the metatarsal area, the heel, the toes, their
prominence and the presence of an arch. The measurements allow us to estimate
the size and compare it to that of other known human populations. The stature
may also be estimated from world data where the length of the foot is approxi-
mately 15% of the height.
Behavioural: By studying the footprints we can make inferences about the way of
walking, whether the individual was running at the time the footprints were made,
and we can also estimate the pace. Of course, this is not always possible, because
it is required that the series of steps lies undisturbed, like in the case of Laetoli,
Tanzania.
Environmental: Together with the human footprints there could be others made
by animals, in which case we will be able to assume if they are contemporary with
172 N. Kyparissi-Apostolika and S. K. Manolis
each other. Related botanical data will allow us to recreate the climatic and
environmental conditions at the given time.
Historical Background
The most ancient footprints in the world are the ones discovered in Laetoli, Tanza-
nia, and have been dated to 3.7 Ma. Therefore, when footprints are discovered, some
of the questions that need to be answered can be: Who were the hominids that
created them? Were they male or female and of what age? How tall were they? How
much did they weigh? Some of these questions may be answered, but comparative
data are required. Next, we will see which of these we have been able to answer
through our research so far.
The most ancient footprints in Europe were found at Happisburgh, UK, dated to
the Early Pleistocene (ca ~10.78 Ma) (Ashton et al. 2014). The following caves or
open-air sites have been dated to the Middle Pleistocene: Roccamonna, Italy,
325,000385,000, open air (Avanzini et al. 2004,2008); Terra Amata, France,
300,000400,000, open air (de Lumley 1966,1967); and Theopetra Cave, Greece,
ca 130,000, cave (Manolis et al. 2000).
Late Pleistocene sites are caves such as the Vârtop Cave in Romania dated 62 ka
(Onac et al. 2005); the Grotta del Cavallo in Italy (the rst known appearance of
anatomically modern humans in Europe, ~44 ka (Benazzi et al. 2011); and several
caves in France such as Lascaux, Niaux, Aldene, Peche Merle, Fontanet, Ariège and
Chauvet, Bàsura in Italy and Ojo Guareña in Spain (Lockley et al. 2008). All these
sites except the Vârtop Cave and Grotta del Cavallo have been dated below
30,000 years and therefore are undoubtedly the trails of anatomically modern
humans. The footprints of children seem to be an important component of the trail
record of Palaeolithic caves. For example, Chauvet Cave in southern France revealed
a trail of footprints of a young boy (8 years old and 1.5 m tall) (Harrington 1999).
Niaux is of signicant interest because it includes footprints that may represent
children (Pales 1976). In the Réseau Clastres, three trails of children are recorded
(Lockley et al. 2008). In the Bàsura Cave in Italy, (Chiapella 1952; see Chap. 14)
Late Pleistocene tracks of children were found.
Materials and Methods
After the necessary cleaning process took place, the footprints were copied and
photographed, and a negative cast was created. Positive casts of the footprints were
also created for the necessity of research (mapping). The questions arising are of the
same kind as the ones mentioned earlier. Nevertheless, a more thorough observation
revealed that all four footprints were made by left feet, and this caused one more
10 Middle Palaeolithic Footprints from Greece 173
question to rise: are these solitary footprints or trails of steps crossing each other,
some of which have been lost?
The suggestion that we are before a series of steps made by different individuals is
considered the most possible, since the rst two prints were made by a foot with
some kind of covering, the third has been made by a bare foot, whereas the fourth
one is not clearly distinguished, either because it has been rotated or because other
footprints have been made on it.
Chronology Dating
Ten burnt int specimens unearthed from the lower part of the Middle Palaeolithic
sequence of the cave (layers II2 and II4) were dated by thermoluminescence (TL),
which gave dates ranging between ~110 and 135 ka (Valladas et al. 2007). The
positions of the TL-dated samples are shown in Fig. 10.3. These results are not
consistent with the earlier
14
C dates (Facorellis and Maniatis 1999), as they support a
much later date for these layers (Facorellis et al. 2013). Facorellis and his colleagues
(2013) note that the depositional sequence of Theopetra Cave is complex with
frequently appearing lled channels and underground tunnels as well as labyrinthine
large burrows. It is well established by sediment micromorphological analysis that
Fig. 10.3 Stratigraphic
prole of the trench I10
showing the position of the
samples. (After Facorellis
et al. 2013)
174 N. Kyparissi-Apostolika and S. K. Manolis
the Pleistocene sediments underwent an intense diagenetic mechanical and chemical
alteration related to the caves hydrological conditions.
Comparison between the
14
C dates with the much older TL dates of 11 burnt int
specimens indicates that most of the charcoal samples have been contaminated by a
progressively increasing unidentied amount of exogenous carbon, thus yielding
more recent dates (Facorellis et al. 2013).
Archaeology Lithic Artefacts
The Middle Palaeolithic (II18) is represented by the most clearly stratied lithic
deposits. The principal characteristic of the layer II2 lithic industry is the use of the
Levallois technique for production of a wide range of tool types. The assemblages
from layer II4 bear technological and morphological characteristics often encoun-
tered in the terminal Middle Palaeolithic industries of the Balkans and the Near East.
Their principal characteristic is the use of both Levallois unipolar and prismatic
bipolar core reduction strategies. The tool inventory of Theopetra Cave contains
Middle and, to a lesser extent, Upper Palaeolithic types (Panagopoulou 1999,2000).
Description of the Footprints
A detailed examination disclosed that all four footprints were made by left feet, and
it remains unresolved whether these are solitary footprints or continuous trails of
steps crossing each other where some intermediate steps are missing. There are four
footprints. The rst footprint in the trail (No 1) lacks the posterior half of the foot and
had been made by a covered foot. The second is complete (Fig. 10.4a) and was also
created by a covered foot. Although the footprint is restricted due to the covering
material (footwear), we can clearly observe the arch region and the support lines at
the external region of the foot. The distance between these two left footprints is small
(25 cm), so there is not enough space for the right footprint, which is not preserved.
This led us to suppose that they belong to different individuals. The third is also
complete but has been made by a bare foot (Fig. 10.4b). The big toe, the ball, the arch
and the heel region are evident in the footprint. The last one is disturbed, and it is
very difcult to be analysed.
Two methods of analysis have been used, which led to almost the same results,
stereophotography and photogrammetry. The latter was used by M. Day for the
analysis of various footprints from Laetoli (Tanzania), Niaux Cave (France), Bàsura
Cave (Italy), and Uskmouth (Great Britain). Both methods include the creation of
contour lines (mapping).
However, when we evaluated this method, we estimated it would be quite
difcult to produce an illustration of the contours. This means that the researchers
might miss important parts of the contours (in different altitudes), and so they would
]
ðÞ¼ þ × ðÞ½]
±
10 Middle Palaeolithic Footprints from Greece 175
Fig. 10.4 Contours shaped by 3D laser scanning of the Theopetra footprints; (a) Theopetra No 2;
(b) Theopetra No 3
have to ll them in, without being certain that the additions were correct. It was
therefore decided (see Manolis et al. 2000) to illustrate the contours of the studied
footprints by using a 3D laser scanner, and the results were indeed impressive. It
should also be noted that this method was innovative, the software was still under
development, and this was the rst ofcial application in anthropological material.
Subsequent studies mainly use 3D laser scanner in the study of footprints.
The length and width of the footprints were measured as well, so that they could
be compared to data from contemporary humans. The formulae for the estimation of
height when sex and age are unknown (as in this case) are the following (after Grivas
et al. 2008):
Height cmðÞ¼17:369 þ5:879 ×right foot length in cmðÞ½
Height cm 17:592 5:861 left foot length in cm
Another formula is the estimation of stature by the foot length as a percentage of
body height. The percentage may vary from 14% to 16% according to the population
measured, although the traditional gure quoted is 15% (Topinard 1877). This
formula applies to both sexes and individuals of all ages. However, when calculating
the size of the foot from footprints rather than contours (footwear), there is a
difference of 1%, meaning that the length of the foot is 14% of the height and not
15%, with a variation of 25.4 mm (Robbins 1985,1986).
176 N. Kyparissi-Apostolika and S. K. Manolis
Results
Contour Analysis (3D Laser Scanner)
In order to proceed with our study, we needed to gather data from the contemporary
Greek population. For this rst stage, footprints from three subadults (one male and
two females) were collected. Comparing the contours of the footprints to those of
contemporary juveniles revealed that the former belong to human children.
We can easily see the typical form of the human footprint in the prints of
Theopetra No 2 and 3 in comparison to the prints produced by contemporary
children. This comparison was done in order to prove what seemed to be well
understood: that the ancient footprints belonged to human beings.
The footprint (Fig. 10.5a) was made by a female whose left foot was covered by
three thin stockings, and thus we recreated a print by a covered foot. Notice the
resemblance in the pattern in both prints (Theopetra No. 2 (Fig. 10.5b) and contem-
porary female).
The footprint (Fig. 10.6a) was made by a male child (bare foot). Notice the similar
pattern (ball, arch and the heel region) with the No 3 footprint of Theopetra
(Fig. 10.6b).
All measurements conrm the rst suggestion and leave us with no doubt that
they are the footprints of young children (Table 10.1).
This seems highly unlikely. By comparing the length of the feet, we can make the
following suggestions.
Fig. 10.5 Contour of the footprint created by covered foot of female, 3 years old (a), and the No
2 footprint of Theopetra Cave (b). Note the almost similar pattern in the heel and toe region
10 Middle Palaeolithic Footprints from Greece 177
Fig. 10.6 Contour of the footprint created by the bare foot; (a) modern child; (b) No 3 footprint of
Theopetra Cave. Note the similarity in the pattern of the toes, ball, arch and heel regions
Table 10.1 Measurements of the complete Theopetra footprints and the reference sample of
modern Greek children (in mm)
Variables
Theopetra
No 2
Theopetra
No 3
Modern female
(2 years old)
Modern male
(3 years old)
Modern male
(3 years old)
Foot
length
150.86 138.11 140.2 142.9 165.0
Heel
width
47.28 51.17 33.84 35.19 40.33
Ball
width
54.03 62.82 58.94 55.39 62.0
Sex and Age
Footprint No. 2: If a male (young boy) made the footprint, then it should be a child of
about 3 years. If, on the other hand, a female made it, it should be between 3 and
4 years of age. Note that we should be particularly careful in our nal statements,
because a covered foot created the print, and so measurements may not necessarily
correspond to the actual dimensions. This means that the print is actually bigger than
the foot that created it. Footprint No 3: this bare footprint was made by a child
between 2 and 2.5 years of age, regardless of sex.
The comparison with mean values of Muller et al. (2012) reveals that most
probably the two footprints (No 2 and No 3) were made by children aged 3 and
2 years, respectively. Nevertheless, an older age can't be excluded as a result of
volume reduction due to diagenesis.
Stature
From the footprints available, the ones that provide us with more information are No
2 and No 3.
The stature for the individual 2 falls in the range 97.3106.5 cm (Robbins 1985,
1986). Applying the percentage of 15% (Topinard 1877), a height of 100.6 cm was
estimated. Finally, the application of the formula of Grivas and his colleagues (2008)
gave a stature of 106.0 cm.
The individual that left the bare footprint No 3 probably was a child between
2 and 2.5 years of age, either male or female. The stature falls in the range
89.197.5 cm (Robbins 1985,1986). Applying the percentage of 15% (Topinard
1877), a height of 89.1 cm was estimated. The application of the formula of Grivas
and his colleagues (2008) gave a stature 98.5 cm.
All these calculations are assumptions, because there are several uncertainties
when working with footprints.
Discussion
From the results of our study thus far, we can summarize the following:
178 N. Kyparissi-Apostolika and S. K. Manolis
Footprint size and form: We have limited knowledge about the rate of growth and
development of Neanderthal children. Trinkaus (1983) implies that during the
rst year of their life, Neanderthal children are identical to the children of
anatomically modern humans, but this is based mostly on cranial remains. We
should also point out that the rate of maturation in Neanderthal children has
challenged many scientists, and Dean et al. (1986) proposed that the rate of
development in Neanderthal children may have been faster than that in the
children of early modern humans. Another recent study of Rosas et al. (2017)
notes that Neanderthalsgrowth rate is very similar to that of Homo sapiens,in
general, but differences have been observed in the development of the brain and
spine of these two human groups. These are the main conclusions of a study
which focussed on Neanderthal child approximately 8 years old, who lived in the
cave of El Sidrón (Spain).
Foot function and footwear: Neanderthals used re; they certainly buried their
dead; they seem to have self-medicated with local plants; and they undoubtedly
used foot coverings. It is very crucial to know whether the Neanderthals and early
humans have the same foot function. Recently Bennett and his colleagues (2016)
conclude that foot function has remained almost unchanged, perhaps experienc-
ing evolutionary homeostasis, for the last 3.66 million years. The archaeological
record has limited evidence of footwear. The most ancient evidence appears to be
in Theopetra Cave. The footwear probably had signicant use among Middle
Palaeolithic humans, who may have had various forms of foot covering, to
provide insulation and protection from cold weather and rough substrate
10 Middle Palaeolithic Footprints from Greece 179
(Trinkaus 2005). This phenomenon of footwear use must have been widespread
although the archaeological ndings are very rare.
Age: Children between 2 and 4 years of age produced the footprints, bearing in
mind that these children could belong to early Homo sapiens or Neanderthals.
The analysis of the footprints gave a clear reconstruction of the facts that occurred
at this time in the cave. Firstly, the cave was in use during the Middle Palaeolithic
(Kyparissi-Apostolika 1999; Karkanas 2001; Valladas et al. 2007; Facorellis et al.
2013). There are traces of re remnants on an ashy wet surface, and the footprints
are very near these remnants. This could mean that the children were walking and
playing in the area surrounding the hearth. We suppose that there are several trails
(at least two) which were made by different individuals. The evidence that an
individual who wore some kind of foot covering made the rst two prints
supports this. A bare foot has made the third print, whereas the fourth one is
not clearly distinguished, either because it has been rotated or because other
footprints have been made on it. The study of the footprints reconstructs and
brings the children in our eyes:We can imagine these children playing in the cave
and leaving their traces in the ashy wet surface around the burnt remnants.
Neanderthal children or not? Through dating of specimens from the layer on
which the footprints were found, it became obvious that at the specic time point
early Homo sapiens and Neanderthals coexisted. The former make an appearance
in Europe early, at about ~210 ka in Apidima Cave, Greece (Harvati et al. 2019).
The latter resided in Europe, and their remains are found all over the continent.
Evaluating all the available information, it is difcult to conclude what kind of
individuals left these footprints. How do we know that they are Neanderthal
children? We do not know this because the form and shape of the footprints are
not consistent with the known anatomy of Neanderthals, from various other sites.
Duveau et al. (2019: 19411) note, They are relatively broader, especially in the
midfoot, than the footprints made by Homo sapiens, which corresponds to a more
robust foot and a less pronounced arch.
The fact that the footprints of Theopetra have been made by children aged
24 years would lead one to hypothesize that probably the foot at this age is not
fully developed. On the other hand, the cultural ndings (lithic artefacts) seem to be
Mousterian (the typical technological expression of Neanderthals) and have led to
the conclusion that the footprints were made by Neanderthal children.
Conclusion
Both the palaeoanthropological and archaeological records suggest that foot cover-
ing was present during the Middle Palaeolithic. The only evidence that helps us in
this case for establishing chronology is the lithic material found in this layer. If
eventually the assessment that the tools are Mousterian (Panagopoulou 1999,2000)
is conrmed, we will positively assume that they were Neanderthal children. But
even if they turn out to be the children of early Homo sapiens, the signicance of the
ndings is still great, since that would prove beyond reasonable doubt the presence
of anatomically modern humans in Europe at approximately ~135 ka much earlier of
what was thought until now.
180 N. Kyparissi-Apostolika and S. K. Manolis
Acknowledgements We thank Dr. Tsitsiloni, Mrs. P. Aleri and Mrs. E. Malliou (for the
permission to cast their childrens fee). In addition, many thanks to Mr. P. Polydoropoulos for
casting the Theopetra footprints and Dr. I. Petroutsa and Mrs. H. Belte for their assistance. Finally,
we thank Dr. P. Karkanas for his valuable information on how the footprints were preserved and
Dr. C. Eliopoulos for the editing of the manuscript.
References
Ashton, N., Lewis, S. G., De Groote, I., Duffy, S. M., Bates, M., et al. (2014). Hominin footprints
from early pleistocene deposits at Happisburgh, UK. PLoS ONE, 9(2), e88329-1-13.
Avanzini, M., De Angelis, M., Mietto, P., Panarello, A., & Rolandi, G. (2004). Pleistocene human
footprints preserved on a zeolite-rich pyroclastic ow (Roccamanna, Italy). In International
Union of Geological Sciences (Ed.), 32nd international geological congress (p. 598). Florence:
Università di Firenze.
Avanzini, M., Mietto, P., Panarello, A., De Angelis, M., & Rolandi, G. (2008). The devils trails:
Middle Pleistocene human footprints preserved in a volcanoclastic deposit of Southern Italy.
Ichnos, 15, 179189.
Benazzi, S., Douka, K., Fornai, C., Bauer, C. C., Kullmer, O., Svoboda, J., Pap, I., Mallegni, F.,
Bayle, P., Coquerelle, M., Condemi, S., Ronchitelli, A., Harvati, K., & Weber, G. W. (2011).
Early dispersal of modern humans in Europe and implications for Neanderthal behaviour.
Nature, 479, 525529.
Bennett, M. R., Reynolds, S. C., Morse, S. A., & Budka, M. (2016). Footprints and human
evolution: Homeostasis in foot function? Palaeogeography, Palaeoclimatology,
Palaeoecology, 461, 214223.
Chiapella, V. G. (1952). Orsi e nomini preistorici nella grotta della Strega, (Genova). Rivista del
Comune, A29,2229.
Day, M. H. (1991). Bipedalism and prehistoric footprints. In Y. Coppens & B. Sénut (Eds.), Origine
(s) de la Bipedie chez les Hominides (pp. 199213). Paris: Cahiers de Paleoanthropologie,
Editions du CNRS.
De Lumley, H. (1966). Les fouilles de Terra Amata a Nice. Premiers rèsultats. Bulletin of the
Museum of Anthropology and Prehistory of Monaco, 13,2951.
De Lumley, H. (1967). Découverte dhabitats de lacheuléen ancien dans les dépôts mindéliens sur
le site de Terra Amata, Nice, Alpes-Maritimes. Compte-Rendus de LAcadémie des Sciences,
264, 801804.
Dean, M. C., Stringer, C. B., & Bromage, T. G. (1986). Age at death of the Neanderthal child from
Devils tower, Gibraltar and the implications for studies of general growth and development in
Neanderthals. American Journal of Physical Anthropology, 70, 301309.
Duveau, J., Berillon, G., Verna, C., Laisné, G., & Cliquet, D. (2019). Neanderthal childrens
footprints offer rare snapshot of Stone Age family life. PNAS, 116(39), 1940919414.
Facorellis, Y., & Maniatis, Y. (1999). Possibilities and accuracy of radiocarbon dating in the
Palaeolithic period. British School at Athens Studies, 3, 179189.
Facorellis, Y., Kyparissi, N., & Maniatis, Y. (2001). The Cave of Theopetra, Kalambaka: Radio-
carbon evidence for 50,000 years of human presence. Radiocarbon, 43(2B), 10291048.
10 Middle Palaeolithic Footprints from Greece 181
Facorellis, Y., Karkanas, P., Higham, T., Brock, F., Ntinou, M., & Kyparissi-Apostolika, N. (2013).
Interpreting radiocarbon dates from the Paleolithic layers of Theopetra Cave in Thessaly,
Greece. Radiocarbon, 55(23), 14321442.
Grivas, T., Mihas, C., Arapaki, A., & Vasiliadis, E. (2008). Correlation of foot length with height
and weight in school age children. Journal of Forensic and Legal Medicine, 15,8995.
Harrington, S. P. M. (1999). Human footprints at Chauvet Cave. Archaeology, 52(5) www.archae
ology.org/9909/newsbriefs/chauvet.html. Accessed 14 Jan 2020.
Harvati, K., Röding, C., Bosman, A. M., Karakostis, F. A., Grün, R., Stringer, C., Karkanas, P.,
Thompson, N. C., Koutoulidis, V., Moulopoulos, L. A., Gorgoulis, V. G., & Kouloukoussa,
M. (2019). Apidima Cave fossils provide earliest evidence of Homo sapiens in Eurasia. Nature,
571, 500504.
Karkanas, P. (2001). Site formation processes in Theopetra Cave: A record of climatic change
during the late Pleistocene and early Holocene in Thessaly, Greece. Geoarchaeology, 16,
373399.
Kyparissi-Apostolika, N. (1998). The signicance of Theopetra cave for Greek prehistory. In
M. Otte (Ed.), Préhistoire dAnatolie Genèse de deux mondes, Actes du colloque international,
Liège, 28 avril-3 mai 1997. ERAUL 85 (pp. 241252).
Kyparissi-Apostolika, N. (1999). The palaeolithic deposit of Theopetra cave, Thessaly (Greece). In
G. E. Bailey, E. Adam, E. Panagopoulou, C. Perles, & K. Zachos (Eds.), The palaeolithic
archaeology of Greece and adjacent areas, proceedings of the ICOPAG conference, studies 3
(pp. 232239). Athens: British School at Athens.
Lockley, M., Roberts, G., & Kim, J. Y. (2008). In the footprints of our ancestors: An overview of
the hominid track record. Ichnos, 15, 106125.
Manolis, S. K., Aiello, L. C., Henessy, R., & Kyparissi-Apostolika, N. (2000). The middle
Paleolithic footprints of Theopetra cave (Thessaly, Greece). In N. Kyparissi-Apostolika (Ed.),
Proceedings of the international conference, Theopetra Cave: 12 years of excavation and
research, Trikala, 67 November 1998 (pp. 8793). Athens: Greek Ministry of Culture and
Institute for Aegean Prehistory.
Muller, S., Carlsohn, A., Muller, J., Baur, H., & Mayer, F. (2012). Static and dynamic foot
characteristics in children aged 113 years: A cross-sectional study. Gait & Posture, 35,
389394.
Onac, B. P., Viehmann, I., Lundberg, J., Lauritzen, S. E., Stringer, C., & Popita, V. (2005). UTh
ages constraining the Neanderthal footprint at Vârtop Cave, Romania. Quaternary Science
Reviews, 24, 11511157.
Pales, L. (1976). Les Empreintes de Pieds humains dans les Cavernes (Memoir 36, Archives de
lInstitut de Paleontologie Humaine (Fondation Albert 1er, Prince de Monaco)). Paris: Masson.
Panagopoulou, E. (1999). The Theopetra middle palaeolithic assemblages: Their relevance to the
middle palaeolithic of Greece and adjacent areas. In G. N. Bailey, E. Adam, A. Panagopoulou,
C. Perles, & K. Zachos (Eds.), The Palaeolithic archaeology of Greece and adjacent areas.
Proceedings of the ICOPAG conference, studies 3 (pp. 252265). British School at Athens:
Athens.
Panagopoulou, E. (2000). The middle palaeolithic lithic assemblages of Theopetra cave: Techno-
logical evolution in the upper pleistocene. In N. Kyparissi-Apostolika (Ed.), Proceedings of the
international conference, Theopetra Cave: 12 years of excavation and research, Trikala,
67 November 1998 (pp. 152161). Athens: Greek Ministry of Culture and Institute for Aegean
Prehistory.
Robbins, L. M. (1985). Footprints: Collection, analysis and interpretation (p. 225). Springeld:
Charles C. Thomas.
Robbins, L. M. (1986). Estimating height and weight from size of footprints. Journal of Forensic
Sciences, 31(1), 143152.
Rosas, A., Ríos, L., Estalrrich, A., Liversidge, H., García-Tabernero, A., Huguet, R., Cardoso, H.,
Bastir, M., Lalueza-Fox, C., de la Rasilla, M., & Dean, C. (2017). The growth pattern of
Neandertals, reconstructed from a juvenile skeleton from El Sidrón (Spain). Science, 357,
12821287.
Topinard, P. (1877). LAnthropologie (Vol. 1, 2nd ed.). Reinwald: Paris.
182 N. Kyparissi-Apostolika and S. K. Manolis
Trinkaus, E. (1983). Functional aspects of Neandertal pedal remains. Foot & Ankle, 3(6), 377390.
Trinkaus, E. (2005). Anatomical evidence for the antiquity of human footwear use. Journal of
Archaeological Science, 32, 15151526.
Valladas, H., Mercier, N., Froget, L., Joron, J. L., Reyss, J. L., Karkanas, P., Panagopoulou, E., &
Kyparissi-Apostolika, N. (2007). TL age-estimates for the Middle Palaeolithic layers at
Theopetra cave (Greece). Quaternary Geochronology, 2, 303308.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Chapter 11
On the Tracks of Neandertals: The
Ichnological Assemblage from Le Rozel
(Normandy, France)
Jérémy Duveau, Gilles Berillon, and Christine Verna
Abstract Hominin tracks represent a unique window into moments in the life of
extinct individuals. They can provide biological and locomotor data that are not
accessible from skeletal remains. However, these tracks are relatively scarce in the
fossil record, particularly those attributed to Neandertals. They are also most often
devoid of associated archaeological material, which limits their interpretation. The
Palaeolithic site of Le Rozel (Normandy, France) located in a dune complex formed
during the Upper Pleistocene has yielded between 2012 and 2017 several hundred
tracks (257 hominin footprints, 8 handprints as well as 6 animal tracks). This
ichnological assemblage is distributed within ve stratigraphic subunits dated to
80,000 years. These subunits are rich in archaeological material that attests to brief
occupations by Neandertal groups and provides information about the activities that
they carried out. The ichnological assemblage discovered at Le Rozel is the largest
attributed to Neandertals to date and more generally the most important for hominin
taxa other than Homo sapiens. The particularly large number of footprints can
provide major information for our understanding of the Palaeolithic occupations at
Le Rozel and for our knowledge of the composition of Neandertal groups.
Keywords Group composition · Morphometry · Footprint · Neandertals · Le Rozel
Introduction
Tracks, and especially footprints, are unique vestiges that provide direct information
aracteristics (
on the locomotor and biological ch e.g. stature, body mass, age) of
hominin groups (e.g. Bennett et al. 2009; Crompton et al. 2011; Bennett and Morse
2014). Such information can be obtained from trackways (e.g. Leakey and Hay
1979; Masao et al. 2016; Roach et al. 2016) or from isolated footprints by using
J. Duveau (*) · G. Berillon · C. Verna
UMR 7194 HNHP, Centre National de la Recherche Scientique, MusÕum national dHistoire
naturelle, Université Perpignan Via Domitia, Paris, France
e-mail: jeremy.duveau@edu.mnhn.fr
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_11
183
morphometric methods (Dingwall et al. 2013; Ashton et al. 2014; Citton et al. 2017)
or expert tracker readings (Pastoors et al. 2015,2017). Ichnological assemblages
require a quick sedimentary burial to be preserved in an open-air context; this differs
from the cave context, where they are usually found at the surface of the soil (see
Chap. 4). As it, they represent an original snapshot on the composition of groups and
their behaviours during their lives (e.g. Mastrolorenzo et al. 2006; Hasiotis et al.
2007; Schmincke et al. 2010; Falkingham 2014). They differ in this respect from
skeletal or lithic material whose accumulations may have occurred during various
and repeated occupations over long periods (Farizy 1994; Pettitt 1997). However,
the study of tracks is usually a challenging task. Indeed, if their morphology reects
the biological and locomotor characteristics of trackmakers, they are also affected by
the nature of substrate and by taphonomic modications (e.g. Allen 1997; Bennett
and Morse 2014; see Chap. 2). In addition, despite several signicant discoveries in
recent years (e.g. Altamura et al. 2018; Bustos et al. 2018; McLaren et al. 2018; see
Chap. 5), the number of sites that yielded hominin tracks is relatively low compared
to sites with archaeological and palaeoanthropological material (e.g. Kim et al. 2008;
Lockley et al. 2008,2016; Bennett and Morse 2014). This rarity is even more
important for the footprints attributed to Neandertals since only nine footprints
184 J. Duveau et al.
found at four sites attributed to this taxon were reported to date (Fig. 11.1).
In this context, here we present the largest ichnological assemblage attributed to
Neandertal discovered at the archaeological site from Le Rozel (Manche, France).
We present rst a synthesis of the previously known footprints attributed to Nean-
dertals. Then the archaeological site of Le Rozel will be presented before describing
Fig. 11.1 Geographical distribution of the footprints attributed to Neandertals
the ichnological sample discovered there between 2012 and 2017. Finally, the
importance of this assemblage in relation to other sites that yielded hominin foot-
prints, and in particular those attributed to Neandertals, will be discussed before
concluding on the potential of these footprints to yield direct information on the
11 On the Tracks of Neandertals: The Ichnological Assemblage from Le Rozel... 185
trackmaker groups that lived at Le Rozel 80,000 years ago.
The Neandertal Footprint Record
The oldest track attributed to Neandertals is also the rst that has been described: it is
a single footprint discovered in 1976 in a silty ground at the Middle Pleistocene site
of Biache-Saint-Vaast (France) (Tuffreau 1978,1988). It is associated with two
fragmentary human skulls that bear Neandertal features as well as with archaeolog-
ical material including lithic industry and 236,000-year-old faunal remains (Tuffreau
1988; Rougier 2003; Guipert et al. 2011; Bahain et al. 2015). The attribution of the
track to a Neandertal individual is based on the cranial remains and on the associated
archaeological material. This footprint is poorly preserved and was probably dam-
aged by bovid trampling making its identication as a hominin footprint and its
analysis difcult (Tuffreau 1988).
Four footprints were discovered in 1996 in the Greek cave of Theopetra. They
were made in a clay substrate dated by thermoluminescence to 130,000 years
(Manolis et al. 2000; Valladas et al. 2007; see Chap. 10). They are associated with
a Mousterian industry that allows to attribute them to Neandertals (Manolis et al.
2000; Valladas et al. 2007). The four footprints were probably made by different
individuals with their left feet. The second and the third footprints are relatively
complete. They are 14 and 15 cm long and were made by young children whose ages
and statures are estimated to 2 and 4 years and to 86 and 100 cm (Manolis et al.
2000). Furthermore, Manolis et al. (2000) suggest that the third footprint was made
by a shod individual, which would represent the oldest occurrence of a shoe among
hominins. Casts of the footprints were realized, and the two most complete were 3D
digitized (Manolis et al. 2000).
Three footprints made in calcareous mud dated by U-Th between 97,000 and
62,000 years were discovered in the Romanian Vârtop Cave (Onac et al. 2005; see
Chap. 12). No archaeological or palaeoanthropological material was associated with
these tracks. The taxonomic attribution to Neandertals is based only on the chrono-
logical age, Neandertals being the only taxon known in Europe for this time period.
The three footprints were made by a single individual (Onac et al. 2005; Harvati and
Roksandic 2016). Two of them are partial, consisting only of either heel or forefoot
impressions. The third footprint is longitudinally complete; it is 22 cm long and was
made by an individual whose height was estimated to 146 cm (Viehmann 1987). It is
characterized by a space described as important (1.6 cm) between the hallux and the
second toe impressions (Onac et al. 2005). Its morphology would reect the robust
Neandertal anatomy (Onac et al. 2005).
More recently, a potential human footprint was discovered in the dune complex of
Catalan Bay at Gibraltar. OSL dating of the aeolian unit where the footprint was
made provided an age of 28,000 years (Muñiz et al. 2019). This footprint is
described as poorly preserved. It is 17 cm long and was made by an individual
whose height is estimated between 106 and 126 cm and who was descending a slope
(Muñiz et al. 2019). No archaeological or palaeoanthropological remains are asso-
ciated with this footprint. Moreover, its morphology does not allow to discard Homo
sapiens as the possible trackmaker (Muñiz et al. 2019). Therefore, the taxonomic
attribution to a Neandertal individual is only based on the discovery a few kilometres
away of archaeological material that would indicate that Neandertal groups may
have lived in the region until 28,000 years BP (Finlayson et al. 2006). However, the
dating of this material is questioned not only as regards stratigraphic consistency
(Delson and Harvati 2006) but also for methodological aspects (Wood et al. 2013).
Therefore, the lack of consensus on these dates combined with the fact that the
186 J. Duveau et al.
footprint would correspond to the last Neandertal occurrence raises questions about
the validity of the taxonomic attribution to a Neandertal individual.
In this synthesis on footprints attributed to Neandertals, it is necessary to mention
those discovered in the Romanian cave of Ciur Izbuc (Webb et al. 2014; see
Chap. 12). The research undertaken at this cave yielded 400 human footprints,
dated between 36,500 and 29,000 years calBP, before three quarters of them were
destroyed (Webb et al. 2014). The absence of archaeological or palaeo-
anthropological material associated with the footprints makes their taxonomic attri-
bution complex. Indeed, the lowest limit of the chronological interval is close to the
last occurrence of Neandertals reported in central and Eastern Europe (Pinhasi et al.
2011; Devièse et al. 2017). However, skeletal remains provided evidence of the
occurrence of Homo sapiens in Romania around the period when the footprints were
made (Trinkaus et al. 2003;Socaru et al. 2007; Higham et al. 2011). It is thus more
likely that these footprints were made by Homo sapiens (Webb et al. 2014).
Lastly, the footprints discovered in the Italian site of Bàsura Cave were for a long
time attributed to Neandertals (Pales 1954,1960). For this attribution, L. Pales used
the presence of a Mousterian industry in a nearby cave and remains of cave bears that
he considered as contemporary to Neandertals. However, subsequent radiocarbon
dating on charcoals discovered in the same layer as the footprints invalidated their
taxonomic attribution to Neandertals, showing instead that they were made by Homo
sapiens (Molleson et al. 1972; De Lumley and Vicino 1984).
The Archaeological Site from Le Rozel
Located on the western coast of the Cotentin (Manche, Normandy) (Fig. 11.1), Le
Rozel (4928'20.92'' N, 150'25.58'' W) is part of a dune formation in a creek
opened in a schist cliff. This dune complex is composed of soft aeolian sand and
was formed during the end of the Eemian and the beginning of the Last Glacial
Period, between 115,000 and 70,000 years ago (Van Vliet-Lanoë et al. 2006). The
site was discovered in the 1960s by Yves Roupin following coastal erosion that
uncovered several faunal bones at the base of the dune. These initial discoveries led
to a survey in 1967 and to the rst excavations in 1969 directed by Frédéric Scuvée
(Scuvée and Verague 1984). The monitoring of the site since the 1980s has revealed
11 On the Tracks of Neandertals: The Ichnological Assemblage from Le Rozel... 187
signicant damages caused by erosion and led to annual excavations under the
direction of D. Cliquet since 2012.
Le Rozel shows a long stratigraphic sequence (Fig. 11.2)dominated by detrital
elements brought by wind dynamics (Van Vliet-Lanoë et al. 2006). This sequence is
delimited at its summit by a 6- to 8-m-thick head above a palaeodune massif. The
archaeological layers discovered since 2012 are located within ve stratigraphic
subunits of this palaeodune (D3b-1 to D3b-5) composed of ne to medium sand
(Cliquet et al. 2018a,b). The OSL dating carried out within the stratigraphic
sequence places these subunits around 80,000 years (Mercier et al. 2019). Further-
more, geochronological and sedimentary analyses have shown that the stratigraphic
subunits were formed and covered quickly (Mercier et al. 2019) which means that
each subunit represents a relatively short and likely single occupation phase. The
rst three subunits (D3b-1 to D3b-3) are composed of subhorizontal organic soils,
brown to black in colour, and consist of degraded dune sand where lithic industries,
charcoals, faunal remains and tracks were discovered (Fig. 11.3). The Palaeolithic
Fig. 11.2 Cross section of the dune complex from Le Rozel and locations of the Palaeolithic
occupations. (Modied from Cliquet et al. 2018b)
188 J. Duveau et al.
activities in these soils seem to be structured around hearths and, for the D3b-2 and
D3b-3 subunits, knapping spots (Cliquet et al. 2018a,b). The D3b-4 and D3b-5
subunits, whose excavations are still in progress, are affected by numerous
intertwined mudows that are intersected by small schist plates. These subunits
yielded lithic industries, knapping spots, hearths, faunal remains and most of the
tracks (Cliquet et al. 2018a,b). Below these stratigraphic subunits are the occupation
layers identied and studied by Scuvée during the 1960s (Scuvée and Verague
1984). Two layers (Scuvée E2 and Scuvée E3) were located at the base of the
dune and included faunal remains and lithic industry, while a third layer (Scuvée F2)
was located inside a rock shelter (TR 67) where hearths, faunal remains and lithic
Fig. 11.3 The archaeological site from Le Rozel; (a) view of the site; (b) Levallois akes; (c)
blades; (d) knapping spot; (e) hearth; (f) butchery area. (Photos D. Cliquet)
11 On the Tracks of Neandertals: The Ichnological Assemblage from Le Rozel... 189
artefacts were discovered (Scuvée and Verague 1984; Van Vliet-Lanoë et al. 2006;
Cliquet et al. 2018a,b).
The analyses carried out on the archaeological material discovered at Le Rozel
show that two techno-cultural worlds were operated by the human groups
80,000 years ago (Cliquet et al. 2018a,b). This dichotomy observed between the
two sets of occupations is particularly visible with the lithic industries. Indeed,
although the raw materials and their relative frequency are similar between the
upper and lower subunits (a majority of local int and to a lesser extent quartz;
anecdotal use of sandstone and mylonite), the characteristics of the industry differ.
The industries discovered in subunits D3b-1 to D3b-3 mainly represent direct
debitage akes and Levallois akes. The D3b-4 and D3b-5 ones correspond to a
higher proportion of lamellar and laminar productions. While some blades come
from the production of direct debitage akes or Levallois akes, a lot of them have
been obtained by semi-rotating or rotating debitages (Cliquet et al. 2018a,b).
The three more recent upper subunits (D3b-1 to D3b-3) provide evidence of
butchery activities (Fig. 11.3), whereas site function for the lower subunits (D3b-4,
D3b-5, Scuvée E2-E3, Scuvée F2) is not yet established (Cliquet et al. 2018a,b).
Within the D3b-1 to D3b-3 subunits, the fauna consumed is largely dominated by
red deer, horse and aurochs, both in terms of number of remains and minimum
number of individuals (Sévêque 2017). The bones of these three species bear the
characteristic stigmata of skinning, dismantling and the recovery of meat. The study
of the slaughter periods of this fauna enabled to estimate that Palaeolithic occupa-
tions took place during bad weather seasons, between autumn and spring (Sévêque
2017; Cliquet et al. 2018b). Other bones belong to straight-tusked elephant, grass-
land rhinoceros, roe deer and rabbit whose nutritional usefulness is not conrmed.
Anthracological analyses show that the hearths were mainly composed of Scots pine
and yews, which could reect a vegetal selection. Anthracological and
zooarchaeological material provide a representation of the environments during the
Palaeolithic occupations of the site (Stoetzel et al. 2016; Sévêque 2017; Cliquet et al.
2018a,b): they are characteristic of a temperate climate and open landscapes,
including humid temperate semi-wooded meadows.
The lower subunits (D3b-4 and D3b-5) are less informative than the rst three;
only the large fauna, which is weakly conserved, provides results. In these layers, red
deer is once again the most frequent, with horse and aurochs (Sévêque 2017; Cliquet
et al. 2018a,b).
In the absence of human osteological remains, the Palaeolithic occupations at Le
Rozel are attributed to Neandertals by considering the chronostratigraphic context
and the characteristics of the archaeological material. On the one hand, the D3b-1 to
D3b-5 subunits are dated to 80,000 years when Neandertals were the only taxon
known in Europe (Benazzi et al. 2011;Nigst et al. 2014; Hublin 2015
with Neandertal remains (Cliquet et al. 2018b). The cultural dichotomy between the
upper and the lower subunits suggests the presence of different groups.
). On the other
hand, the technological features of the archaeological material, especially that of the
upper subunits, have already been observed on other Mousterian sites associated
190 J. Duveau et al.
Material and Methods
The analysis of the ichnological set discovered between 2012 and 2017 at Le Rozel
led to the identication of 271 tracks including 257 hominin footprints, 8 hominin
handprints (Fig. 11.4) as well as 6 animal tracks (Duveau et al. 2019).
These tracks were identied by morphological criteria; they had to reect the
anatomy and the locomotor behaviour of the trackmakers. More particularly, human
footprints reect a rounded heel, a narrow midfoot, relatively short toes including a
robust and adducted hallux (e.g. Aiello and Dean 1990; Klenerman and Wood 2006;
Morse et al. 2010; Bennett and Morse 2014). The heel and forefoot impressions are
deeper than that of the midfoot (Crompton and Pataky 2009; Morse et al. 2010;
Bennett and Morse 2014). Moreover, the identication of the human footprints was
reinforced by using the morphometric test developed by Morse et al. in 2010
(Duveau et al. 2019). Human handprints are recognizable by the impressions of a
rounded palm, relatively wider than the heel of the foot, and the ngers are relatively
long except the thumb which is smaller. This thumb has an abduction capacity unlike
the human hallux (e.g. Aiello and Dean 1990; Jones and Lederman 2006). At last,
the animal tracks were identied and taxonomically attributed thanks to identica-
tion criteria from the literature (e.g. Bang et al. 2001; Murie and Elbroch 2005).
Each track was photographed and described in situ. Casts were made of 64 tracks
between 2013 and 2016. Seventy original tracks were directly extracted after a
chemical consolidation of the substrate in 2017. The casts and extracted footprints
are curated in the premises of the Direction Régionale des Affaires Culturelles
(DRAC, Caen, France). 180 tracks including 170 footprints were digitized in 3D.
Fig. 11.4 Hominin footprints and handprints discovered at Le Rozel. (Photos D. Cliquet)
(scale bar : 10 and 40 cm)
Seventy-seven footprints were digitized by using photogrammetry with the Agisoft
Protocan software (v.1.4.0) and a Canon EOS 1300D camera. 137 footprints were
3D modelled by using a Noomeo OptiNum surface scan. The use of these different
acquisition techniques required that we run morphometric comparisons between
them prior to analysis. These comparisons did not detect any differences between
the types of acquisition (Duveau et al. 2019).
11 On the Tracks of Neandertals: The Ichnological Assemblage from Le Rozel... 191
Besides,each track was measured in situ. These measurements were controlled
and specied on the tracks digitized in 3D by using Geomagic Studio 2013. The
length was measured along the longitudinal axis. For footprints, this axis is from the
most proximal point of the heel to the distal end of the second toe. Therefore, length
measurement requires that the footprint is longitudinally complete and that the toe
impressions can be differentiated from the rest of the print. In addition, we measured
the maximum width of the forefoot, along a mediolateral axis perpendicular to the
longitudinal axis. The lengths and widths of the footprints from Le Rozel were
compared to those of the other footprints attributed to Neandertals using published
data: two of the four footprints from Theopetra Cave (Manolis et al. 2000), the most
complete footprint from Vârtop Cave (Onac et al. 2005) and the potential footprint
from Catalan Bay (Muñiz et al. 2019).
Results
Preservation and Distribution of the Tracks
The 271 discovered tracks have been preserved thanks to a rapid sedimentary cover
by aeolian sand. Indeed, experimental observations carried out in situ have shown
that without this protection, the tracks could have been damaged, if not entirely
destroyed, in a few tens of minutes. Due to the erosive action of the wind on the
tracks, as well as other taphonomic agents, this ichnological assemblage probably
represents only a sample of the initial assemblage left by the trackmakers
~80,000 years ago.
The tracks come from the subunits D3b-1 to D3b-5 and were discovered in the
same layers as the archaeological material. Nearly 80% of the reported tracks come
from the D3b-4 stratigraphic subunit, which extends over more than 90 m
2
; 11% of
the tracks come from the D3b-5 subunit; the rest of the tracks are similarly distrib-
uted among the three other subunits. Among the 271 tracks, 198 were made in sandy
mud and 73 in dune sand. The tracks made in dune sand, which mainly come from
the D3b-1 to D3b-3 subunits, are less well preserved (i.e. they reect less anatomical
details and in particular less clear toe impressions) than those made in sandy mud
that come from the D3b-4 and D3b-5 subunits. This differential conservation partly
explains the differences in distribution between the subunits. The depth of the tracks
is highly variable, from a few millimetres to 5 cm, and may suggest varying moisture
conditions when they were made.
192 J. Duveau et al.
Human Footprints
Description: Among the 257 footprints, 5 trackways composed of 2 to 3 footprints
were reported (Fig. 11.4), the rest of the footprint set consisting of isolated tracks.
They include 112 left prints, 115 right prints and 30 impressions of indeterminate
laterality. Footprint morphology is variable which is common for footprints made in
soft substrate (e.g. Allen 1997; Morse et al. 2013; Bustos et al. 2018), such as dune
sand or sandy mud. The quality of the prints is variable, and some are partial. Ten
prints correspond only to the heel impressions, and three reect only the forefoot.
Eighty-eight footprints are longitudinally complete since they show proximally the
impressions of rounded heels and distally clear impressions of the tip of the toes. Of
these longitudinally complete footprints, not all the toes are systematically printed.
The hallux impression and to a lesser extent that of the second toe are the most
common and the deepest toe impressions. With one exception, the hallux impression
is always visible when the impressions of toes can be distinguished from the rest of
the footprint. The remaining 156 footprints reect a relatively complete foot outline
but do not provide evidence, such as variation in depth, allowing to distinguish the
toe impressions. It is therefore difcult to attest that they are longitudinally complete.
The best-preserved footprints reect morphological features close to those of
humans including a fully adducted hallux and a midfoot mediolaterally narrow.
Moreover, the heel and forefoot impressions are the deepest areas of the footprints;
the forefoot is on average deeper than the heel. The midfoot impression is shallow
and has a slight outline. This depth distribution and the narrowing of the midfoot
impression are consistent with an architecture of the foot in vault. These architectural
characteristics are less pronounced for the smallest footprints, which suggest a atter
foot for the youngest individuals. They are also less marked compared to footprints
made by Homo sapiens (Duveau et al. 2019), which is consistent with our knowl-
edge of the anatomy of the Neandertal foot, which was more robust and had a less
pronounced plantar arch than the Homo sapiens foot (Trinkaus et al. 1991; Berillon
2000).
Comparative morphometry: The 3D modelling of 169 footprints allows accurate
morphometric comparisons according to the subunits (Fig. 11.5a). These compari-
sons were carried out on footprints sufciently complete that were made on hori-
zontal layers and that do not show any evidence of sliding. The footprints from the
D3b-4 stratigraphic subunit, the densest in tracks, have lengths ranging from 11.4 to
28.4 cm (mean, 19.2 cm) and widths from 4.5 to 12.8 cm (mean, 8.4 cm). The
exploitable footprints from the other stratigraphic subunits fall within these ranges
(Fig. 11.5). The footprints from the D3b-1 subunit are shorter (12.318.4 cm) and
narrower (mean, 5.18.4 cm) than the average of those from the D3b-4. The
footprints from the D3b-2 and D3b-3 subunits are on average longer (respectively,
21.4 cm and 22.0 cm) but have close average widths (8.1 cm and 8.7 cm). Finally,
the footprints from the D3b-5 subunit are biometrically close to those from the
D3b-4 for both length (mean, 19.7 cm) and width (mean, 7.9 cm). The lengths and
widths of the other footprints attributed to Neandertals fall within the ranges of those
11 On the Tracks of Neandertals: The Ichnological Assemblage from Le Rozel... 193
from Le Rozel (Fig. 11.5b). The two footprints from Theopetra Cave (14 and 15 cm
long, 5 and 6 cm wide) and the footprint from Catalan Bay (17 cm long and 7 cm
wide) are relatively smaller than the averages of the Le Rozel footprints (19,2 cm
long and 8,4 cm wide). On the other hand, the footprint from Vârtop Cave is
relatively longer (22 cm) and wider (11 cm).
Fig. 11.5 Dimensions of the Le Rozel footprints digitized in 3D; (a) depending on their locations
in the stratigraphic subunits; (b) compared to the other footprints attributed to Neandertals
Human Handprints
The eight handprints all reect a right laterality. As with footprints, their morphology
is variable. Six handprints are longitudinally complete with lengths ranging from
11.4 to 16.1 cm. The two other handprints show ngerprints but not clearly the base
of the palm. The handprints are characterized by a broad palm, deep and long
ngerprints (relatively longer than toe impressions) and a short thumb with a
capacity for abduction.
194 J. Duveau et al.
Animal Tracks
Six animal tracks were discovered in the peripheral areas of the D3b-4 subunit. Their
low level of conservation complicated their precise taxonomic attribution. Five of
them are attributed to Carnivora (Felidae, Canidae and Mustelidae) and the last one
to a Ruminantia (probably a Cervidae).
Discussion
Since 2012, the eld missions yielded a large ichnological assemblage that makes Le
Rozel a major track site. First of all, the 257 footprints represent to date the largest
footprint sample attributed to a hominin taxon other than Homo sapiens. In partic-
ular, they form more than 95% of all the footprints attributed to Neandertals since
only nine footprints had so far been attributed to this taxon (Fig. 11.1). Moreover,
even for footprints attributed to Homo sapiens, such a large number is exceptional;
the sites from the Willandra Lakes (e.g. Webb et al. 2006; Webb 2007) and the
Hawaii Volcanoes National Park (Moniz Nakamura 2009) are among the few sites
that yielded more footprints than Le Rozel. Also notable are the eight handprints
recorded at Le Rozel that are to date the only Neandertal handprints available with
the hand discovered at Maltravieso (Hoffmann et al. 2018). Besides the well-known
positive and negative painted hands in rock art (e.g. Bahn 1998; Guthrie 2005), only
a few handprints are known for Pleistocene hominins (Zhang and Li 2002; Mietto
et al. 2003; Ledoux et al. 2017; Panarello et al. 2018). In addition, animal tracks
provide information on fauna that lived near the site during the Palaeolithic occu-
pations. For example, they attest to the presence of several carnivores of which no
osteological remains had been found on the site (Cliquet et al. 2018a,b). Le Rozel
tracks also represent an important discovery because of their association with
archaeological material that attests to the occupations of the site by Neandertals.
Such occupation contexts are rare among the other hominin footprint sites (Altamura
et al. 2018); the majority of them only reect passage areas (Masao et al. 2016;
Roach et al. 2016; Bustos et al. 2018).
Except for the dimensions (Fig. 11.5b), it is difcult to morphologically compare
the Le Rozel footprints with other footprints attributed to Neandertals because of
their rarity as well as the differences in conservation and deposition conditions. Few
anatomical details are reected by these other footprints. Only a gap described as
important between the hallux and the second toe has been reported for the most
complete footprint discovered at Vârtop Cave (Onac et al. 2005). Such a space is not
observed on the Le Rozel footprints, but this may be related to the nature of the
substrate.
In that vein, some morphological features of the Le Rozel footprints, such as the
lack of clear toe impressions on relatively complete footprints, raise a question: the
possibility of shod feet. Such a feature could have a signicant impact on our
e
knowledge of Neandertal culture. No direct remains of shoes are known for Nean-
dertals, and the earliest occurrences were discovered in Holocene sites (e.g. Kuttruff
et al. 1998; Pinhasi et al. 2010). However, anatomical studies on the robustness of
phalanges suggested a possible wearing of shoes as early as 30,000 years ago
(Trinkaus and Shang 2008; see Trinkaus et al. Chap. 7). In addition, some footprints,
such as one of those discovered at Theopetra Cave (Manolis et al. 2000;se
Kyparissi-Apostolika and Manolis Chap. 10), have been described as prints of
shod feet. Nevertheless, the associations between footprints and footwear are not
certain, being generally based on qualitative criteria or on outliers in footprint
dimensions (Bennett et al. 2010). Experimental studies on the same substrate
conditions as at Le Rozel, and investigating morphometric differences between
barefoot footprints and footwear (including shoes of varied rigidity), may provide
signicant information on this issue in the future.
11 On the Tracks of Neandertals: The Ichnological Assemblage from Le Rozel... 195
Finally, we have shown that the 20122017 assemblage described here could
provide direct information on the size and composition of the trackmaking groups;
the assemblage from D3b-4 stratigraphic subunit represents a small group, most
likely composed of 1013 individuals, and 90% of the footprints correspond to
children or adolescents (Duveau et al. 2019). This high proportion of children and
adolescents raises questions about the distribution of activities (hunting, carcass
transport, lithic industry, etc.) within the group. It is also currently impossible to
know why so few adults were on the site at this time. Future analyses of the spatial
distribution of footprints and their relationship to associated archaeological remains
could provide valuable information on these important issues. Importantly, the two
last eld missions in 2018 and 2019 allowed the discovery of around 800 new
potential footprints (most of them coming from the D3b-4 subunit). Ongoing studies
of these new tracks will rst have to validate or not their identication as hominin
footprints. Then, morphometric analyses will aim to clarify our knowledge of the
size and the composition of the groups who occupied Le Rozel 80,000 years ago.
To sum up, the tracks discovered at Le Rozel represent the most important
ichnological assemblage attributed to Neandertals to date and more generally the
most important for hominin taxa other than Homo sapiens. The analysis on the
footprints provides not only essential data in order to understand the Palaeolithic
occupations at Le Rozel 80,000 years ago but also could provide access to unique
information on the composition of groups at a timescale unusual in prehistoric
archaeology, that of a snapshot. In this perspective, the crossing of ichnological
data with archaeological data (occupation structures, spatial distribution of activities,
etc.) will bring closer to the life history of the Pleistocene human groups.
Acknowledgements Data concerning the site of Le Rozel was communicated by Dominique
Cliquet in charge of the excavation and included in this contribution with his formal agreement;
we thank him for this. Furthermore, we thank the owners of the Le Rozel site, Mrs. Lecomte, Mrs.
Deregeaucourt, Mrs. Guillotte and Mrs. Maurouard. We wish to acknowledge all the volunteers
who participated in the excavations, making the discovery and analysis of the archaeological and
ichnological material possible. We are grateful to M. Friess and F. Detroit (Musée de lHomme,
Paris, France) for their help during the photogrammetric digitization. The eld work is supported by
the French Ministry of Culture and the French department of La Manche, and data acquisition has
been funded by the CNRSInstitut Écologie et Environnement International Research Network
IRN-GDRI0870. The rst author was granted funding for a doctoral degree (Muséum national
dHistoire naturelle, Paris, France).
196 J. Duveau et al.
References
Aiello, L., & Dean, C. (1990). An introduction to human evolutionary anatomy. New York:
Academic Press.
Allen, J. R. (1997). Subfossil mammalian tracks (Flandrian) in the Severn Estuary, SW Britain:
Mechanics of formation, preservation and distribution. Philosophical Transactions of the Royal
Society of London B: Biological Sciences, 352(1352), 481518.
Altamura, F., Bennett, M. R., DAoût, K., Gaudzinski-Windheuser, S., Melis, R. T., et al. (2018).
Archaeology and ichnology at Gombore II-2, Melka Kunture, Ethiopia: Everyday life of a
mixed-age hominin group 700,000 years ago. Scientic Reports, 8(1), 2815.
Ashton, N., Lewis, S. G., De Groote, I., Duffy, S. M., Bates, M., et al. (2014). Hominin footprints
from early Pleistocene deposits at Happisburgh, UK. PLoS One, 9(2), e88329.
Bahain, J.-J., Falgueres, C., Laurent, M., Dolo, J.-M., Shao, Q., et al. (2015). ESR/U-series dating of
faunal remains from the paleoanthropological site of Biache-Saint-Vaast (Pas-de-Calais,
France). Quaternary Geochronology, 30, 541546.
Bahn, P. G. (1998). The Cambridge illustrated history of prehistoric art. Cambridge: Cambridge
University Press.
Bang, P., Dahlstrøm, P., & Walters, M. (2001). Animal tracks and signs. Oxford: Oxford University
Press.
Benazzi, S., Douka, K., Fornai, C., Bauer, C. C., Kullmer, O., et al. (2011). Early dispersal of
modern humans in Europe and implications for Neandertal behaviour. Nature, 479(7374), 525.
Bennett, M. R., & Morse, S. A. (2014). Human footprints: Fossilised locomotion? Heidelberg:
Springer.
Bennett, M. R., Harris, J. W., Richmond, B. G., Braun, D. R., Mbua, E., et al. (2009). Early
hominin foot morphology based on 1.5-million-year-old footprints from Ileret, Kenya. Science,
323(5918), 11971201.
Bennett, M. R., Gonzalez, S., Huddart, D., Kirby, J., & Toole, E. (2010). Probable Neolithic
footprints preserved in inter-tidal peat at Keng, South Wales (UK). Proceedings of the
GeologistsAssociation, 121,6676.
Berillon, G. (2000). Le pied des hominoïdes miocènes et des hominidés fossiles: Architecture,
locomotion et évolution. Paris: CNRS.
Bustos, D., Jakeway, J., Urban, T. M., Holliday, V. T., Fenerty, B., et al. (2018). Footprints preserve
terminal Pleistocene hunt? Human-sloth interactions in North America. Science Advances, 4(4),
eaar7621.
Citton, P., Romano, M., Salvador, I., & Avanzini, M. (2017). Reviewing the upper Pleistocene
human footprints from the Sala dei Misteriin the Grotta della Bàsura (Toirano, Northern Italy)
cave: An integrated morphometric and morpho-classicatory approach. Quaternary Science
Reviews, 169,5064.
Cliquet, D., Auguste, P., Coutard, J.-P., Dupret, L., Duveau, J., et al. (2018a). Le site paléolithique
moyen du « Pou » au Rozel (Manche) : des aires de travaux spécialisés et des habitats vieux
denviron 80 000 ans. Un premier bilan. In J.-P. Olivier (Ed.), Journées archéologiques de
Normandie, Rouen, 30 septembre-1er octobre 2016 (pp. 1335). Mont Saint Agnan: Presses
universitaires de Rouen et du Havre.
Cliquet, D., Coutard, J.-P., Van Vliet-Lanoë, B., Jamet, G., Auguste, P., et al. (2018b). Le Rozel
(Manche), Le Pou, Section B n486 et 572 (Rapport de fouille). Caen: SRA Normandie.
Crompton, R. H., & Pataky, T. C. (2009). Stepping out. Science, 323, 11741175.
11 On the Tracks of Neandertals: The Ichnological Assemblage from Le Rozel... 197
Crompton, R. H., Pataky, T. C., Savage, R., DAoût, K., Bennett, M. R., et al. (2011). Human-like
external function of the foot, and fully upright gait, conrmed in the 3.66 million year old
Laetoli hominin footprints by topographic statistics, experimental footprint-formation and
computer simulation. Journal of the Royal Society Interface, 9(69), 707719.
de Lumley, H., & Vicino, G. (1984). New data concerning the dating and interpretation of human
footprints present in the grotta della basuraat toirano (Savona, Northern Italy). Results of an
international round table. Journal of Human Evolution, 13(6), 537540.
Delson, E., & Harvati, K. (2006). Palaeoanthropology: Return of the last Neandertal. Nature,
443(7113), 762.
Devièse, T., Karavanić, I., Comeskey, D., Kubiak, C., Korlević, P., et al. (2017). Direct dating of
Neandertal remains from the site of Vindija cave and implications for the middle to upper
Paleolithic transition. PNAS, 114(40), 1060610611.
Dingwall, H. L., Hatala, K. G., Wunderlich, R. E., & Richmond, B. G. (2013). Hominin stature,
body mass, and walking speed estimates based on 1.5 million-year-old fossil footprints at Ileret,
Kenya. Journal of Human Evolution, 64(6), 556568.
Duveau, J., Berillon, G., Verna, C., Laisné, G., & Cliquet, D. (2019). The composition of a
Neandertal social group revealed by the hominin footprints at Le Rozel (Normandy, France).
PNAS, 116(39), 1940919414.
Falkingham, P. L. (2014). Interpreting ecology and behaviour from the vertebrate fossil track
record. Journal of Zoology, 292, 222228.
Farizy, C. (1994). Spatial patterning of middle Paleolithic sites. Journal of Anthropological
Archaeology, 13(2), 153160.
Finlayson, C., Pacheco, F. G., Rodríguez-Vidal, J., Fa, D. A., López, J. M. G., et al. (2006). Late
survival of Neandertals at the southernmost extreme of Europe. Nature, 443(7113), 850.
Guipert, G., de Lumley, M. A., Tuffreau, A., & Mafart, B. (2011). A late middle Pleistocene
hominid: Biache-saint-Vaast 2, North France. Comptes Rendus Palevol, 10(1), 2133.
Guthrie, R. D. (2005). The nature of Paleolithic art. Chicago: University of Chicago Press.
Harvati, K., & Roksandic, M. (2016). The human fossil record from Romania: Early upper
Paleolithic European mandibles and Neandertal admixture. In K. Harvati & M. Roskandic
(Eds.), Paleoanthropology of the Balkans and Anatolia (pp. 5168). Dordrecht: Springer.
Hasiotis, S. T., Platt, B. F., Hembree, D. I., & Everhart, M. J. (2007). The trace-fossil record of
vertebrates. In S. T. Hasiostis et al. (Eds.), Trace fossils: Concepts, problems, prospects
(pp. 196218). Amsterdam: Elsevier.
Higham, T., Compton, T., Stringer, C., Jacobi, R., Shapiro, B., et al. (2011). The earliest evidence
for anatomically modern humans in northwestern Europe. Nature, 479(7374), 521.
Hoffmann, D. L., Standish, C. D., García-Diez, M., Pettitt, P. B., Milton, J. A., et al. (2018). U-Th
dating of carbonate crusts reveals Neandertal origin of Iberian cave art. Science, 359(6378),
912915.
Hublin, J.-J. (2015). The modern human colonization of western Eurasia: When and where?
Quaternary Science Reviews, 118, 194210.
Jones, L. A., & Lederman, S. J. (2006). Human hand function. Oxford: Oxford University Press.
Kim, J. Y., Kim, K. S., Lockley, M. G., & Matthews, N. (2008). Hominid ichnotaxonomy: An
exploration of a neglected discipline. Ichnos, 15(34), 126139.
Klenerman, L., & Wood, B. (2006). The human foot: A companion to clinical studies. Heidelberg:
Springer.
Kuttruff, J. T., DeHart, S. G., & OBrien, M. J. (1998). 7500 years of prehistoric footwear from
Arnold research cave, Missouri. Science, 281(5373), 7275.
Leakey, M. D., & Hay, R. L. (1979). Pliocene footprints in the Laetolil Beds at Laetoli, Northern
Tanzania. Nature, 278(5702), 317323.
Ledoux, L., Fourment, N., Maksud, F., Delluc, M., & Costamagno, S. (2017). Traces of human and
animal activity (TrAcs) in Cussac cave (Le Buisson-de-Cadouin, Dordogne, France): Prelimi-
nary results and perspectives. Quaternary International, 430, 141154.
198 J. Duveau et al.
Lockley, M., Roberts, G., & Kim, J. Y. (2008). In the footprints of our ancestors: An overview of
the hominid track record. Ichnos, 15(34), 106125.
Lockley, M., Meldrum, J., & Kim, J. Y. (2016). Major events in hominin evolution. In M. G.
Mángano & L. A. Buatois (Eds.), The trace-fossil record of major evolutionary events volume 2:
Mesozoic and Cenozoic springer (pp. 411448). Dordrecht: Springer.
Manolis, S., Aiello, L., Henessy, R., & Kyparissi-Apostolika, N. (2000). Middle Palaeolithic
footprints from Theopetra cave (Thessaly, Greece). In N. Kyparissi-Apostolika (Ed.), Theopetra
cave. Twelve years of excavation and research 19871998 (pp. 8793). Greek Ministry of
Culture and Institute for Aegean Prehistory: Athens.
Masao, F. T., Ichumbaki, E. B., Cherin, M., Barili, A., Boschian, G., et al. (2016). New footprints
from Laetoli (Tanzania) provide evidence for marked body size variation in early hominins.
eLife, 5, e19568.
Mastrolorenzo, G., Petrone, P., Pappalardo, L., & Sheridan, M. F. (2006). The Avellino 3780-yr-BP
catastrophe as a worst-case scenario for a future eruption at Vesuvius. PNAS, 103(12),
43664370.
McLaren, D., Fedje, D., Dyck, A., Mackie, Q., Gauvreau, A., et al. (2018). Terminal Pleistocene
epoch human footprints from the Pacic coast of Canada. PLoS One, 13(3), e0193522.
Mercier, N., Martin, L., Kreutzer, S., Moineau, V., & Cliquet, D. (2019). Dating the palaeolithic
footprints of Le Rozel(Normandy, France). Quaternary Geochronology, 49, 271277.
Mietto, P., Avanzini, M., & Rolandi, G. (2003). Palaeontology: Human footprints in Pleistocene
volcanic ash. Nature, 422(6928), 133.
Molleson, T. I., Oakley, K. P., & Vogel, J. C. (1972). The antiquity of the human footprints of Tana
della Basura. Journal of Human Evolution, 1(5), 467471.
Moniz Nakamura, J. J. (2009). Hominid footprints in recent volcanic ash: New interpretations from
Hawaii Volcanoes National Park. Ichnos, 16(12), 118123.
Morse, S. A., Bennett, M. R., Gonzalez, S., & Huddart, D. (2010). Techniques for verifying human
footprints: Reappraisal of pre-Clovis footprints in Central Mexico. Quaternary Science Reviews,
29, 25712578.
Morse, S. A., Bennett, M. R., Liutkus-Pierce, C., Thackeray, F., McClymont, J., et al. (2013).
Holocene footprints in Namibia: The inuence of substrate on footprint variability. American
Journal of Physical Anthropology, 151(2), 265279.
Muñiz, F., Cáceres, L. M., Rodríguez-Vidal, J., de Carvalho, C. N., Belo, J., et al. (2019). Following
the last Neandertals: Mammal tracks in Late Pleistocene coastal dunes of Gibraltar (S Iberian
Peninsula). Quaternary Science Reviews, 217, 297309.
Murie, O. J., & Elbroch, M. (2005). Aeld guide to animal tracks. Singapore: Houghton Mifin
Harcourt.
Nigst, P. R., Haesaerts, P., Damblon, F., Frank-Fellner, C., Mallol, C., et al. (2014). Early modern
human settlement of Europe north of the Alps occurred 43,500 years ago in a cold steppe-type
environment. PNAS, 111(40), 1439414399.
Onac, B. P., Viehmann, I., Lundberg, J., Lauritzen, S.-E., Stringer, C., et al. (2005). UTh
ages constraining the Neandertal footprint at Vârtop cave, Romania. Quaternary Science
Reviews, 24(1011), 11511157.
Pales, L. (1954). Les empreintes de pieds humains de la Tana della Bàsura(Toirano). Rivista di
studi liguri, 20,512.
Pales, L. (1960). Les empreintes de pieds humains de la grotta della Basura.Rivista di studi liguri,
26,2590.
Panarello, A., Mazzardo, L., & Mietto, P. (2018). The Devils Touch: A rst dataset from what
could be the oldest handprint ever found (Central-Southern Italy). Alpine and Mediterranean
Quaternary, 31(1), 3747.
Pastoors, A., Lenssen-Erz, T., Ciqae, T., Kxunta, U., Thao, T., et al. (2015). Tracking in caves:
Experience based reading of Pleistocene human footprints in French caves. Cambridge Archae-
ological Journal, 25(3), 551564.
11 On the Tracks of Neandertals: The Ichnological Assemblage from Le Rozel... 199
Pastoors, A., Lenssen-Erz, T., Breuckmann, B., Ciqae, T., Kxunta, U., et al. (2017). Experience
based reading of Pleistocene human footprints in Pech-Merle. Quaternary International, 430,
155162.
Pettitt, P. B. (1997). High resolution Neandertals? Interpreting middle palaeolithic intrasite spatial
data. World Archaeology, 29(2), 208224.
Pinhasi, R., Gasparian, B., Areshian, G., Zardaryan, D., Smith, A., et al. (2010). First direct
evidence of chalcolithic footwear from the near eastern highlands. PLoS One, 5(6), e10984.
Pinhasi, R., Higham, T. F., Golovanova, L. V., & Doronichev, V. B. (2011). Revised age of Late
Neandertal occupation and the end of the middle Paleolithic in the northern Caucasus. PNAS,
108(21), 86118616.
Roach, N. T., Hatala, K. G., Ostrofsky, K. R., Villmoare, B., Reeves, J. S., et al. (2016). Pleistocene
footprints show intensive use of lake margin habitats by Homo erectus groups. Scientic
Reports, 6, 26374.
Rougier, H. (2003). Étude descriptive et comparative de Biache-Saint-Vaast 1 (Biache-Saint-Vaast,
Pas-de-Calais, France). PhD thesis, Université de Bordeaux I.
Schmincke, H.-U., Rausch, J., Kutterolf, S., & Freundt, A. (2010). Walking through volcanic mud:
The 2,100 year-old Acahualinca footprints (Nicaragua) II: The Acahualinca people, environ-
mental conditions and motivation. International Journal of Earth Sciences, 99(1), 279292.
Scuvée, F., & Verague, J. (1984). Paléolithique supérieur en Normandie occidentale: lAbri-sous-
Roche de la Pointe du Rozel (Manche). Cherbourg: LITTUS CEHP.
Sévêque, N. (2017). Variabilité des comportements alimentaires au Paléolithique moyen en France
septentrionale: apports des études archéozoologiques. PhD thesis, Unviersité de Lille 3.
Socaru, A., Petrea, C., Doboş, A., & Trinkaus, E. (2007). The human cranium from the Peştera
Cioclovina Uscată, Romania: Context, age, taphonomy, morphology, and paleopathology.
Current Anthropology, 48(4), 611619.
Stoetzel, E., Koehler, H., Cliquet, D., Sévêque, N., & Auguste, P. (2016). New data on Late
Pleistocene small vertebrates from northern France. Comptes Rendus Palevol, 15(6), 681695.
Trinkaus, E., & Shang, H. (2008). Anatomical evidence for the antiquity of human footwear:
Tianyuan and Sunghir. Journal of Archaeological Science, 35(7), 19281933.
Trinkaus, E., Churchill, S. E., Villemeur, I., Riley, K. G., Heller, J. A., et al. (1991). Robusticity
versus shape: The functional interpretation of Neandertal appendicular morphology. Journal of
the Anthropological Society of Nippon, 99(3), 257278.
Trinkaus, E., Moldovan, O., Bîlgăr, A., Sarcina, L., Athreya, S., et al. (2003). An early modern
human from the Peştera cu Oase, Romania. PNAS, 100(20), 1123111236.
Tuffreau, A. (1978). Les fouilles du gisement paléolithique de Biache-Saint-Vaast (Pas-de-Calais):
années 1976 et 1977-premiers résultats. Quaternaire, 15(1), 4655.
Tuffreau, A. (1988). Les habitats du Paléolithique inférieur et moyen dans le Nord de la France
(Nord, Pas-de-Calais, Somme). Revue archéologique de Picardie, 1(1), 91104.
Valladas, H., Mercier, N., Froget, L., Joron, J.-L., Reyss, J.-L., et al. (2007). TL age-estimates for
the middle Palaeolithic layers at Theopetra cave (Greece). Quaternary Geochronology, 2(14),
303308.
van Vliet-Lanoë, B., Cliquet, D., Auguste, P., Folz, E., Keen, D., et al. (2006). Labri sous-roche du
Rozel (France, Manche): un habitat de la phase récente du Paléolithique moyen dans son
contexte géomorphologique. Quaternaire, 17(3), 207258.
Viehmann, I. (1987). Prehistoric human footprints in Romanias caves. Theoretical and Applied
Karstology, 3, 229234.
Webb, S. (2007). Further research of the Willandra Lakes fossil footprint site, southeastern
Australia. Journal of Human Evolution, 52(6), 711.
Webb, S., Cupper, M. L., & Robins, R. (2006). Pleistocene human footprints from the Willandra
Lakes, southeastern Australia. Journal of Human Evolution, 50(4), 405413.
Webb, D., Robu, M., Moldovan, O., Constantin, S., Tomus, B., et al. (2014). Ancient human
footprints in Ciur-Izbuc cave, Romania. American Journal of Physical Anthropology, 155(1),
128135.
200 J. Duveau et al.
Wood, R. E., Barroso-Ruíz, C., Caparrós, M., Pardo, J. F. J., Santos, B. G., et al. (2013).
Radiocarbon dating casts doubt on the late chronology of the middle to upper Palaeolithic
transition in southern Iberia. PNAS, 110(8), 27812786.
Zhang, D. D., & Li, S. H. (2002). Optical dating of Tibetan human hand-and footprints: An
implication for the palaeoenvironment of the last glaciation of the Tibetan Plateau. Geophysical
Research Letters, 29(5), 1611.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Chapter 12
Hominin Footprints in Caves from
Romanian Carpathians
Bogdan P. Onac, Daniel S. Veres, and Chris Stringer
Abstract The Romanian karst hosts numerous caves and shelters that over time
provided remarkable archaeological and anthropological vestiges. Altogether they
show that humans must have entered caves in Romania at least as early as
170,000 years ago. However, ancient human footprints are very rare in the fossil
record of East-Central Europe, with only two known locations in the Apuseni
Mountains of western Romania. Vârtop Cave site originally preserved three fossil
footprints made about 67,800 years ago by a Homo neanderthalensis, whereas Ciur
Izbuc Cave was probably home of early H. sapiens that left almost 400 footprints
(interspersed with spoors of cave bears), which were indirectly dated to be younger
than ~36,500 years.
Keywords Karst · Cave · Prehistoric people · Footprints · Romania
Introduction
The major karst areas of Romania occur in the East and South Carpathians, Apuseni
Mountains, and Dobrogea (Onac and Goran 2019) (Fig. 12.1), all hosting key
archaeological cave sites (for comprehensive reviews, see Boroneanț2000;
Anghelinu and Boroneanț2019). Over the last 150 years, archaeological and
paleontological researches focused on a signicant number of shelters and cavities,
most of them concentrated in the south-western part of the South Carpathians and
B. P. Onac (*)
School of Geosciences, University of South Florida, Tampa, FL, USA
Emil RacovițăInstitute of Speleology, Cluj-Napoca, Romania
e-mail: bonac@usf.edu
D. S. Veres
Emil RacovițăInstitute of Speleology, Cluj-Napoca, Romania
C. Stringer
Centre for Human Evolution Research, Natural History Museum, London, UK
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_12
201
202 B. P. Onac et al.
Fig. 12.1 (a) Romania within Europe; (b) location of Ciur Izbuc and Vârtop caves in Apuseni
Mountains karst
Apuseni Mountains that also produced human remains (e.g. Cioclovina, Muierii,
Polovragi, Oase). These investigations documented important Middle to Upper
Palaeolithic sites, with the latter ones being far more abundant (Mertens 1996;
Cârciumaru 1999;Păunescu 2001). Other archaeological and anthropological nd-
ings indicate that the early modern humans had a more constant presence in the
Romanian caves (Cârciumaru 1988; Trinkaus et al. 2003; Olariu et al. 2005;Socaru
et al. 2007; Clottes et al. 2012; Webb et al. 2014; Harvati and Roksandic 2016). Until
very recently, the cave-based Middle and Upper Palaeolithic in Romania offered
almost exclusively archaeological collections with limited reliable chronological
control (Cosac et al. 2018; Anghelinu and Boroneanț2019). Abri 122 from Vârghiș
karst (East Carpathians; Veres et al. 2018) has produced so far the most important
Middle Palaeolithic lithic assemblage in the Carpathian region, including evidence
of use-wear on fragmented bone tools and cut marks on a bos/bison tibial diaphysis
(Cosac et al. 2018). Multiple-method luminescence dating indicates that human use
of this site commenced sometime between 141 ±12 ka and 174 ±37 ka (Veres et al.
2018). These ages corroborate other evidence of Middle Palaeolithic occurrences in
that chronological span within the Dobrogean karst and loess records near the Black
Sea (Balescu et al.2015).
Except for a handful of sites worldwide, human footprints are not that common in
the fossil record. Onac et al. (2005) showed that the earliest footprints documenting
direct human incursions into a Carpathian cave also come from Middle Palaeolithic
(i.e. Homo neanderthalensis) and date back to more than 62,000 years ago. The only
other ancient footsteps preserved in a Romanian cave were made in soft clay that
partly hardened and then remained undisturbed until recently (Rusu et al. 1969;
Rișcuția and Rișcuția 1970). Their age could be as old as 36,500 years, which heralds
them as the oldest direct traces left by anatomically modern humans in a European
cave (Webb et al. 2014). The two sites from which these footprints were
documented, i.e. Vârtop and Ciur Izbuc caves in western Apuseni Mountains
(Fig. 12.1), deserve further attention.
12 Hominin Footprints from Romania 203
Vârtop Cave (Bihor Mountains)
Forty-ve years ago, cavers from Emil RacovițăSpeleological Club in Cluj-Napoca,
led by the late Iosif Viehmann, organized a winter camp at Casa de Piatră(Stone
House), a remote hamlet that counts only a few scattered houses in the heart of the
Bihor Mountains. One of the objectives of this camp was a study visit into the Vârtop
Glacier Cave (hereafter Vârtop), a short (340 m), but very well-decorated cavity
discovered in 1955 and declared a natural monument in 1957 (Bleahu and Viehmann
1963). In the middle section of the cave, just before entering the Domes Room (Sala
Domului), a small, east-trending side passage opens, with its oor covered almost
completely by a shallow lake. As the access to this section of the cave was somewhat
more difcult prior to this expedition, no one ventured beyond the small chamber
hosting the lake. However, in February 1974, Iulia Szekely and Ioan Bucur passed
the lake and climbed a steep owstone in the north-eastern part of the Lake Room
(Sala Lacului). After barely passing between two large stalagmites, they entered a
rather small, low-ceilinged chamber. Just behind the stalagmite obstruction, over a
at surface of ca. 1.5 m
2
, they noticed a well-preserved single human footprint
(Fig. 12.2a). When I. Viehmann investigated the site, two other less clear prints were
noticed, one of a heel and the other made by the toes. A few months later, the cave
site was visited by Cantemir Rișcuția, a well-known Romanian anthropologist who
after some preliminary ichnological measurements suggested a possible age of
ca. 15,000 years (Viehmann 1975). After numerous other biometric measurements
and photographs were taken in situ, the decision was made to cut out the best
preserved footprint and safeguard it in the Museum of the Institute of Speleology
Fig. 12.2 (a) Vârtop footprint (dashed line represents the CT cross-section shown in b); (b)
transversal CT image of the footprint displaying the stalagmite (stg), embedded soda-straws (ss),
and the U-series ages (in thousands of years)
in Cluj-Napoca. Although the decision to remove the print from the cave sparked
controversy at the time, a few years later, when the illegal disappearance of the other
traces was discovered, the usefulness of the undertaken approach was understood.
204 B. P. Onac et al.
Room of the Steps and the Vârtop Footprints
The area of the cave that hosts the footprints consists of a small chamber (Room of
the Steps) that continues into a short ascending gallery, with its oor covered with
limestone boulders and red-brownish clay. The presence of these materials indicates
an older cave entrance at the upper end of this corridor that collapsed more than
15,000 years ago (Onac et al. 2005). The age has been established after dating the
base of one of the scattered stalagmites growing over the clayed cave oor by means
of U-series technique. The existence of a different cave access point makes total
sense, since it is unlikely that the prehistoric human crawled and climbed into the
Room of the Steps using the present-day cave entrance.
The footprints were fossilized into a moonmilk deposit that accumulated between
the cave wall and an alignment of stalagmites (Fig. 12.2). At the time the human left
the prints, the moonmilk blanket covering the oor must have been soft and pliable,
but later hardened into a calcareous tufa type deposit. The best preserved footprint is
22 cm in length and rather wide (10.6 cm) and shows a wide gap (1.6 cm) between
the great toe of the foot and the rest of the toes (Fig. 12.2a). This is not necessary a
distinctive feature of the foot (i.e. hallux varus), but likely the gap formed when
stepping in soft clay, barefoot. This could also be the reason for the overall width of
the footprint. These two observations and a comparison with the human footprint
from Bàsura Cave (Toirano, Italy), then assigned to a Neanderthal (Blanc and Pales
1960), led Viehmann (1987) to suggest (without any dating information) that the
Vârtop Cave footprint is ca. 80,000 years old. However, the antiquity of the human
footprints discovered in the Italian cave was revised (based on radiometric dating) to
be just 14,000 to 12,000 years old (Molleson et al. 1972; De Lumley et al. 1984), a
fact that called for a re-evaluation and a better way to estimate the age of the
Vârtop Man.
Geochronology
As described below, a suite of favourable settings allowed an international group of
researcher to successfully date the Vârtop footprint using the U-series method (Onac
et al. 2005). The moonmilk deposit accumulated in the Room of the Steps was an
ideal surface and material for casting human footprints, especially because it hard-
ened, becoming a compact calc-tufa layer. Computer tomography (CT) imaging
suggests the upper 1 cm of this deposit has a low density (grey), followed by a higher
density (brighter) indicating less porous calcite, and a more compact layer (1 cm).
The lower part (~4 cm) instead shows a rather low density (dark) material, which
would correspond to a porous but homogeneous texture (Fig. 12.2b). Our interpre-
tation of the CT is that the human stepped in the denser and mechanically more
competent layer, causing lateral displacement of the softer material below. Since the
moonmilks porous nature is far from ideal for U-series dating, we applied the
isochron method to correct for admixed detritus with uniform
230
Th/
232
Th. Based
on seven coeval subsamples having different U and Th concentrations and conse-
quently distinct detrital components, an isochron age of 97,000 years (1σ) was
obtained for the lower 5 cm of calc-tufa deposit. Statistically speaking, the age is
not very robust due to a large uncertainty, but nevertheless implies a rapid accumu-
lation of the moonmilk sometime during MIS 5.
12 Hominin Footprints from Romania 205
Constraining the footprint age was possible due to the presence of a small
stalagmite that grew over the footprint mould right below the big toe (Fig. 12.2)
and a piece of soda straw embedded in the calc-tufa layer directly overlying the
human print. The latter one was revealed by the CT scan (Fig. 12.2), which also
showed the depth to which the footprint was imprinted on moonmilk. The soda straw
from this undated layer returned a U-Th age of ~67,800 years. Three ages obtained
from the base of the small stalagmite that was growing in the footprint mould cluster
around 62,000 years. The last layer of moonmilk that partly lled the footprint was
dated to 22,300 years, whereas a calcite fragment from of a soda straw cemented on
the surface of the uppermost calc-tufa layer appears to have formed 20,000 years ago
and then broke and fell to the oor. To further consolidate the chronology of the
entire sequence, the base of a stalagmite which precipitated directly over the reddish
clayey oor was dated to 15,400 years. This could be considered the earliest time at
which the old entrance collapsed, preventing soil and other sediments from entering
the cave.
Based on the calc-tufa stratigraphy and the above chronology, our interpretation
of the Vârtop footprint is as follows: some 97,000 years ago, a period with
documented speleothem growth near Vârtop Cave (Onac 2001), and other parts of
Romania (Onac and Lauritzen 1996), moonmilk accumulated on the oor of the
Room of the Steps. A prehistoric human entered Vârtop Cave using a different
entrance than today and left her/his footprints impressed in the upper, more compe-
tent layer of the moonmilk deposit not earlier than 67,800 years ago, when a soda
straw of this age fell off the cave ceiling and was later embedded in a thin (undated)
moonlike layer that covers the footprint. In Romania, the period between 78,000 and
67,000 years ago was mild and wet, favouring speleothem precipitation (Onac and
Lauritzen 1996; Staubwasser et al. 2018). Similar conditions must have existed
~62,000 years ago when the small stalagmite nested in the middle part of the
footprint mould begun its growth. Since the publication of the original paper
reporting these footprints (Onac et al. 2005), the newly dated soda straw
(67,800 years) indisputably conrms that the Vârtop prints cannot be younger than
67,800 years; thus they clearly belong to a Homo neanderthalensis.
206 B. P. Onac et al.
Ciur Izbuc Cave (Pădurea Craiului Mountains)
Ciur Izbuc Cave is part of the Toplița-Ciur-Tinoasa karst system located in the south-
eastern part of the Pădurea Craiului Mountains on the Runcuri Karst Plateau (Rusu
et al. 1970). Although the cave entrance must have been known to locals for
centuries, the rst documented visit happened in 1962 when T. Rusu,
I. Viehmann, and S. Avram surveyed ~150 m out of its total length of 1030 m
(Viehmann et al. 1970). During the exploration and mapping of the cave, a team of
researchers from the Emil RacovițăInstitute of Speleology in Cluj-Napoca
(I. Viehmann, T. Rusu, G. Racoviță, and V. Crăciun) discovered in November
11, 1965, about 400 barefooted human footprints (Rusu et al. 1969; Viehmann
et al. 1970). These imprints are interspersed with cave bear (Ursus spelaeus)
footmarks in the clayey oor of the caves upper level, in what is now known as
Sala Pașilor (Footprint Room). Within 3 years from the time of the discovery, ~230
of the best-preserved prints were tagged with numbered metal ags, some of which
still at their original position (Fig. 12.3). This process served two purposes,
(i) inventory (for systematic observations) and (ii) raising awareness (protection),
to those entering the cave. Nevertheless, decades of indiscriminate visitation of the
cave led to the disappearance of many of these ags. Since most footprint casts were
anyway hard to see in the red-brown clay and others became lled with bat guano,
covered by a sub-millimetre-thick calcite dust, or affected by mud cracks, many of
them were damaged or even completely destroyed.
Following the discovery of the human footprints in Ciur Izbuc Cave, an article
announcing the ndings was published by Rusu et al. (1969) in Ocrotirea Naturii
(Nature Conservation), a Romanian popular science magazine. Despite the
nontechnical character of the publication, the paper includes very important
Fig. 12.3 (a) Photo of the Footprint Room in Ciur Izbuc Cave showing part of the tagged footprints
(Photograph A. Posmoșanu); (b) close-up view of a well-preserved footprint. (Photograph
G. Ponta)
scientic information regarding the evolution of the cave, documents the traces left
by the cave bears, and for the rst time illustrates the human footprints. It was also
noticed that there was a lack of any footprints or cave bear bones between the current
entrance and the Footprint Room, whose northern end is only 50 m away and 810 m
below the sinking point in which Tinoasa stream disappears into the cave. Due to this
geomorphological setting and because the Footprint Room is too far and difcult to
reach using the present day cave entrance, it has been speculated that in the past,
humans and cave bears likely used a different access point (Rusu et al. 1969).
Relying solely on the presence of some polished cave bear bones (used as tools?)
and a human gurine rudimentarily engraved on the root of a cave bear canine tooth,
the authors attributed the footmarks to a Homo sapiens who lived ~15,000 to
~10,000 years ago.
12 Hominin Footprints from Romania 207
Three other studies appeared in a book printed on the occasion of Emil Racovițăs
(founder of the worldsrst Speleological Institute in Cluj, Romania) birth cente-
nary. The paper by Rusu et al. (1970) tackles the geomorphology and hydrology of
the Toplița-Ciur-Tinoasa karst system but also includes a paragraph on the prehis-
toric footprints, along with a photograph. Viehmann et al. (1970) present a couple of
observations that shed light on the presence of human and cave bear footmarks.
Without having any radiocarbon ages, but from the apparent relationship between
the human and Ursus spelaeus prints, the authors claimed that the human footmarks
must be younger than those of Solutrean people (21,000 to 17,000 years). The same
study suggested based on the large number of prints that the visits were not
occasional and the humans had deliberately entered the cave.
The rst standard ichnological analyses were undertaken by Rișcuția and Rișcuția
(1970), who measured ve morphometrical parameters for 188 footprints. By
relating the maximum length of the foot (Fl) with the height of individuals (h),
using the classic relationship Fl ¼15% h, the authors concluded that two adults and
a child were the Ciur Izbuc cave trackmakers. They reported a height of 157 cm for
the woman and 174.9 cm for the man, but for the child, they only indicated an age
range (9 to 11 years old).
A more recent study conducted by Webb et al. (2014) measured the width for the
ball and heel and the maximum length of 51 footprints that were still visible on the
cave oor. Using the print lengths (range between 157 and 318 mm), the authors
estimated the minimum number of individuals and their stature range. Contrary to
the previous studies, Webb et al. (2014) suggest a group of six to seven individuals
left their footprints in Ciur Izbuc Cave. Considering that the printmakers travelled
only ~75 m from the former cave entrance towards the inner part of the Footprint
Room, the same study concluded that it would have taken 9 min for an individual
(and far less for 67 people) to leave behind those 400 footprints originally counted.
The estimations regarding the human stature (calculated by either regression or
percentage method) overlapped well, both studies reporting heights between 106.4
and 216.1 cm.
The real novelty in the study of Webb et al. (2014) is the approach taken by the
authors to estimate the age of the footprints. In the absence of any artefacts or human
remains, the direct dating of the tracks was impossible. Nevertheless, considering
that a few human footmarks appear to be overprinted by cave bears, and having
radiocarbon dated two bear bones, the study concluded that the Ciur Izbuc people
might have ventured in the cave anytime since ~36,500 years ago. Based on this age,
the authors suggest that the footprints belong to either early H. sapiens or sapiens/
neanderthalensis hybrids, but without being able to place them in a clear cultural
context. It is now known that even if the humans were as old as 36,500 years, they
would probably have been too young to be direct hybrids, as they post-dated the last
known appearances of Neanderthals in Europe (Higham et al. 2014).
208 B. P. Onac et al.
It is not surprising that human footprints have been found within the Romanian
Carpathian caves. The area has often been considered a refugial area for humans and
ecosystems during stadials (Staubwasser et al. 2018), and the potential dispersal
routes into Central Europe intersect those north of the Black Sea along the
Carpathian arch and the Danube Valley (e.g. Iovita et al. 2012). As such, south-
eastern Europe has long been considered one of the most likely routes for hominin
spreads across the continent, including anatomically modern humans, with the Oase
Cave fossils amongst the oldest modern human fossils in Europe (Trinkaus et al.
2003). It is thus expected that more intensive research will signicantly augment the
number of cave archaeological sites, as well as our understanding of migration
routes, genetic turnover, and past human population dynamics.
Acknowledgements A. Posmoșanu and G. Ponta are thanked for providing the human footprints
photographs from the Sala Pașilor in Ciur Izbuc Cave. Chris Stringers research is supported by the
Calleva Foundation and the Human Origins Research Fund.
References
Anghelinu, M., & Boroneanț, A. (2019). The archaeology of caves in Romania. In G. M. L. Ponta &
B. P. Onac (Eds.), Cave and karst systems of Romania (pp. 501517). Springer: Cham.
Balescu, S., Tuffreau, A., & Dobrescu, R. (2015). Mamaia Sat (Romania): A late middle Pleisto-
cene Palaeolithic site? LAnthropologie, LIII(12), 157166.
Blanc, A. C., & Pales, L. (1960). Le vestigia umane della Grotta della Basura a Toirano. Rivista di
Studi Liguri, 26,190.
Bleahu, M., & Viehmann, I. (1963). Ghețarul de la Vârtop. Ocrotirea Naturii, 7, 147152.
Boroneanț, V. (2000). Arheologia peșterilor și minelor din România. București: Institutul de
Memorie Culturală(cIMeC).
Cârciumaru, M. (1988). Lart pariétal préhistorique en Roumanie. LAnthropologie, 92(1),
239254.
Cârciumaru, M. (1999). Le Paléolithique en Roumanie. Grenoble: Editions Jérôme Millon.
Clottes, J., Besesek, M., Gely, B., Kenesz, M., Lascu, V. T., Meyssonnier, M., Philippe, M.,
Plichon, M., Prudhomme, F., Radu, V. A., Rus, T., & Tociu, R. L. (2012). Découverte dune
nouvelle grotte ornée paléolithique en Roumanie, dans le département du Bihor. In J. Clottes
(Ed.), LArt Pléistocéne dans le Monde (Bulletin de la Societé préhistorique de lAriège, 65-66)
(pp. 513528).
Cosac, M., Murătoreanu, G., Veres, D., Niță, L., Schmidt, C., Hambach, U., Alexandru, R.,
Cuculici, R., Buzea, D. L., Mărgărit, M., Dumitrașcu, V., Vasile, Ș., Petculescu, A., & Dénes,
I. (2018). Multi-proxy archaeological investigations of a middle Palaeolithic occupation context
in eastern Transylvania, Romania. Quaternary International, 485, 115130.
12 Hominin Footprints from Romania 209
De Lumley, H., Giacobini, G., Vicino, G., & Yokoyama, Y. (1984). New data concerning the dating
and interpretation of human footprints present in the grotta della Basuraat Toirano (Savona,
Northern Italy). Results of an international round table. Journal of Human Evolution, 13,
537540.
Harvati, K., & Roksandic, M. (2016). The human fossil record from Romania: Early upper
Paleolithic European mandibles and Neanderthal admixture. In K. Harvati & M. Roksandic
(Eds.), Paleoanthropology of the Balkans and Anatolia: Human evolution and its context
(pp. 5168). Dordrecht: Springer.
Higham, T., Douka, K., Wood, R., Bronk Ramsey, C., Brock, F., et al. (2014). The timing and
spatiotemporal patterning of Neanderthal disappearance. Nature, 512, 306309.
Iovita, R., Fitzsimmons, K. E., Dobos, A., Hambach, U., Hilgers, A., & Zander, A. (2012). Dealul
Guran: Evidence for lower Palaeolithic (MIS 11) occupation of the lower Danube loess steppe.
Antiquity, 86, 973989.
Mertens, S. (1996). The middle Paleolithic in Romania. Current Anthropology, 37(3), 515521.
Molleson, T. I., Oakley, K. P., & Vogel, J. C. (1972). The antiquity of human footprints of Tana
della Basura. Journal of Human Evolution, 1, 467471.
Olariu, A., Skog, G., Hellborg, R., Stenström, K., Faarinen, M., Persson, P., & Alexandrescu,
E. (2005). Dating of two Paleolithic human fossil bones from Romania by accelerator mass
spectrometry. In A. Olariu, K. Stenström, & R. Hellborg (Eds.), Applications of high-precision
atomic and nuclear methods (pp. 222226). Bucharest: Editura Academiei Române.
Onac, B. P. (2001). Mineralogical studies and uranium-series dating of speleothems from
Scărișoara Glacier cave (Bihor Mountains, Romania). Theoretical and Applied Karstology,
13-14,3338.
Onac, B. P., & Goran, C. (2019). Karst and caves of Romania: A brief outlook. In G. M. L. Ponta &
B. P. Onac (Eds.), Cave and karst systems of Romania (pp. 3740). Cham: Springer.
Onac, B. P., & Lauritzen, S. E. (1996). The climate of the last 150,000 years recorded in
speleothems: Preliminary results from North-Western Romania. Theoretical and Applied
Karstology, 9,91.
Onac, B. P., Viehmann, I., Lundberg, J., Lauritzen, S. E., Stringer, C., & Popiţă, V. (2005). U-Th
ages constraining the Neanderthal footprint at Vârtop cave, Romania. Quaternary Science
Review, 24(1011), 11511157.
Păunescu, A. (2001). Paleoliticul și Mezoliticul din spațiul Transilvan. Bucharest: Ed AGIR.
Rișcuția, C., & Rișcuția, I. (1970). Étude anthropologique des empreintes humaines du la grotte
Ciurului-Izbuc. In T. Orghidan (Ed.), Livre du centenaire Émile G. Racovitza
(pp. 529533). Editura Academiei RSR: Bucharest.
Rusu, T., Viehmann, I., Racoviță, G., & Crăciun, V. (1969). Primele urme de pași ale omului
preistoric din peșterile României (First footprints of prehistoric humans from Romanian caves).
Ocrotirea Naturii, 13, 191200.
Rusu, T., Racoviță, G., & Crăciun, V. (1970). Le système karstique Toplița - Ciur - Tinoasa (Les
Monts Pădurea Craiului). In T. Orghidan (Ed.), Livre du centenaire Émile G. Racovitza
(pp. 627650). Editura Academiei RSR: Bucharest.
Socaru, A., Doboş, A., & Trinkaus, E. (2007). Early modern humans from the Peștera Muierii,
Baia de Fier, Romania. PNAS, 48(103), 1719617201.
Staubwasser, M., Drăgușin, V., Onac, B. P., Assanov, S., Ersek, V., Hoffmann, D. L., & Veres,
D. (2018). Climate change impact on the transition of Neanderthals to modern humans in
Europe. PNAS, 115(37), 91169121. https://doi.org/10.1073/pnas.1808647115.
Trinkaus, E., Moldovan, O., Milota, S., Bîlgăr, A., Sarcină, L., Athreya, S., Bailey, S. E., Rodrigo,
R., Gherase, M., Higham, T., Ramsey, C. B., & Van Der Plicht, J. (2003). An early modern
human from the Peștera cu Oase, Romania. PNAS, 100(20), 1123111236.
Veres, D., Cosac, M., Schmidt, C., Murătoreanu, G., Hambach, U., Hubay, K., Wulf, S., &
Karátson, D. (2018). New chronological constraints for Middle Palaeolithic (MIS 6/5-3) cave
sequences in Eastern Transylvania, Romania. Quaternary International, 485,103114. https://
doi.org/10.1016/j.quaint.2017.07.015.
210 B. P. Onac et al.
Viehmann, I. (1975). Urme de viațăde acum 15000 de ani (Traces of life from 15000 years ago).
Radio Tele Școală,3334.
Viehmann, I. (1987). Prehistoric footprints in Romanias cave. Theoretical and Applied Karstology,
3, 229234.
Viehmann, I., Racoviță, G., & Rișcuția, C. (1970). Découvertes tracéologiques concernant la
présence de lhomme et de lours de caverne dans la grotte Ciurului-Izbuc des Monts Padurea
Craiului. In T. Orghidan (Ed.), Livre du centenaire Émile G. Racovitza(pp. 521527).
Editura Academiei RSR: Bucharest.
Webb, D., Robu, M., Moldovan, O., Constantin, S., Tomuș, B., & Neag, I. (2014). Ancient human
footprints in Ciur Izbuc cave, Romania. Amercian Journal of Physical Anthropology, 155,
128135.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapters Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapters Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Chapter 13
Episodes of Magdalenian Hunter-Gatherers
in the Upper Gallery of Tuc dAudoubert
(Ariège, France)
Andreas Pastoors, Tilman Lenssen-Erz, Tsamgao Ciqae, /Ui Kxunta,
Thui Thao, Robert Bégouën, and Thorsten Uthmeier
Abstract The Tuc dAudoubert cave (Ariège, France) offers unique insights into
the life of Late Pleistocene hunters-gatherers due to its exceptionally good preser-
vation conditions. This is especially true for the 300 footprints in the upper gallery of
the cave. Even for the layperson, some trackways are easily recognized. Short
episodes of past life become tangible. The spectrum of scientic analytic methods
used in western science has not yet provided an option to interpret these visible
episodes satisfactorily. For this reason, tracking experts, i.e. indigenous ichnologists,
were invited to analyse the footprints in Tuc dAudoubert. With their dynamic
approach of identication, they are able to do justice to the dynamics embodied in
the footprints. In total, eight main concentrations in four different locations were
studied. Two hundred fty-ve footprints were identied and grouped into 24 events.
In view of the group compositions and the assumption that humans did not climb
alone into the upper gallery for security reasons, it can be concluded that a maximum
of ve visits by two to six subjects were carried out. Among the events, the couple of
an adult man and an adult woman, who appear together in a total of ten different
spots, is particularly noteworthy. Altogether, this study is a rst step of a multi-stage
procedure. Further analyses based on measurements and plantar pressure analyses
will follow.
A. Pastoors (*) · T. Uthmeier
Institut für Ur- und Frühgeschichte Friedrich-Alexander-Universität Erlangen-Nürnberg,
Erlangen, Germany
e-mail: andreas.pastoors@fau.de
T. Lenssen-Erz
African Archaeology, University of Cologne, Cologne, Germany
e-mail: lenssen.erz@uni-koeln.de
T. Ciqae · /Ui Kxunta · T. Thao
Nyae Nyae Conservancy, Tsumkwe, Namibia
R. Bégouën
Association Louis Bégouën, Musée de Pujol, Montesquieu-Avantès, France
©The Author(s) 2021
A. Pastoors, T. Lenssen-Erz (eds.), Reading Prehistoric Human Tracks,
https://doi.org/10.1007/978-3-030-60406-6_13
211
Keywords Footprints · Upper Palaeolithic · Event identication
212 A. Pastoors et al.
Introduction
Over the past years, ichnology has acquired a new relevance in prehistoric archae-
ology of caves, as shown in a number of scientic studies (e.g. Ledoux 2019;
Romano et al. 2019; Ortega Martínez and Martín Merino 2019; Pastoors et al.
2017; Pastoors et al. 2015) and the International Conference on Prehistoric Human
Traces held in Germany (Cologne, May 2017). It is within this framework that the
prehistoric human tracks in Tuc dAudoubert are analysed in a multi-stage proce-
dure, combining static with dynamic approaches. In the rst phase the tracks have
been studied by indigenous ichnologists in 2018, and their results will be presented
in this contribution. As static analyses, i.e. Cussac, Fontanet, Bàsura Cave and Pech-
Merle of human footprints in caves have shown, this method is not appropriate for
exploring the entire information potential of human tracks (cf. Ledoux 2019;
Romano et al. 2019; Duday and García 1983). A dynamic method of reading
footprints in a morpho-classicatory way offers signicantly more possibilities.
The good preservation of most of the footprints in Tuc dAudoubert provides an
ideal framework for this investigation.
Quantitative, static analyses are not yet done but will in a next step serve as an
important complement and cross-check. In this way, a maximum of information can
be drawn from the prehistoric footprints of Tuc dAudoubert.
At this point, it is important to note that this contribution focuses exclusively on
the footprints which are not directly related to the making of drawings or clay
models, in the broader sense of art. This clear separation of the aforementioned
spoors in terms of activity and space makes such a distinction meaningful. The
results of the analysis of the spoors from the Salle des Talons are only included here
in particular cases as far as they are published.
The following chapter examines traces that document the locomotion in space
and the interaction between humans and bear bones in the various locations along the
upper gallery. But it is the intention to go beyond the reconstruction of the activities
of every subject. The focus is on the identication of events from the lives of the
individual subject as well as groups.
Design of the Project
For the study, three indigenous ichnologists were engaged who have already worked
in the Tracking in Caves project (Pastoors et al. 2015,2017; Lenssen-Erz et al. 2018)
but also as professional trackers for commercial hunting and, especially, as eco-
nomic support for their families and villages through traditional hunting practices.
Eight main concentrations of human tracks in the upper gallery of Tuc dAudoubert
13 Episodes of Magdalenian Hunter-Gatherers in Tuc dAudoubert 213
Entrance
Lower Gallery
Upper Gallery
Intermediate Gallery
Galerie de la Colonne
Galerie des Effondrements
Galerie des Empreintes
Galerie des Petits Pieds
Salle des Talons
Galerie des Bisons d‘Argile
Siphon 50 m
Paris
Bordeaux
Toulouse
Tuc d‘Audoubert
13.1). There, the
three ichnologists were asked to investigate the discernible footprints and other
traces, while the archaeologists accompanying them were assigned to document
Fig. 13.1 Simplied plan of Tuc dAudoubert with designation of the locations mentioned in the
text. (Illustration Association Louis Bégouën)
their analysis. The research in Tuc dAudoubert took place from 10 to 21 October
2018.
were selected according to a list of priorities for the quality and quantity of human
footprints in the following locations: Galerie des Effondrements, Galerie des
Empreintes, Galerie des Petits Pieds and Salle des Talons (Fig.
Participants
The main researchers of this project were three indigenous ichnologists from the
Nyae Nyae Conservancy around Tsumkwe (Namibia): Thui Thao, /Ui Kxunta and
Tsamgao Ciqae. The rst two of them are certied Master Trackers of the Cyber-
Tracker system (see www.cybertracker.org), while the third, having learned tracking
in a traditional way, has mainly helped to translate into English the analysis of the
other two ichnologists, which were in Ju|hoansi language. In addition, T. Ciqae also
holds a level 2 certicate as a tourist guide and is currently preparing a Namibian
hunting licence, so he is very familiar with species terminology (English and Latin).
214 A. Pastoors et al.
Materials
The Volp Caves
The three caves of the Volp, Enlène, Trois-Frères and Tuc dAudoubert, have
already been widely described in previous publications (cf. Bégouën et al. 2009,
2014,2019) and will be presented here only in short form.
The caves are located in the extension of each other under a limestone massif
mostly forested, covered with dolines and bizarre rocks with channel-like furrows
(southeastern France, Ariège). The limestone massif runs from east to west in this
northern Pyrenean part formed of parallel ranges between the Plantaurel in the north
and the Arize massif in the south. It is placed in the territory of the community of
Montesquieu-Avantès, 14 km southwest of Mas dAzil. The landscape is contrasted,
since the regular and undulating forms of the Cenomanian hills are brutally opposed
to the classical phenomena of karst. Under one of these hills, only a few kilometres
after its source, the Volp has carved out a large three-level hydrographic network.
The lower gallery is the one where the Volp ows, interspersed with two impassable
siphons, making the 875 m course impossible to navigate between its loss and its
resurgence. The intermediate gallery only exists in the downstream zone at 3 m
above the Volp bed. It is in the uppermost level that the upper gallery of Tuc
dAudoubert and the caves of Enlène and Trois-Frères are located.
The Cave of Tuc dAudoubert
The cave of Tuc dAudoubert is 640 m long with the resurgence of the Volp as its
entry, and because the Volp did not ow during certain periods of the late glacial
(Bégouën et al. 2009), this allowed humans easy access to the intermediate gallery
(Fig. 13.1). This gallery has preserved many archaeological ndings and parietal art,
remains of diverse prehistoric activities. A 12-m-high chimney leads from the Salle
Nuptiale to the upper gallery, which extends over 465 m. The course in this network,
sometimes very difcult, is closely marked for preservation reasons by two cords up
to the Bisons dArgile, a unique masterpiece of its kind. Throughout the route, traces
of the humanspassage are visible on either side of the trail: footprints and heels of
adults and children, ngerprints in the clay on the ground, broken bear skulls with
extracted teeth, jewellery objects placed on the ground, etc. Parietal art is present in
the entire intermediate gallery and in the rst part of the upper gallery.
Archaeological Context
From 1992 to 2009 a comprehensive research project was carried out in Tuc
dAudoubert (Bégouën et al. 2009). The aims of this 17-year project were to carry
out a broad prospection to evaluate the archaeological potential; to develop a
systematic investigation of divers nd categories, their documentation and analysis
(rock art, depositions, excavations, sondages and dating); and to publish an
encompassing monography (Bégouën et al. 2009).
13 Episodes of Magdalenian Hunter-Gatherers in Tuc dAudoubert 215
According to this publication, a total of 356 graphic elements were recorded,
101 of which show motifs from the animated world. Among these, depictions of
steppe bison (41%) clearly dominate over horse (16%). Reindeer, ibex, snake, lion,
bear and unreal beings complete the ensemble of these motifs. In addition, the 140 P-
and Q-shaped claviform signs stand out. Apart from these numbers, the multiple
depictions of bison couples (male and female) is exceptional. But, most spectacular
are the clay sculptures preserved in Tuc dAudoubert. They represent a male and a
female bison, each being 60 cm long and placed in the centre of the last chamber of
the upper gallery. On their surfaces human traces as marks of the production process
are well preserved (e.g. smoothing with hands and ngerprints on the mane).
Furthermore, technical details of the production are still visible: Horns and ears
were attached, eyes modelled as craters or elevations and beards cut with a
sharp tool.
The cave walls were not only used as canvas for drawings, but their niches and
ssures serve for the deposition of various artefacts. A total of 18 objects were found
in Tuc dAudoubert in such situations. Usually these are bone fragments, but lithic
artefacts, projectiles and red ochre were also found. The objects are wedged or ready
to hand. Only in rare cases they are hidden and difcult to nd.
Human presence in Tuc dAudoubert is evinced for autumn-winter, between
17,200 and 16,500 calBP. Only one single nd layer was found at each of the ve
limited excavations in different chambers. Remarkable is the diversity of the
reconstructed activities, their probable contemporaneity and relation to the cave
topography (Pastoors 2016).
Various reconstructed activities reect concrete movements in space and show
that the cave as a natural structure has been anthropogenized. This is important to
memorize for the analysis of the prehistoric footprints in the upper gallery.
In Tuc dAudoubert, 21 specicnd concentrations were identied at which, on
the one hand, substantial activities (N ¼2) and, on the other hand, limited, qualied
activities were carried out (Fig. 13.2). These limited, qualied activities include
drawing activities (N ¼16) and the consumption of introduced provisions (N ¼2).
All 21 nd concentrations are in the dark zone of the cave.
The selection of chambers for the various activities of prehistoric humans in Tuc
dAudoubert shows a clear pattern (Fig. 13.2). While substantial and consumption
activities were carried out in chambers that were wide and high, drawing activities
were carried out in the entire spectrum of chamber types used in Tuc dAudoubert. It
is noticeable, however, that concentrations with only drawing activities are located
in narrow or low chambers. From the picture emerge two chambers with substantial
or consumption activities in narrow, low chambers (Galerie du Bouquetin and
Diverticule des Dessins).
Find concentrations with substantial or consumption activities do not show any
pattern at rst sight due to their placement in the path network (Fig. 13.2). They are
216 A. Pastoors et al.
Entrance Diverticule des Dessins
Galerie du Bouquetin
Balcon II
Balcon I
Salle du Cheval Rouge
Diverticule des Claviformes
Passage des Monstres
Galerie des Bisons d‘Argile
Diverticule du Siphon
10
5
5
5
5
>100
50
<5
30
Salle des Talons
15
Siphon
?
substantial
dark
half-shade
daylight
consumption
unknown qualified
undocumented
drawing
narrow-high
medium-high
wide-high
narrow-low
medium-low
wide-low
human activity light zone chamber type
side passage
passageway
junction
crossing
dead-end
walking
crawling
climbing
people (n)
path network mode of movement available space
>100
50 m
Fig. 13.2 Distribution of concentrations of prehistoric remains with structured information on the use of space in Tuc dAudoubert. (Illustration Association
Louis Bégouën)
located at junctions, at a side passage, at passageway and in a dead end. The
differentiated consideration of the two types of activities shows that at least the
concentrations with substantial activities are in a strategically favourable position in
Tuc dAudoubert path network due to their immediate proximity to the central trafc
axis of the lower gallery. The concentrations with drawing activities, on the other
hand, show a clear relation to certain components of the path network. In particular,
dead ends and side passages were selected. It is interesting to note that junctions
were selected for drawing activities when also other activities were carried out there.
Passageways seem to have been of little interest.
13 Episodes of Magdalenian Hunter-Gatherers in Tuc dAudoubert 217
The concentrations with substantial or consumption activities are comfortably
accessible upright (Fig. 13.2 mode of movement). This also applies to the majority
of concentrations of drawing activities. In addition, two concentrations can only be
passed crawling. Another two concentrations have to be climbed. A total of three
concentrations require combined modes of movement: walking and crawling or
climbing and crawling.
For the substantial and consumption activities, premises were selected that offer
sufcient space for several people at the same time (Fig. 13.2 chamber type). Small
chambers were avoided for these activities. Exactly the opposite is true for the
drawing activities. Here, space was selected that could accommodate a maximum
of ve people at the same time.
The spatial distribution of the depots corresponds very well with that of the
concentrations with substantial or consumption activities. Here a direct relationship
between the different activities seems to be evident. The only exception is a fragment
of bone deposited at the branch of the Diverticule des Claviformes diverting from the
Galerie du Bouquetin in a niche 6 m above the ground in a shaft leading upwards.
The analyses of the archaeological nds exhibit a short stay in Tuc dAudoubert
with different activities of basic supplies, consumption, raw material extraction and
drawing activities. In the course of this stay, the entire cave was explored with
sporadic visits to the upper gallery. This large spectrum of qualied activities in
connection with substantial activities is similar to base camp activities in open-air
sites. Thus Tuc dAudoubert plays a comparable role within the network of sites of
Magdalenian hunter-gatherers in the Pyrenees for a limited period of time and
represents in this respect an autonomous subsystem.
The inferences from this detailed picture for the basic understanding of the
episodes fossilized in the oor of the upper gallery are the following:
Base camp activities suggest the presence of the entire group of hunter and
gatherers with members from each age class.
The anthropization of the intermediate gallery of the cave testies to a behaviour
based in experience with the conditions of a complex cave system.
218 A. Pastoors et al.
Human Tracks
Tracks of humans and cave bears were noticed and respected right from the rst day
of the discovery of the upper gallery (10 October 1912). This is a very thoughtful
behaviour for that time and the basis for the preservation of all tracks into the twenty-
rst century. For the monograph of 2009, a rst tentative count of the human tracks
was carried out (Table 13.1).
Of the total of 302 human footprints, passage-related traces are by far the most
abundant, and among them, the heels are mainly grouped in the Salle des Talons.
Apart from the latter, whose count corresponds to all that is visible in this place, the
87 feet inventoried elsewhere represent only a sample. To preserve the soil, the
distant identication of human tracks in the vicinity of bear tracks has proven to be
difcult, sometimes impossible (human presence always after that of the cave bear).
Moreover, the entire gallery could not be prospected because the virgin surfaces
were too fragile. The actual number of footprints must be signicantly higher.
Moving towards the deep part of the cave, the rst footprints appear in the Salle
des Lacis. They can be related to the last engravings when coming from the entrance
but also to the rst displaced bear bones and accumulated concretions. This associ-
ation of footprints and manipulated objects, moved or broken, becomes a constant
phenomenon in the deep part of the upper gallery. However, two categories can be
distinguished: on the one hand, footprints reecting dynamic movement and, on the
other hand, concentrations of imprints over small areas, indicating a stopover or
short-distance comings and goings. The former are related to the progression of
humans in the gallery and the latter to activities requiring a stopover. The activities
during the stopover were sparse because there is no intense trampling as the
footprints are clearly discernible and overlaps are infrequent. Thus in the Galerie
des Effondrements, about 20 footprints, ngerprints and broken concretions encircle
the mandible of a cave bear. Further on, about 40 footprints are spread over 30 m in
four concentrations: about 20 in the rst, then 8 around the broken cave bear skull,
19 at least in an area with scattered manipulated cave bear bones and nally some at
Table 13.1 Number of prehistoric human tracks in the upper gallery of Tuc dAudoubert (Bégouën
et al. 2009)
Location Footprint Heelprint Div. spoors Total
Salle des Lacis 1 1
Galerie du 10 Octobre 1 1
Galerie de la Colonne 1 1 2
Galerie des Effondrements 11 6 8 25
Galerie des Empreintes 72 3 1 76
Galerie des Petits Pieds 4 1 5
Galerie des Bisons dArgile 4 4
Salle des Talons 183 5 188
Total 87 196 19 302
13 Episodes of Magdalenian Hunter-Gatherers in Tuc dAudoubert 219
the end of the gallery. Five footprints are concentrated at a prominent location in the
Galerie des Petits Pieds.
The next important concentration is none other than the Salle des Talons
(Fig. 13.1). Since 1912, these prints have raised many questions, most of which
remain unanswered. We have seen that the asymmetrical shape of the cups and their
level of sinking into the ground correspond well to heels. Rare circular cups
represent knee imprints. It was until more than 100 years after the discovery that
complete footprints were discovered for the rst time by indigenous ichnologists
(Pastoors et al. 2015) in 2013. The distribution of heel imprints indicates activities
around the extraction of clay and the making of various drawings on the cave oor
(Bégouën et al. 2009; Pastoors 2016).
In summary, for the Salle des Talons initially there were the assumptions that
young individuals have left behind ve to six sequences of tracks. According to
Bégouën, ritual dance or initiation (Bégouën 1928) was the motivation for this.
Vallois is much more neutral and sees here young individuals, deliberately walking
on heels (Vallois 1931). A further interpretation of the events in the Salle des Talons
that led to the distribution of the footprints was carried out by the indigenous
ichnologists in 2013 and 2018. They identied two subjects who crossed the
chamber twice to a clay extraction pit (Pastoors et al. 2015). In addition, further
footprints are associated with drawing activities on the oor.
Investigations about the identity of the trackmakers in the upper gallery of Tuc
dAudoubert were carried out only unsystematically up to the present work. Vallois
examined a selection of the best-preserved footprints and took the rst measure-
ments. Two complete footprints measure 218 mm or 200 mm in length and 53 mm or
62 mm in heel width. Further dimensions were taken from heel prints, which
accumulate at various points in the cave. Accordingly, the examined heels have a
maximum width of 72, 68, 67, 60, 54 and 52 mm (Vallois 1931). The step width of
these heel imprints is between 25 and 28 cm. In the Salle des Talons, also measures
of the maximum width of the heels were taken. Thus they are 58, 55, 53, 52 and
50 mm wide. The step width of these prints examined is a maximum of 20 to 25 cm.
Methods
Prehistoric human traces are considered to be the most personal, nonmaterial
legacies that have remained. These are mainly footprints, but also traces of hands,
knees and other body parts. Curiously, it does not yet seem possible to do justice to
these information-rich traces with synthetic classication and quantitative methods.
A critical inspection of the possibilities and above all the limits of current methods
clearly shows that on empirical basis only the number of different trackmakers can
be calculated (Bennett and Morse 2014; see Chap. 2). In the ideal case, statements
about the gait and the walking speed are also possible. On the basis of quantitative
analyses, it is currently not possible to say anything dependable about the identity of
people and the episodes stored in the tracks. It looks as if these static analyses are not
appropriate for exploiting the information potential of this multifaceted nd category
of dynamic processes (see Chap. 6).
To pursue this issue more closely, the methodological process for the analysis of
prehistoric footprints in Tuc dAudoubert follows a multistage procedure. This
includes the identication of the traces left behind by prehistoric humans according
to the principle of the preiconographic description by Panofsky (see next paragraph;
Panofsky 1962).Traces are recognized, put in relation to each other and summarized
as events. In a further step, the identied human tracks are analysed quantitatively
following basic measurements (cf. Bennett and Morse 2014). Footprint outline- and
landmark-based geometric-morphometric analysis (cf. Bennett et al. 2009; Bennett
et al. 2016) and pixel-based quantitative analysis of the whole plantar pressure
(cf. Crompton et al. 2011)are also planned.
Practical experience (familiarity with objects and phenomena) is an absolute
prerequisite for a successful application of the preiconographic description, from
which a positive correlation between experience and descriptive accuracy can be
derived. In the case that the spectrum of personal experience is not sufcient, this
spectrum must be extended by consulting publications or experts. Practical experi-
ence, in turn, helps to determine which publication or expert is to be consulted
(Panofsky 1962: 9). In prehistoric archaeology, it is a common practice to compen-
sate the lack of practical experience with experiments (e.g. Bourguignon et al. 2001).
In the layout of the current research project we decided against the generation of
experience through experimental archaeology. Instead, we use expert knowledge of
indigenous ichnologists building on their outstanding experience in reading tracks
(Liebenberg 1990; Gagnol 2013; see also Chap. 6and 19).
The process of recording the workow of the indigenous ichnologists in reading
prehistoric human spoors has been substantially further developed compared to the
one applied in 2013. First of all, lists were compiled with information on each
individual footprint examined. The following aspects were documented:
220 A. Pastoors et al.
Subject number: The subject number identies each individual (trackmaker)
independently of the study area within the cave. This makes it easy to follow
each subject through the cave.
Age: The results of the morpho-classicatorical analysis of age are given very
precisely by the indigenous ichnologists. In consideration of the fact that such a
precise age indication by means of footprints seems problematic and should
always be seen against the background of the reference collection used or
personal experience, the data of the indigenous ichnologists are grouped together
in age classes according to Martin (Martin 1928)neonatus, infans I
(0.56 years), infans II (713 years), juvenis (1420 years), adultus
(2140 years), maturus (4160 years) and senilis (>60 years).
Sex: If the sex of the subject can be identied, it is recorded as female or male.
Physique: Under this aspect, information about the body shape is given. Here,
too, it is more a matter of deviations from a normal physique than of a precise
denition of a certain shape.
13 Episodes of Magdalenian Hunter-Gatherers in Tuc dAudoubert 221
Handicap: Under handicap, observations are recorded that relate to deviations
from a well-balanced human being. No statements are made about the medical
causes.
Spoor number: The spoor number designates each individual human trace exam-
ined and listed in the project. Subject and spoor number together form a distinc-
tive unit. They are continuous and thus allow an unambiguous assignment of the
human traces in each part of the cave.
Spoor type: Species the exact body part that caused the traces. This includes the
foot, hand, knee, elbow and others (e.g. tools).
Side: If the side of the body part can be identied, it is recorded as left or right.
Additional weight: The additional weight refers to the characteristics of a subject
that deviate from the normal gait or depth of imprint.
Gait: Under this point, statements are made about the manner of the executed
locomotion. This includes safety and speed, as well as movement in a group or
alone.
Direction: The direction of movement is documented in cardinal direction.
Specic local information is given for better orientation in the cave.
Trackway: Hereunder it is noted whether the footprint is part of a series of
footprints of the same subject or whether it is isolated.
Event identication: Summary of traces of individual or several subjects in
temporal, spatial and content-related connection with each other.
Taphonomy: This aspect refers to the state of preservation of the various traces
which can be inuenced by both natural and anthropogenic factors.
Substrate: The substrate refers to the sediment in which the spoor was formed.
Reliability of identication: Particularly important for the comprehensibility of
the analysis is the judgement of its reliability on the basis of preservation and
visibility. For this purpose, a subjective ve-stage classication was carried out
from very good (1) to unsatisfactory (5). The intermediate stages are good (2),
satisfactory (3) and sufcient (4).
Remarks: An open eld for comments of any kind.
The position of every spoor was located on plans or sketches. All work sequences
were recorded on lm. In this way, not only the results can be checked and compared
with each other, but also further linguistic research can be carried out. At the end
stands a database (catalogue) with the results of the morpho-classicatorical analysis
and event identication. For future work, photogrammetric records of the examined
footprints will be generated with the help of Structure from Motion (e.g. Mallison
and Wings 2014).
In order to understand how a combination of footprints is identied as a track, and
how several tracks sometimes are being interpreted as a coherent event, it is helpful
to look at perception psychology and Gestalt principles in particular. By Gestalt is
meant:
a unitary whole of varying degrees of detail, which, by virtue of its intrinsic articulation and
structure, possesses coherence and consolidation and thus detaches itself as a closed unit
from the surrounding eld. (Maynard 2005:501 citing Gurwitsch 1964).
222 A. Pastoors et al.
The concept of Gestalt was introduced by Max Wertheimer (Wertheimer 1923),
and since then research into Gestalt formation focuses on the perception and
interpretation of grouped objects as well as on small entities within larger environ-
ments and is of relevance still today (Wagemans et al. 2012a,b). So-called Gestalt
laws (Fitzek and Salber 1996) or principles are particularly vital in the advertisement
industry (e.g. Graham 2008), and, besides psychology (e.g. Wörgötter et al. 2004),
they have also received quite some attention in computer science and mathematical
approaches (e.g. Zhu 1999; Elder and Goldberg 2002; Wen et al. 2010). Some of the
Gestalt principles are gure-ground articulation, proximity, common fate, similarity,
continuity, closure, past experience and good Gestalt (Todorovic 2008). All these
principles are at work in perception when regarding spoor, single or in trails, and
making sense of their complex and combining information.
Results
In the following part, the results of the identications of the prehistoric footprints
from Tuc dAudoubert by the indigenous ichnologists are presented in spatial units,
advancing into the depth of the upper gallery. Starting point in each section is the
specication of the chamber with its prominent nds and features, which are based
on the descriptions by Bégouën et al. (2009). After this intro, the results are grouped
according to the events identied. In this chapter, two different systems are used to
identify the individual spoors. On the one hand, the numbering of the spoors as
published by Bégouën et al. (2009) (e.g. TUC-291) is used as a reference while on
the other hand, since it is more detailed, the project-internal numbers of the tracks
(e.g. S81, S82...) (for cross-references, see Table 13.2). The rating of the
reliability of identication is assembled in the same table.
Galerie des Effondrements
This gallery is about 50 m long and comprises a passage between various geological
phenomena that have marked this place (Fig. 13.3). Prehistoric humans followed this
itinerary, leaving their traces throughout this same passage. The oor of the Galerie
des Effondrements is largely made up of stalagmitic oors, especially on the
southern side of the path. On the northern side there are clay areas with various
human spoors. Apart from these traces, the most spectacular testimonies are the bear
bones removed from their original deposits and placed along the path. After a large
chute, the gallery widens but remains marked by bear bone deposits, still located in
the axis of the passage.
Just after a stalagmitic obstacle, the path makes a sharp turn to the right. On its left
side, at 60 cm from the passage, a human heel (TUC-266) is visible with its well-
marked clay ridge. Not far from the previous one, over a length of about 1 m, there
(continued)
13 Episodes of Magdalenian Hunter-Gatherers in Tuc dAudoubert 223
Table 13.2 List of prehistoric tracks identied by the indigenous ichnologists in the upper gallery
of Tuc dAudoubert
Spoor
n
Reliability of identication
Cross-
reference Location
Toes
Ball of
foot Midfoot Heel Event
S31 1 2 2 2 E24 TUC-331 Galerie des Petits Pieds
S32 3 4 4 4 E15 TUC-308 Galerie des Empreintes
eastern centre
S33 5 5 5 3 E15
S41 1 1 1 2 E24 TUC-332 Galerie des Petits Pieds
S42 5 4 4 3 E21 TUC-330
S51 3 4 4 3 E22
S61 2 3 3 3 E23
S71 5 3 3 1 E17 TUC-308 Galerie des Empreintes
eastern centre
S72 5 5 4 4 E17
S73 5 5 4 3 E16
S74 2 3 3 3 E16
S75 1 3 3 2 E17
S76 4 4 4 4 E17
S77 5 5 5 2 E13 TUC-303
S78 5 5 5 2 E13
S79 5 5 2 2 E13
S710 5 4 2 2 E13
S711 4 4 3 2 E13
S712 4 4 4 3 E14
S713 1 3 5 5 E11 TUC-293 Galerie des Empreintes
western centre
S714 5 5 4 4 E11
S715 3 3 3 3 E10
S716 5 5 3 2 E10
S717 4 4 5 5 E10
S718 4 4 4 4 E8 TUC-291 Galerie des Empreintes
western end section
S719 5 5 5 4 E8
S720 2 3 5 5 E8
S721 4 4 3 3 E8
S722 4 4 4 3 E7
S723 5 5 4 3 E7
S724 4 4 3 3 E7
S725 1 1 1 1 E2 TUC-273 Galerie des Effondrements
S726 1 1 1 1 E2
S727 5 5 1 1 E2
S728 1 1 1 1 E2
S729 1 1 1 1 E2
S730 3 2 1 1 E4
S731 5 5 2 1 E1 TUC-267 Galerie des Effondrements
S732 buttock (2) E1
nEvent reference Location
Toes Midfoot Heel
(continued)
224 A. Pastoors et al.
Table 13.2 (continued)
Spoor
Reliability of identication
Cross-
Ball of
foot
S733 5 4 3 3 E18 TUC-324 Galerie des Empreintes
eastern end section
S734 4 4 4 4 E18
S735 1 1 4 3 E19
S736 4 3 4 2 E19
S737 2 4 2 4 E19
S81 4 3 2 1 E17 TUC-308 Galerie des Empreintes
eastern centre
S82 5 5 4 2 E17
S83 2 4 4 3 E16
S84 2 3 4 5 E16
S85 1 3 4 5 E16
S86 1 2 5 5 E11 TUC-293 Galerie des Empreintes
western centre
S87 5 4 4 1 E10
S88 4 4 4 1 E10
S89 5 4 4 2 E10
S810 4 4 4 3 E10
S811 4 4 4 2 E9 Galerie des Empreintes
between western centre and
western end section
S812 4 4 3 1 E9
S813 5 3 2 2 E8 TUC-291 Galerie des Empreintes
western end section
S814 5 3 2 2 E8
S815 2 2 1 1 E8
S816 2 3 3 1 E8
S817 2 4 4 3 E8
S818 1 2 4 4 E8
S819 1 3 4 4 E8
S820 1 4 4 4 E8
S821 5 5 2 1 E7
S822 5 5 4 2 E7
S823 5 3 2 1 E3 Galerie des Effondrements
S824 1 1 2 2 E4 TUC-285
S825 5 3 1 1 E1 TUC-266 Galerie des Effondrements
S826 4 3 3 2 E18 TUC-324 Galerie des Empreintes
eastern end section
S827 4 4 4 4 E18
S828 4 4 4 3 E18
S829 4 4 4 3 E18
S830 5 5 2 2 E19
S831 1 2 3 3 E19
S91 3 3 3 2 E20
S101 3 3 3 3 E15 TUC-308 Galerie des Empreintes
eastern centre
S111 4 4 4 2 E15
S121 5 5 5 3 E15
S131 3 5 5 5 E15
nToes
3
13 Episodes of Magdalenian Hunter-Gatherers in Tuc dAudoubert 225
Table 13.2 (continued)
Spoor
Reliability of identication
Cross-
reference Location
Ball of
foot Midfoot Heel Event
S141 2 3 4 4 E12 TUC-293 Galerie des Empreintes
western centre
S142 2 3 4 4 E5 TUC-291 Galerie des Empreintes
western end section
S143 3 2 2 E6
Table 13.3 Quantication of tracks identied during the Tracking in Caves project in 2018; the
published data refer to Bégouën et al. (2009). Tracks from Salle des Talons were not analysed
equally detailed as all other tracks
Location Number of footprints
Published Identied in 2018 Proportion of published footprints
Galerie des Effondrements 25 11 44%
Galerie des Empreintes 76 67 88%
Galerie des Petits Pieds 5 5 100%
Salle des Talons 188 172 91.5%
Total 294 255 86.1%
are at least one footprint and a slide (TUC-267), on a relief near a depression.
According to the indigenous ichnologists, the traces described above came from
one single event.
The most visible human footprints are on the northern side of the passage, two
small heels of probably identical dimensions and appearance (TUC-280 and
TUC-281) heading to the east, facing the deep part of the cave. Another footprint
(TUC-285) is 1.5 m from the path, close to a natural crack in the clay, perpendicular
to the axis of the gallery. It is a right foot well printed in clay with ve clearly visible
toes. The most prominent area with human spoors is 3 m to the north from the path,
in the largest part of the gallery (TUC-273). Their presences indicate human
activities over an area of 6 m
2
. Apart from footprints, the edges of a depression
have retained two parallel and aligned nger marks, one of them near to a cave bear
mandible without its canine. In this area there are about 15 well-preserved footprints.
So far, 25 human tracks have been published of the Galerie des Effondrements
(Bégouën et al. 2009). In the course of the investigations by the indigenous
ichnologists, two further footprints were discovered, so that now overall 27 footprints
are known. Of these, only 11 were interpreted more closely by the trackers (44%)
(Table 13.3). The other footprints were either hidden or there was nothing reliable to
report about them. The 11 footprints were made by 2 adults, 1 female (subject S8)
and 1 male (subject S7), and derived from 4 events:
226 A. Pastoors et al.
Event 1: Just in the sharp turn, two subjects, subject S7, male adult, and subject
S8, female adult, walked together fast in direction to the entrance (Fig. 13.4).
From subject S7, male adult, is a left footprint that results from slipping (S731)
and led to a curiosity. The trackmaker couldnt keep his balance and has sat down
on his buttocks (S732) right on the edge of the depression mentioned above.
Event 2: The second event happened in the most prominent area with human
spoors in the Galerie des Effondrements (Fig. 13.3). Here subject S7, male adult,
has left several spoors in a sequence of ve successive footprints left (S725),
right (S727), left (S726), right (S729) and left (S728). The subject was
standing there, picking something up, probably the mandible of the cave bear
that is in front of the footprints (Fig. 13.4). While working there with the body
aligned to the northern wall of the gallery, the trackmaker was alone at the place.
No footprints of other subjects are visible in this restricted area.
Event 3: Just around 3 metres from the rst event, there is a new isolated left
footprint from subject S8 (S823), female adult (Fig. 13.3). It is on the north side,
about 1.5 m from the path. The subject was walking in the direction of the deep
part of the cave. At this point, the subject slipped a little and walked with slow
speed.
Event 4: The last event identied by the indigenous ichnologists is located near
the natural crack in the clay perpendicular to the axis of the gallery (Fig. 13.3).
Here, on a slightly rising ground, subjects S7, male adult, and S8, female adult,
walked fast together to the entrance of the cave. Both trackmakers were carrying
little additional weight at that place. From subject S8, female adult, a right
isolated footprint has been identied (S824). The right isolated footprint
(S730) from subject S7, male adult, was hitherto unknown.
The footprints in the Galerie des Effondrements document very well a short-term
activity of a male adult (subject S7) in the environs of the mandible of the cave bear,
a dynamic locomotion of a female adult (subject S8) in direction to the deep part of
the cave and a dynamic and fast locomotion of the two adults (subjects S7 and S8)
together carrying each a little additional weight back to the entrance of the cave.
Galerie des Empreintes
On a wide and not very calcinated surface, contrasting in this respect with the
previous gallery, the Galerie des Empreintes measures nearly 60 m long and 7 to
8 m wide and high (Fig. 13.3). Coming from the Galerie des Effondrements, the
entrance to the gallery is marked by an impressive stalagmite cascade. Shortly
afterwards, the eye immediately catches the long marked path that follows the central
axis of the gallery to its right. The left part is made up of a vast clayey expanse
entirely covered by bear tracks. There is also evidence of human activities, but the
fragility of the soil has not allowed a full exploration of this area. The omnipresence
of the bear is evident throughout the entire path of the Galerie des Empreintes. Its
13 Episodes of Magdalenian Hunter-Gatherers in Tuc dAudoubert 227
50 m
Galerie de la Colonne
Galerie des Effondrements
Galerie des Empreintes
Galerie des Petits Pieds
Galerie des Bisons d‘Argile
Salle des Talons
2
9
5-6
7
10
11 13
15
18
21-24
12
14 16
17 19 20
8
4
3
1
1
subject S7, male adult
subject S8, female adult
subjects S7 & S8
other subjects
event number
Fig. 13.3 Localization of the events in the upper gallery of Tuc dAudoubert. (Illustration Association Louis Bégouën/Tracking in Caves)
228 A. Pastoors et al.
Fig. 13.4 Complete events 1 and 2 and excerpt of event 8 in the upper gallery of Tuc dAudoubert
with the respective spoor number. (Photo Association Louis Bégouën/Tracking in Caves) the red
laser points to a part of the buttock imprint (S732) of event 1, whereas the green laser points to the
slip track (S813) in event 8
scattered bones are visible all over, and its tracks, slips and traces of hair and claws in
the clay and on the walls, broken concretions, make the presence of the bear almost
tangible.
A total of 76 human footprints were recorded during an initial counting. These
tracks are mostly concentrated in a total of four well-dened sections (Fig. 13.3). At
rst, the right part of the path runs along a low ceiling (1.4 m) under which footprints
are visible (western end section) in spatial relation to a drawing made by ngers on
the oor. After about 20 m, the gallery widens to the right into a semicircular room.
Here and a few metres before it, the soil has kept traces of passages and intense
prehistoric human activities (TUC-293 to TUC-327) (western and eastern centres)
(Fig. 13.3).
In the second part of the Galerie des Empreintes, the gallery then becomes slightly
open where another concentration of prehistoric human activities is visible (eastern
end section) (Fig. 13.3). A few metres further on, before a narrowing of the space
between barriers of concretions, the right wall marks its end. This narrow place has
been chosen to deposit three perforated teeth and red ochre on the oor, right against
the wall.
Western End Section
Coming from the Galerie des Effondrements, on the right side, under the lower roof,
21 footprints printed in the loamy soil were counted over a length of 3 m, some of
them later calcined (TUC-291). Nineteen of them were interpreted by the indigenous
ichnologists. The most complete, a right foot, is located very close to the path
(Fig. 13.3).
In the western end section, three subjects were identied. These are the same two
subjects (S7 and S8), who were already identied in the Galerie des Effondrements
and were underway together. There are four trackways with up to eight footprints of
this couple. Furthermore, a third subject (subject S14) left two isolated footprints in
the western end section. According to the observations of the experienced trackers,
subject S14 was solo on this spot. The western end section is a passage zone along a
low ceiling with various blocks and stalagmites on the oor. The passage was used
for the way into the deep part of the cave as well as to the entrance. In total of four
events can be summarized:
13 Episodes of Magdalenian Hunter-Gatherers in Tuc dAudoubert 229
Event 5: The isolated left footprint (S142) of subject S14, female infans II,
describes the rst event within this section of the Galerie des Empreintes. With a
fast speed the trackmaker moved to the deep part of the cave, lost her grip and
slipped with the toes against a rock which probably caused some pain.
Event 6: From the same subject S14, female infans II, a second footprint is from
her right foot (S143). Again with fast speed, she moved this time towards the
entrance (Fig. 13.3).
Event 7: The next event happened in a corridor with a low roof close to the right
wall. Here the two subjects S7, male adult, and S8, female adult, walked fast
230 A. Pastoors et al.
together in direction to the deep part of the cave (Fig. 13.3). From this event is left
arst trackway of subject S7 that is composed of three footprints right (S722),
left (S723) and right (S724). In a shorter trackway with only two footprints, the
way of subject S8 left (S821) and right (S822) is documented. Since the
roof is very low in this part of the cave, subject S8 moving through the lowest
passage had to walk bent over.
Event 8: The last event in this section of the Galerie des Empreintes took place
close to the actual path in the central axis of the gallery. Again, the two subjects
S7, male adult, and S8, female adult, walked fast together, each carrying little
additional weight in direction to the entrance (Fig. 13.3). Subject S7 is
documented by a trackway of four footprints right (S718), left (S719), right
(S720) and right (S721). From subject S8, female adult, is the longest trackway
known in the upper gallery left (S813), left (S814), right (S815), left
(S816), right (S817), left (S818), left (S819) and left (S820). Some tracks
are missing due to the changing substrate conditions. Close to a stalagmite that
disturbs the direct passage of subject S8, an interesting incident took place
(Fig. 13.4). With her left foot (S813), subject S8 lost her grip and slipped. But
it did not end in a fall because she found the balance by an interruption of her
forward movement, regaining a rm stand again (S814) and (S815) by
putting both feet side by side. Quite rare in Tuc dAudoubert are identications
of overlapping footprints. A very good example is provided within the described
event 8. Footprint S718 was clearly overstepped by S815 and S719 by S819
(Fig. 13.4). This proves that subject S7 went in front of subject S8 at this point of
the cave when walking back to the cave entrance.
Between western end section and western centre just close to the nger drawing
(Bégouën et al. 2009: 262), no tracks were left by the artist. The only identiable
footprints come from subject S8, female adult who has passed this section. This short
event is evinced by a short trackway with two footprints.
Event 9: Subject S8, female adult, left two footprints right (S811) and left
(S812), which lead to the entrance (Fig. 13.3). She was solo at this point and
passed fast this section close to the actual path.
Western Centre
Near the path, still on the right, about 15 footprints remain around a small prehistoric
excavation (TUC-293), 11 of them interpreted by the ichnologists. On the left side at
this point of the path, on the previously mentioned trampled slope, three barely
visible footprints seem to descend towards the path (TUC-294 - TUC-296). They are
too far from where one can regard them without damaging the substrate to identify
any details.
The western centre, according to the footprints, is a passage zone that three
subjects have passed. The path leads over a limestone block on the ground and
past a second one. In three events the same subjects appear as already met in the
western end section (S7, S8 and S14). They are represented here with three track-
ways with up to four footprints and two isolated footprints. The identied walking
directions lead, in both directions, to the depth of the cave as well as to the entrance.
13 Episodes of Magdalenian Hunter-Gatherers in Tuc dAudoubert 231
Event 10: In the rst and most complex event in this section, the two subjects S7,
male adult, and S8, female adult, walked fast together one after another in
direction to the entrance, both carrying something (Fig. 13.3). From subject S7
three footprints have been identied left (S715), right (S716) and left
(S717). The trackway with four footprints of subject S8 is longer left
(S87), right (S88), left (S89) and right (S810). It seems that the trackmaker
has probably supported herself in the vicinity of the footprints S89 and S810
with her left hand on the rock jutting into the passage. On the basis of
superposition S87 and S88 were overstepped by S715 it can be concluded
that subject S8 was the rst to pass this spot.
Event 11: Beyond the limestone block crossed by both, traces of the subjects S7
and S8 can be found again, this time pointing in the other direction (Fig. 13.3).
The two went together almost in the direction to the deep part of the cave. Subject
S7, male adult, has left a short trackway of two footprints right (S713) and left
(S714). This time subject S8 has left only an isolated left footprint (S86).
Event 12: The third event in this section happened in the same area as that of
event 10, but this time with subject S14, female infans II, that has left only an
isolated right footprint (S141) (Fig. 13.3). She was moving fast slightly slipping,
in the direction to the deep part of the cave.
Eastern Centre
On the right side of the small room, the at oor has abundant animal tracks,
including very large claws. Over a distance of about 10 m, human activity focused
on collecting and handling bear bones that would stick out from the clay soil. On the
natural anvil formed by a nascent stalagmite, a skull of a bear was smashed with the
probable purpose of extracting the teeth, none of which remain nearby (TUC-302).
Eight footprints (TUC-303) are printed in the clay to the left-hand side of the skull.
The face broke into fragments scattered all around the skull. A prehistoric excavation
located 1 m further to the left (TUC-305) can reasonably be considered as the
extraction site of the skull. On a strip 1.5 m wide, along the path, at least 19 footprints
mark the bottom of a slight depression (TUC-308), all covered with calcite. A little
further on to the deep part of the cave, 50 cm from the path, scattered on the ground,
there is a coxal bear bone, a bear rib and a complete left human footprint (TUC-318).
On the rib, there are clear traces of the brown clay crust that coated it before it was
extracted. Twenty-three out of the mentioned 28 footprints have been identied by
the indigenous ichnologists.
In the Galerie des Empreintes, the eastern centre represents the main activity area
in which seven subjects left their footprints. It seems that there was the couple
subject S7 and S8 again relocating bear bones, but also another group of subjects
(S3, S10, S11, S12 and S13) on their way. In total four trackways with up to ve
footprints and ten isolated footprints have been identied that constitute altogether
six events.
232 A. Pastoors et al.
Event 13: The following sequence of footprints is certainly one of the most
spectacular events (Fig. 13.3). These are ve consecutive footprints of subject
S7, male adult. The sequence begins with an isolated right footprint (S77). The
following two belong together and indicate a squatting position left (S78) and
right (S79) (Fig. 13.5). The same applies to the following two footprints: right
(S710) and left (S711). In this posture, an activity was performed close to the
oor, turned in direction of the entrance. Since the skull of a cave bear described
above is located directly in front of the footprints, a direct connection is most
likely. At this point subject S7 acted alone.
Event 14: Just behind the described event an isolated left footprint of subject S7,
male adult (S712), is found (Fig. 13.3). The path leads in direction to the cave
wall. Subject S7 was at this point alone.
Event 15: Several metres deeper in the cave, a group of subjects (S3, S10, S11,
S12 and S13) were identied that walked together at that point. The picture left by
the footprints is not to be interpreted as clearly as it was the case in other events.
The footprints point in different directions and are most likely to be understood as
walking around the gallery (Fig. 13.3). Subject S3, male infans I, is represented
by two isolated footprints. The rst is a right isolated footprint (S32). With fast
speed he went to the deep part of the cave. The second isolated footprint of
subject S3 derives again from a right foot (S33). It also shows a fast walking
speed towards the deep part of the cave. In the same area, a left footprint (S101)
from subject S10, female infans I, is leading into the direction of the deep part of
the cave. Furthermore, a left footprint (S111) comes from subject S11, female
adult, who walked fast in direction to the entrance and carried something. She
stepped over two footprints of the subject S7 (S71 and S76) (Fig. 13.5). Apart
from this, a right footprint (S121) comes from subject S12, male juvenis, who
walked in direction to the wall of the gallery. The last isolated track in this event
comes from subject S13, male infans I, and represents a non-specic footprint
(S131) pointing towards the deep part of the cave.
Event 16: In the same area in which event 15 happened, the two subjects S7, male
adult, and S8, female adult, walked fast together in direction to the entrance
(Fig. 13.3). Subject S7 is present with a trackway that consists of only two
footprints left (S7-3) and right (S7-4). The best visible and even recognizable
for a layperson is the trackway of subject S8, female adult, consisting of three
footprints right (S83), left (S84) and right (S85) (Fig. 13.5).
Event 17: The last event identied in this section of the Galerie des Empreintes
happened again with the two subjects S7, male adult, and S8, female adult
(Fig. 13.3). This time they walked fast together towards the deep part of the
cave after subject S7 had picked up probably some cave bear bones. This
particular trackway from subject S7 consists of three footprints left (S75),
right (S76) and right (S71) (Fig. 13.5). From the squatting position (S75 and
13 Episodes of Magdalenian Hunter-Gatherers in Tuc dAudoubert 233
Fig. 13.5 Complete event 13 and excerpt of events 15, 16 and 17 in the upper gallery of Tuc
dAudoubert with the respective spoor number. (Photo Association Louis Bégouën/Tracking in
Caves)
S76), facing the centre of the gallery, subject S7 turned the right foot to the right
and produced the footprint S71. From here the subject moved towards the deep
part of the cave. While sitting in a squatting position, subject S7 probably did
something with the bone in front. Just close to this a right footprint (S81) of
subject S8, female adult, was identied. Some metres from the described scenario
the couple left again their traces. According to the indigenous ichnologists, they
belong to the same event 17 as the other footprints just described. Subject S7 left
an isolated left footprint (S72). Close to it a right footprint (S82) from subject
S8 was identied.
Eastern End Section
Having passed the narrow passage at the sinter basin with the colubrid skeleton, the
gallery widens again (Fig. 13.3). There are bear bones scattered around, including a
right mandible placed on a rock, deprived of its canine tooth. Three metres further
on, on the left, heels and footprints (TUC-324) precede a young bear skeleton
(TUC-325) lying in the clay soil. The vertebral column is anatomically connected.
There are also two holes made by a at tool used as a lever to loosen the bones. Two
vertebrae taken by prehistoric humans are deposited next to it. By its dimensions, a
left mandible with missing teeth recalls the rst one located 6.5 m away. Due to the
connection with the results of the investigations at this place by the indigenous
ichnologists, the following location 6 metres further on is remarkable. Here on the
left side of the path, another bear skeleton (TUC-326) is scattered over a small area at
the foot of the cave wall. The distance from walkable areas prohibits detailed
observation, but it is an accumulation of diverse bones that are no longer in
anatomical connection. At the end of the Galerie des Empreintes, three perforated
teeth, two bison incisors and a fox canine, are aligned on the oor along the wall. Ten
centimetres before these teeth, a niche in the wall is completely stained with red
ochre.
According to the footprints, three subjects were on their way in this area. The
indigenous ichnologists identied three events: again the couple of subjects S7, male
adult, and S8, female adult, whose trackways lead exactly to a cave chamber in
which bear bones were dug out and back towards the entrance of the cave. Further-
more subject S9, male infans II, has not yet been identied in Tuc dAudoubert. The
three subjects have left a total of four trackways with up to four footprints and one
isolated track.
234 A. Pastoors et al.
Event 18: The rst event describes