Hofmeyr: A Late Pleistocene Human Skull from South Africa (Vertebrate Paleobiology and Paleoanthropology) 3031074254, 9783031074257

This edited volume provides the historical and geological background, as well as the details pertaining to the morpholog

124 85 12MB

English Pages 270 [261] Year 2022

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Foreword
Preface
Contents
Contributors
1 Introduction: The Fossil Record of Homo sapiens in Africa – Morphological Variability in the Late Quaternary and the Significance of the Hofmeyr Skull
Abstract
Acknowledgements
References
Background and History
2 Lost and Found: The Discovery and Rediscovery of the Hofmeyr Skull
Abstract
Introduction
The Discovery
The First Analyses
The Rediscovery
Hofmeyr and Changing Methods in the Study of Human Origins
Acknowledgements
References
3 Genetic Divergence Within Southern Africa During the Later Stone Age
Abstract
Introduction
The Cradle of Modern Humans
Click-Speaking Hunter-Gatherers Within Africa
Substructure Among the Khoe-San
Late Holocene Gene Flow
Future Directions
Acknowledgements
References
4 Geological Setting of the Hofmeyr Locality
Abstract
Introduction
Geography
Basement Geology
Karoo Supergroup
Karoo Bedrock Minerology
Gondwana Breakup
Post-breakup Southern Africa
Cenozoic Geology
Conclusions
Acknowledgements
References
5 Isotopic Evidence for the Geographic Origin, Movement and Diet of the Hofmeyr Individual
Abstract
Introduction
Background
Strontium Isotopes
Oxygen Isotopes
Carbon Isotopes
Geology
Methods
Hofmeyr Skull Samples
Bioavailable Strontium Isotope Study
Results
Bioavailable Strontium Study
Strontium of the Hofmeyr Specimen
Carbon, Oxygen, and Nitrogen of the Hofmeyr Specimen
Enamel Structure and Crown Development in Hofmeyr
Discussion
Bioavailable Strontium Isotope Distributions in the Karoo
Potential Origin and Movements of the Hofmeyr Individual
Dietary Ecology and Paleoclimate
Conclusion
Notes
Acknowledgements
References
Morphological and Morphometric Analyses of the Skull as a Whole
6 Description and Comparative Morphology of the Hofmeyr Skull
Abstract
Introduction
Comparative Samples
Holocene and Recent Humans
Late Pleistocene Humans
Background
Preservation, Damage and Reconstruction
Ontogenetic Age of the Skull
Evidence from the Dentition
Evidence from Cranial Suture Fusion
The Neurocranium
Neurocranial Size and Shape
Individual Components of the Neurocranium
Frontal
Parietal
Temporal
The Face
Overall Facial Size and Shape
Individual Components of the Face
The Circumorbital Region
The Midface
The Mandibular Ramus
The Dentition
Comparisons Summary
Acknowledgements
Appendix 6.A1
References
7 The Position of the Hofmeyr Skull within Late Pleistocene and Holocene African Regional Diversity: 2D and 3D Morphometric Analyses
Abstract
Introduction
Case Study One
Materials and Methods
Results of Case Study One
Discussion of Case Study One
Case Study Two
Material and Methods
Results of Case Study Two
Discussion of Case Study Two
Comparison of the Two Case Studies and Conclusions
Implications for Hypotheses About Population History
Acknowledgements
Appendix 7.A1
References
8 Cranial Form of the Hofmeyr Skull: Comparative 3D Geometric Morphometrics
Abstract
Introduction
Material and Methods
Results
Discussion
Acknowledgements
References
Morphological and Morphometric Analyses of Particular Aspects of the Skull
9 The Endocast of the Late Pleistocene Human Skull from Hofmeyr
Abstract
Introduction
Methods and Comparative Material
The Comparative Sample
Reconstruction of the Hofmeyr Endocast
Analytical Procedures
Morphological Description of the Hofmeyr Endocast
Hofmeyr Endocranial Volume
Comparisons of Endocranial Shape
Hofmeyr and Present-Day Humans
Hofmeyr and Other Fossils
Conclusions
Acknowledgements
References
10 The Hofmeyr Bony Labyrinth: Morphological Description and Affinity
Abstract
Introduction
Materials and Methods
Results
Univariate and Bivariate Analyses
Multivariate Analyses
Discussion and Conclusions
Acknowledgements
Appendix 10.A1
References
11 The Paranasal Sinuses of the Hofmeyr Cranium
Abstract
Introduction
Background
Frontal Sinus Anatomy
Maxillary Sinus Anatomy
Frontal and Maxillary Sinuses in Paleoanthropology
Materials and Methods
MicroCT Scanning and Visualization of the Hofmeyr Paranasal Sinuses
Determination of Hofmeyr Sinus Volumes and Linear Dimensions
An Overview of Analytical Comparisons
Results – the Frontal Sinus
Frontal Sinuses of the Hofmeyr Cranium
Comparisons with Previously Published Data
Results – the Maxillary Sinus
The Maxillary Sinus of the Hofmeyr Cranium
Comparisons with Previously Published Data
Discussion
Maxillary Sinus Size
Frontal Sinus Size
Methodological Considerations and Future Directions
Conclusion
Acknowledgements
References
12 The Dentition of the Hofmeyr Skull
Abstract
Introduction
Hofmeyr: Preservation and Damage
Morphological Observations from the 1965 Photographs
Tooth Wear
Maxillary Midline Diastema
Morphological Observations on the Existing Teeth
M2 Crown Morphology
M2 Roots
M3 Crown Morphology
M3 Roots
M3 Pulp Chamber
M3 Crown Morphology
M3 Enamel
M3 Root
M3 Pulp Chamber
Molar Size and Shape
Comparisons of MD and BL Crown Diameters
Comparisons of Crown and Cervical Diameters
Discussion
Conclusion
Acknowledgements
References
Summary Perspective on Hofmeyr
13 Hofmeyr: A Summary Perspective on the Context and Morphology of a Late Pleistocene Human Skull from South Africa
Abstract
Introduction
The Significance of MIS 3 – Environmental Variability
The Significance of MIS 3 – Archaeological Variability
The Significance of MIS 3 – The Scanty Human Fossil Record
Part I
Part II
Part III
Conclusion
Acknowledgements
References
14 Correction to: Hofmeyr: A Late Pleistocene Human Skull from South Africa, Vertebrate Paleobiology and Paleoanthropology
Correction to: F. E. Grine (ed.), Hofmeyr: A Late Pleistocene Human Skull from South Africa, Vertebrate Paleobiology and Paleoanthropology, https://doi.org/10.1007/978-3-031-07426-4
464135_1_En_BookBackmatter_OnlinePDF
Recommend Papers

Hofmeyr: A Late Pleistocene Human Skull from South Africa (Vertebrate Paleobiology and Paleoanthropology)
 3031074254, 9783031074257

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Vertebrate Paleobiology and Paleoanthropology Series

Frederick E. Grine Editor

Hofmeyr A Late Pleistocene Human Skull from South Africa

Hofmeyr

Vertebrate Paleobiology and Paleoanthropology Series Edited by Eric Delson City University of New York and American Museum of Natural History NY, USA [email protected] http://www.nycep.org/ed/

Eric J. Sargis Yale University, New Haven, USA [email protected] http://www.yale.edu/anthro/people/esargis.html Focal topics for volumes in the series will include systematic paleontology of all vertebrates (from agnathans to humans), phylogeny reconstruction, functional morphology, Paleolithic archaeology, taphonomy, geochronology, historical biogeography, and biostratigraphy. Other fields (e.g., paleoclimatology, paleoecology, ancient DNA, total organismal community structure) may be considered if the volume theme emphasizes paleobiology (or archaeology). Volumes in the series may either be monographic treatments (including unpublished but fully revised dissertations) or edited collections, especially those focusing on problem-oriented issues, with multidisciplinary coverage where possible. The two Series Editors are assisted by an Editorial Advisory Board. All contributions in the series (whether monographs or chapters in edited volumes) will be peer-reviewed by at least three readers, at the level of a journal submission.

Editorial Advisory Board: Ross D.E. MacPhee (American Museum of Natural History, New York, NY, USA), Peter Makovicky (University of Minnesota, Minneapolis, MN, USA), Sally McBrearty (University of Connecticut, Storrs, CT, USA), Jin Meng (American Museum of Natural History, New York, NY, USA), Tom Plummer (Queens College/CUNY, Queens, NY, USA).

Hofmeyr A Late Pleistocene Human Skull from South Africa

Edited by

Frederick E. Grine Departments of Anthropology and Anatomical Sciences, Stony Brook University, Stony Brook, NY, USA

123

Editor Frederick E. Grine Departments of Anthropology and Anatomical Sciences Stony Brook University Stony Brook, NY, USA

ISSN 1877-9077 ISSN 1877-9085 (electronic) Vertebrate Paleobiology and Paleoanthropology ISBN 978-3-031-07425-7 ISBN 978-3-031-07426-4 (eBook) DOI 10.1007/978-3-031-07426-4 © Springer Nature Switzerland AG 2022, corrected publication 2023 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific 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 affiliations. Cover image: Photo of the Karoo landscape of the region near Hofmeyr taken from the lower slopes of the Bamboesberg Mountains. The conical hill, known as Spitskop, is located directly east of the Hofmeyr locality. The Burgersdorp Formation is exposed on the proximate range of hills. Photo by F.E. Grine. Inset photo of the Hofmeyr cranium by F.E. Grine. Cover design by F.E. Grine and K. Thompson. This Springer imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

Foreword

Over the past decade and a half, the Hofmeyr skull has ignited much curiosity and interest from members of the public, both international and local, who visit the East London Museum. As a well-travelled specimen resting in its original home after its discovery almost seven decades ago, the skull has rekindled a reminder of the rich fossil heritage of the Eastern Cape Province of South Africa and a deeper connection to the story of the evolution of our species. Today the sleepy Karoo town of Hofmeyr has its name carried far and wide by a chance find of a hominin skull from the eroded banks of the Vlekpoort River in 1954. The town has now been elevated to scientific significance through the dedicated research of the authors of the chapters in this volume and the persistent leadership and guidance of its editor, Frederick Grine. The skull rose to worldwide attention in 2007 with the first seminal paper published by Fred Grine and colleagues in Science entitled ‘Late Pleistocene human skull from Hofmeyr, South Africa, and modern human origins’. Indeed, the importance of Hofmeyr was recognized by Time magazine, which declared this Science paper to have conveyed one of the top ten scientific discoveries of 2007. A number of questions have had their answers revealed in the chapters of this book which serve the growing fascination people have regarding our evolutionary history. The skull remained unattended and ‘lost’ to science until Alan Morris ‘re-discovered’ it as part of audited research on museum human skeletal remains. This was a first and important step in bringing the Hofmeyr skull back into the scientific domain and more importantly bridging a gap between mere curiosity and its link in the study of palaeoanthropology. This up-to-date collection of work on the Hofmeyr skull elucidates a detailed multidisciplinary synthesis of research on the specimen ranging from the history of its discovery to all aspects of its morphology, geological setting, the movement and diet of the Hofmeyr individual, an interpretation of the endocast, the bony labyrinth and paranasal sinuses and its dentition. It also provides encouragement to museum curators who have human skeletal material in their care never to underestimate the value of any specimen that has not been rigorously investigated particularly in the light of the rapid advances in dating techniques, 2D and 3D morphometric analysis and collaborative research which made sense of the medley of attributes that a single find revealed. This volume of work makes a further important contribution to the manner in which museums and heritage authorities value paleoanthropological specimens. For the East London Museum, it has placed the Hofmeyr skull into a wide paleoanthropological research orbit allowing a more comprehensive volume of facts to be synthesized for educational programmes which are integral to the vision and objectives of our South African heritage institutions. In addition, through the efforts of Fred Grine, a cast of the Hofmeyr skull has found a place in one of the most comprehensive human origins displays in any museum in the world—the Koch Hall of Human Origins at the Smithsonian Institution’s National Museum of Natural History in Washington, DC. This has brought great credence to the Eastern Cape Province and the curatorial home of the skull. At a local level, a second cast has been used extensively for v

vi

Foreword

outreach programmes to promote the study of palaeoanthropology, and this has been a popular point of call for hundreds of learners at the annual Scifest Africa held in Makhanda (previously Grahamstown). Considering that the Hofmeyr site is only one of two in Africa to reveal a reasonably complete human cranium dating to Marine Isotope Stage 3 (the other being at Nazlet Khater in Egypt) it was vitally important that high-quality research was undertaken on the specimen. The various chapters of this volume carry the themes of its background and history, the morphological and morphometric analyses of the skull as a whole and detailed analyses of selected aspects. A final concluding summary highlights the significance of the specimen in understanding human evolution. The Eastern Cape Province harbours a rich mosaic of coastal middens, Khoisan rock art and lithic sites and the addition of a Late Pleistocene skull to this palette enriches our understanding of the prehistory of Homo sapiens in the region. The body of work to follow will cement this mosaic by revealing the importance of a rare discovery, the first of its kind for the region, and encourage further paleoanthropological research in this unique part of the world. ,

Kevin Cole Principal Natural Scientist, East London Museum East London, South Africa

Preface

There are some issues in human paleontology that are seemingly timeless. One of these is the emergence and subsequent evolution of our own species. An abundance of genetic and paleontological evidence implicates Africa not only as the geographic source of Homo sapiens during Marine Isotope Stage (MIS) 7 but also as the home to subsequent morphological and behavioral development until the first successful, lasting emigration into Eurasia and beyond in MIS 3. The African fossil record for our species over this lengthy period is comprised largely of isolated teeth or small cranial and postcranial fragments. Only a dozen or so sites have provided particularly informative crania and/or skeletons, and the majority of these date to MIS 7–5. None are known from MIS 4, and only two that date to MIS 3 at between 37.6 and 36.2 ka—Nazlet Khater and Hofmeyr—have yielded reasonably complete skulls. MIS 3 is of paleoanthropological significance because of its climatic variability across Africa and especially in southern Africa as the continent experienced a transition from a moderate climatic period in MIS 4 to a severe glacial period, the Last Glacial Maximum in MIS 2. The period between 57 and 20 ka also saw a mosaic of archaeological technologies with the transition from the Middle Stone Age (MSA) to the Later Stone Age (LSA). The lithic record in this span speaks to a geographically adventitious transition to greater bipolar reduction and lithic miniaturization but with persistent regional variability and asynchronous change through time. In contrast to the detailed coverage that has been afforded the MIS 3 human remains and the site of Nazlet Khater, Hofmeyr has been provided with comparatively little documentation. Hofmeyr is undoubtedly one of the most important Late Pleistocene specimens in Africa. The present volume provides a compendium of research papers by some of the world’s leading experts in their respective fields that deal with the background and morphology of this paleoanthropological treasure. These contributions provide an excellent view of what we know about human evolution and adaptation in the Late Pleistocene, and what additional knowledge we seek in the quest to more fully appreciate the Late Pleistocene palaeoanthropology of Africa. I am grateful to all of the colleagues who have devoted their time and expertise to this effort. I hope that they have gained fulfillment from contributing to the documentation, analysis and understanding of Hofmeyr. I am also grateful to Springer’s Vertebrate Paleobiology and Paleoanthropology Series Editor, Eric Delson, for his commitment, unwavering assistance and support throughout all stages of this volume’s development. I thank all 29 peer-reviewers for their cogent comments and suggestions on the chapters that comprise this volume. I am grateful to the Production Staff at Springer for their care and devotion to excellence in the production of this book. We didn’t always make it easy for them! I thank Dr. Katharine E. T. H. Thompson for her skillful assistance in creating the cover illustration and helping with finalizing many of the figures. Stony Brook, USA

Frederick E. Grine

vii

The original version of the book was revised: the copyright year in all in-line references of the chapters was incorrect. This has now been corrected. The correction to the book is available at https://doi.org/10.1007/978-3-031-07426-4_14

Contents

1

Introduction: The Fossil Record of Homo sapiens in Africa – Morphological Variability in the Late Quaternary and the Significance of the Hofmeyr Skull . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Frederick E. Grine

Part I

1

Background and History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2

Lost and Found: The Discovery and Rediscovery of the Hofmeyr Skull . . . . Alan G. Morris

9

3

Genetic Divergence Within Southern Africa During the Later Stone Age . . . Dana R. Al-Hindi, Austin W. Reynolds, and Brenna M. Henn

19

4

Geological Setting of the Hofmeyr Locality . . . . . . . . . . . . . . . . . . . . . . . . . Johann Neveling

29

5

Isotopic Evidence for the Geographic Origin, Movement and Diet of the Hofmeyr Individual . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Sandi R. Copeland, Vaughan Grimes, Johann Neveling, Julia A. Lee-Thorp, Frederick E. Grine, Zhaoping Yang, Christopher Dean, and Michael P. Richards

Part II

47

Morphological and Morphometric Analyses of the Skull as a Whole . . .

6

Description and Comparative Morphology of the Hofmeyr Skull . . . . . . . . . Frederick E. Grine

7

The Position of the Hofmeyr Skull within Late Pleistocene and Holocene African Regional Diversity: 2D and 3D Morphometric Analyses . . . . . . . . . . Isabelle Ribot, Yassmine Ghalem, and Isabelle Crevecoeur

119

Cranial Form of the Hofmeyr Skull: Comparative 3D Geometric Morphometrics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Philipp Gunz and Sarah E. Freidline

143

8

Part III

9

71

Morphological and Morphometric Analyses of Particular Aspects of the Skull . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

The Endocast of the Late Pleistocene Human Skull from Hofmeyr . . . . . . . . Simon Neubauer

153

ix

x

Contents

10 The Hofmeyr Bony Labyrinth: Morphological Description and Affinity . . . . Isabelle Crevecoeur, Adrien Thibeault, Linda Bouchneb, Marie Matu, Bruno Maureille, and Isabelle Ribot

165

11 The Paranasal Sinuses of the Hofmeyr Cranium . . . . . . . . . . . . . . . . . . . . . Lauren N. Butaric, Laura T. Buck, Antoine Balzeau, Anton du Plessis, and Frederick E. Grine

179

12 The Dentition of the Hofmeyr Skull . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Wendy Black and Frederick E. Grine

213

Part IV

Summary Perspective on Hofmeyr. . . . . . . . . . . . . . . . . . . . . . . . . . . .

13 Hofmeyr: A Summary Perspective on the Context and Morphology of a Late Pleistocene Human Skull from South Africa . . . . . . . . . . . . . . . . . Frederick E. Grine

237

Correction to: Hofmeyr . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Frederick E. Grine

C1

Subject Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

253

Specimen / Site Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

257

Contributors

Dana R. Al-Hindi Department of Anthropology, Davis Genome Center, University of California, Davis, CA, USA Antoine Balzeau PaleoFED Team, UMR 7194, CNRS, Département Homme et Environnement, Muséum national d’Histoire naturelle. Musée de L’Homme, Paris, France; Department of African Zoology, Royal Museum for Central Africa, Tervuren, Belgium Wendy Black Department of Archaeology, Iziko Museums of South Africa, Cape Town, South Africa Linda Bouchneb Laboratoire de la Préhistoire à l’Actuel: Culture, Environnement et Anthropologie, Université de Bordeaux, UMR 5199-PACEA, Pessac Cedex, France Laura T. Buck School of Biological and Environmental Sciences, Liverpool John Moores University, Liverpool, UK Lauren N. Butaric Department of Anatomy, Des Moines University, Des Moines, IA, USA Sandi R. Copeland Environmental Stewardship Group, Los Alamos National Laboratory, P.O. Box 1663 Los Alamos, NM, USA Isabelle Crevecoeur Laboratoire de la Préhistoire à l’Actuel: Culture, Environnement et Anthropologie, Université de Bordeaux, UMR 5199-PACEA, Pessac Cedex, France; Chargée de Recherche CNRS, Université de Bordeaux, UMR 5199-PACEA, Pessac Cedex, France Christopher Dean Centre for Human Evolution Research, Natural History Museum, Cromwell Road, London, UK; Department of Cell and Developmental Biology, University College London, Gower Street, London, UK Anton du Plessis Department of Physics, Stellenbosch University, Stellenbosch, South Africa; Object Research Systems, 460 Saint-Catherine St. W, Montreal, Quebec H3B 1A7, Canada Sarah E. Freidline Department of Anthropology, University of Central Florida, Orlando, FL, USA Yassmine Ghalem Département d’Anthropologie, Université de Montréal, Montréal, QC, Canada Vaughan Grimes Department of Archaeology, Memorial University, St. John’s, NL, Canada; Department of Earth Sciences, Memorial University, St. John’s, NL, Canada

xi

xii

Frederick E. Grine Department of Anthropology, Stony Brook University, Stony Brook, NY, USA; Department of Anatomical Sciences, Stony Brook University School of Medicine, Stony Brook, NY, USA Philipp Gunz Department of Human Evolution, Max Planck Institute for Evolutionary Anthropology, Leipzig, Germany Brenna M. Henn Department of Anthropology, Davis Genome Center, University of California, Davis, CA, USA Julia A. Lee-Thorp School of Archaeology, University of Oxford, Oxford, UK Marie Matu Laboratoire de la Préhistoire à l’Actuel: Culture, Environnement et Anthropologie, Université de Bordeaux, UMR 5199-PACEA, Pessac Cedex, France Bruno Maureille Laboratoire de la Préhistoire à l’Actuel: Culture, Environnement et Anthropologie, Université de Bordeaux, UMR 5199-PACEA, Pessac Cedex, France Alan G. Morris Department of Human Biology, University of Cape Town, 7700 Observatory, Cape Town, South Africa Simon Neubauer Institute of Anatomy and Cell Biology, Medical Faculty, Johannes Kepler University Linz, MED Campus I, ADM 4, Krankenhausstrasse 5, 4020, Linz, Austria; Department of Human Evolution, Max Planck Institute for Evolutionary Anthropology, Leipzig, Germany Johann Neveling Geoscience Mapping Unit, Council for Geoscience, Pretoria, South Africa Austin W. Reynolds Department of Anthropology, Davis Genome Center, University of California, Davis, CA, USA; Department of Anthropology, Baylor University, Waco, TX, USA Isabelle Ribot Département d’Anthropologie, Université de Montréal, Montréal, QC, Canada Michael P. Richards Department of Archaeology, Simon Fraser University, Burnaby, BC, Canada Adrien Thibeault Laboratoire de la Préhistoire à l’Actuel: Culture, Environnement et Anthropologie, Université de Bordeaux, UMR 5199-PACEA, Pessac Cedex, France Zhaoping Yang Geological Survey of Canada, Ottawa, ON, Canada

Contributors

Chapter 1

Introduction: The Fossil Record of Homo sapiens in Africa – Morphological Variability in the Late Quaternary and the Significance of the Hofmeyr Skull Frederick E. Grine

Abstract Late Quaternary African fossils from MIS 8 to MIS 2 provide some idea of the considerable morphological diversity that accompanied the emergence and subsequent evolution of Homo sapiens. Fossils that are universally accepted as belonging to our species first appear in MIS 7 at about 200 ka. The intervening 190,000 years until just prior to the Holocene finds human fossils from approximately 50 sites across the African continent, but the vast majority are represented by isolated teeth or small fragments of bone. Only a dozen or so particularly informative crania are known from this lengthy span of time. Most of these date to between MIS 7 and MIS 5, none are known from MIS 4, and only two date to MIS 3. The final key phase of human evolutionary development, our worldwide expansion from Africa, occurred in MIS 3 between about 60 and 40 ka. Hofmeyr represents the only example of sub-Saharan cranial morphology that dates to the temporal span of MIS 3. The chapters in this volume document the background to this specimen and provide details of its anatomy in the context of recent African populations and penecontemporaneous Late Pleistocene crania from North Africa and Eurasia.



Keywords Cranial variability Diversity Evolution



MIS 8



MIS 7



MIS 3



The Late Quaternary fossil record of Africa provides a tantalizing glimpse into the considerable morphological diversity that accompanied the emergence and subsequent evolution of Homo sapiens. Genomic evidence indicates that

F. E. Grine (&) Department of Anthropology, Stony Brook University, Stony Brook, NY 11794-4364, USA e-mail: [email protected] Department of Anatomical Sciences, Stony Brook University School of Medicine, Stony Brook, NY, USA

the lineage that culminates in us may have separated from that leading to Neandertals and Denisovans prior to 430 ka (Meyer et al., 2016). The majority of genetic studies indicate the earliest population divergence within Homo sapiens (i.e., between the Khoesan and all other living humans [Veeramah & Hammer, 2014]) occurred sometime between ca. 200– 150 ka (Chan et al., 2015; Gronau et al., 2011; Henn et al., 2018; Mallick et al., 2016; Schiffels & Durbin, 2014; Schlebusch et al., 2012; Song et al., 2017; Veeramah et al., 2012), although even earlier divergence dates of 300–250 ka (Scally & Durbin, 2012) or perhaps 350–260 ka (Schlebusch et al., 2017) have been proposed. With regard to the fossil evidence pertaining to the emergence of our species, it is, of course, possible to attribute any specimen to H. sapiens that is seen to exhibit any trait by which it resembles modern humans more than archaic forms such as H. erectus. Depending upon one’s predilection, a fossil that evinces even a single synapomorphy with us can be viewed as an “early archaic” member of our species. Thus, for example, Bräuer (2001, 2008) referred the Kabwe 1 cranium – the holotype of Homo rhodesiensis – along with the Saldanha calotte and the Bodo cranium to “early archaic Homo sapiens” that was viewed as being transitional to us from “developed” Homo erectus. However, fossils that arguably display sufficient morphology to qualify as members of populations with obviously close affinities to H. sapiens or perhaps even as early members of the species itself appear only in MIS 8 at about 300–260 ka. These include specimens such as the cranium from Florisbad that has been dated to approximately 259 ± 35 ka (e.g., Bräuer, 2008; Bruner et al., 2013; Grün et al., 1996; Lahr & Foley, 1998; Pearson, 2013; Rightmire, 1978, 2009; Stringer, 1996, 2016). Other specimens that are possibly somewhat older are known from the Moroccan site of Jebel Irhoud. These fossils were initially dated to 190– 90 ka from ESR estimates on associated mammalian teeth (Grün & Stringer, 1991), and this age was supported by a U-series/ESR profile of 160 ± 16 ka for a human tooth

© Springer Nature Switzerland AG 2022 F. E. Grine (ed.), Hofmeyr: A Late Pleistocene Human Skull from South Africa, Vertebrate Paleobiology and Paleoanthropology, https://doi.org/10.1007/978-3-031-07426-4_1

1

2

fragment (Smith et al., 2007). However, subsequent thermoluminescence (TL) dating of burnt flints suggested an age of 315 ± 34 ka and this has been supported by a “recalculated” U-series ESR date of 286 ± 32 ka for hominin tooth enamel (Richter et al., 2017). One is left wondering whether these recalculated dates are perhaps being driven by the similarly ancient molecular divergence estimates (Meyer et al., 2016; Scally & Durbin, 2012; Schlebusch et al., 2017). The Jebel Irhoud specimens possess undivided supraorbital tori, no chin, and an elongate rather than a globular neurocranium. In many respects, they are no more modern than the Florisbad cranium. Although some workers have perceived the Jebel Irhoud assemblage as being attributable to H. sapiens (e.g., Hublin, 1992; Hublin et al., 2017; Stringer & Buck, 2014), others have regarded it as representing a late-surviving archaic population (e.g., Bruner & Pearson, 2013). The relationships between MIS 8 populations as represented by specimens like those from Florisbad and Jebel Irhoud and living Homo sapiens are far from resolved. That is not to say that all specimens that derive from this temporal period (MIS 8 at between ca. 300–243 ka) are necessarily attributable to H. sapiens or even to a lineage leading to it. Indeed, bones of an undoubtedly distinct species, Homo naledi, whose phylogenetic relationships remain unclear, have been recovered from the Rising Star Cave system, South Africa with an estimated age of some 335– 236 ka (Berger et al., 2015; Dirks et al., 2017; Hawks et al., 2017). The oldest fossil over which there is unanimity concerning its undoubted attribution to H. sapiens is the Omo I skull and partial skeleton from the upper part of Member 1 of the Kibish Formation, Ethiopia. Omo I is one of three fossils recovered from these deposits in 1967 by a team led by Leakey (1969). Omo II comprises a well-preserved calvarium and Omo III consists of neurocranial and facial fragments. Omo II presents somewhat more archaic morphology than Omo I, which has led to suggestions of a single population that showed considerable individual variation (Day, 1969; Rightmire, 1976; Trinkaus, 2005), the contemporaneous presence of two different populations (Day & Stringer, 1982), or that Omo I is younger and possibly intrusive (e.g., Bräuer, 2001; Bräuer et al., 1997; Chavaillon, 1982). However, subsequent fieldwork led by J. G. Fleagle resulted in the discovery of additional parts of the Omo I skeleton that had weathered out of the level excavated in 1967 by Leakey. Some of these refitted with bones discovered in 1967 (Pearson et al., 2008). This discovery, together with more comprehensive stratigraphic analyses of the site (Brown & Fuller, 2008; McDougall et al., 2005) have provided ample evidence for the contemporaneity of the Member 1 hominin specimens. They were deposited during a period of high-water level in Lake Turkana correlated with Mediterranean sapropel 7 at c. 197 ka (McDougall et al.,

F. E. Grine

2008). Additionally, the Nakaákire Tuff, which outcrops just below the level of the fossils, has been dated to 196 ka (McDougall et al., 2005), and the KHS Tuff at the base of Member 2, which overlies the hominin recovery sites, has been dated by correlation with other tuffs in East Africa to 172 ka (Brown et al., 2012). This served to securely bracket the Omo I and II fossils to between 197 and 172 ka, with their stratigraphic position placing them closer to ca. 195 ka. Aubert et al. (2012) conducted U-series dating on a fragment of the Omo I cranium, which yielded a minimum age of some 187–155 ka. However, a more recent study has concluded that the KHS tuff in the overlying Member II of the Kibish Formation has an age of 233 ± 22 ka based on correlations with tuffs elsewhere in Ethiopia (Vidal et al., 2022). As such, the Omo I skeleton would appear to be just over 233 ka old. Fossils that are universally recognized as being attributable to Homo sapiens, as defined by the possession of a number of morphological synapomorphies shared with us, appear for the first time in Africa about 200 ka (Weaver, 2012) in the temporal span between MIS 7 and MIS 6. Not only at this time, but subsequently throughout the Pleistocene, the degree of morphological variation among African crania can be rather striking, and its significance continues to be the subject of discussion (e.g., Hammer et al., 2011; Scerri et al., 2018; Stringer, 2007). Unfortunately, the human fossil record in Africa over the next 170,000 years until MIS 2 is decidedly paltry. Indeed, it is only after some 15,000 years ago (e.g., at North African sites such as Afalou-bou-Rhummel, Jebel Sahaba 117 and Grotte des Pigeons) that reasonably large numbers of skeletons are known (Grine, 2016). African Homo sapiens fossils that are constrained between 200 ka (MIS 7) and just prior to the Holocene (in MIS 2) are known from approximately 50 sites. While this number is impressive, evidence from the vast majority of sites consists of isolated teeth or small cranial and postcranial fragments. Only a dozen or so sites have provided particularly informative specimens, and the majority of these (e.g., those from Herto, Singa, Ngaloba, Klasies River, Dar-Es-Soltan, Eyasi and Aduma) date to MIS 7 - MIS 5. None are known from MIS 4, and only two sites that date to MIS 3 have yielded reasonably complete crania (Grine, 2016). The first of these sites, Nazlet Khater in Egypt (Bouchneb & Crevecoeur, 2009; Crevecoeur, 2008, 2012; Crevecoeur et al., 2009; Vermeersch, 2002), has produced two specimens dated to 37.6 ± 3.5 ka (AMS 14C) and 38 ± 6 ka (ESR on tooth enamel). The cranium, which has been described as “robust,” presents an overall modern appearance with a prominent chin, rounded cranial form, modest dental dimensions and no supraorbital torus. At the same time, however, Nazlet Khater evinces several archaic features, including a thick cranial vault, a broad ramus and a robust

1

Introduction

mandibular corpus. Its inner ear morphology is also unusual among recent humans but occurs with some frequency among both Middle and Upper Paleolithic specimens from Eurasia (Bouchneb & Crevecoeur, 2009; Crevecoeur, 2012). The second specimen from MIS 3 is the partial skull from Hofmeyr, South Africa (Grine et al., 2007). The skull was discovered in 1954 by a farmer digging for sand in a dry channel bed of the Vlekpoort River. It was found without any associated archaeological or faunal evidence, and its geological context precluded any assessment of its geochronological age. Although the bone lacked sufficient collagen to be amenable to direct AMS 14C dating, extraction of the heavily indurated carbonate sand matrix that largely filled the endocranial cavity enabled assessment using a combination of OSL and uranium-series dating methods. This resulted in an estimate of its burial time at 36.2 ± 3.3 ka (Grine et al., 2007). The cranium is morphologically modern overall, but it possesses a moderately strong supraorbital torus and projecting glabella as well as marked alveolar prognathism (Grine et al., 2007). Preliminary morphometric assessment indicated that while the cranium could be accommodated within the 95% confidence ellipse of some sub-Saharan African samples, it was rather distinct from Khoesan crania and shared distinct affinities with penecontemporaneous Upper Palaeolithic specimens from Eurasia (Grine et al., 2007). Subsequent comparisons between Hofmeyr and remains such as Nazlet Khater point to a greater range of variation among these Late Pleistocene human crania than is evident today (Crevecoeur et al., 2009). The Hofmeyr skull relates to what Bergström et al. (2021) have identified as the final of three key phases related to modern human evolution, namely our worldwide expansion between about 60 and 40 ka. Bergström et al. (2021) have also emphasized the high degree of morphological variability that is evident in this and the earlier fossil records of Homo sapiens in Africa. Indeed, this morphological variability seems to be mirrored by recent analyses that have highlighted genomic variation across Africa and its potential to speak to past migration and selection events (Choudhury et al., 2020a, 2020b; Hollfelder et al., 2021; Ko et al., 2013; Schlebusch et al., 2020; Sirugo et al., 2019; Tishkoff et al., 2009; Vincente & Schlebusch, 2020; Williams et al., 2021). Whereas the Nazlet Khater specimens have received considerable attention and analysis, there has been comparatively little work done on the Hofmeyr skull beyond its initial description (Grine et al., 2007), comparison with Nazlet Khater (Crevecoeur et al., 2009), and its partial reconstruction (Grine et al., 2010). The present volume documents the background of this unique specimen and details its morphology in the context of recent African populations and penecontemporaneous Late Pleistocene samples from North Africa and Eurasia. The contributions in Part 1 present information that enables

3

appreciation of the contextual significance of the skull. Although it has not been possible to extract endogenous aDNA from Hofmeyr, a knowledge of the Later Stone Age genetic landscape permits the specimen to be placed in context. Hofmeyr is unique among South African human fossils in deriving from the Karoo of the Eastern Cape Province, and appreciation of its geological setting is essential to interpret the data relating to its biogeochemical assessment. Part 1 also includes a discussion of the history of the discovery and subsequent mishandling of the skull. The three chapters in Part 2 deal with the morphological and morphometric comparisons of the skull as a whole. The first of these provides a detailed morphological description of the specimen, and the other two entail 2D and 3D geometric morphometric studies that place the cranium in the milieu of recent African as well as penecontemporaneous Eurasian population samples. The contributions in Part 3 provide detailed analyses of the endocranium, the bony labyrinth, the frontal and maxillary sinuses, and the dentition. A final chapter provides a summary and synthesis of the contributions to this volume. Acknowledgements I am grateful to Bernard Wood, John Fleagle and Eric Delson for their cogent comments, which considerably improved the text of this chapter.

References Aubert, M., Pike, A. W. G., Stringer, C., Bartsiokas, A., Kinsley, L., Eggins, S., et al. (2012). Confirmation of a late Middle Pleistocene age for the Omo Kibish 1 cranium by direct uranium-series dating. Journal of Human Evolution, 63, 704–710. Berger, L. R., Hawks, J., de Ruiter, D. J., Churchill, S. E., Schmid, P., Delezene, L. K., et al. (2015). Homo naledi, a new species of the genus Homo from the Dinaledi Chamber, South Africa. eLife, 4, e09560. Bergström, A., Stringer, C., Hajdinjak, M., Scerri, E. M. L., & Skoglund, P. (2021). Origins of modern human ancestry. Nature, 590, 229–237. Bouchneb, L., & Crevecoeur, I. (2009). The inner ear of Nazlet Khater 2 (Upper Paleolithic, Egypt). Journal of Human Evolution, 56, 257– 262. Bräuer, G. (2001). The KNM-ER 3884 hominid and the emergence of modern human anatomy in Africa. In P. V. Tobias, M. A. Rath, J. Moggi-Cecchi, & G. A. Doyle (Eds.), Humanity from African naissance to coming millennia (pp. 191–197). Firenza University Press. Bräuer, G. (2008). The origin of modern anatomy: By speciation or intraspecific evolution? Evolutionary Anthropology, 17, 22–37. Bräuer, G., Yokoyama, Y., Falguères, C., & Mbua, E. (1997). Modern human origins backdated. Nature, 386, 337. Brown, F. H., & Fuller, C. (2008). Stratigraphy and tephra of the Kibish Formation, southwestern Ethiopia. Journal of Human Evolution, 55, 366–403. Brown, F. H., McDougall, I., & Fleagle, J. G. (2012). Correlation of the KHS Tuff of the Kibish Formation to volcanic ash layers at other sites, and the age of early Homo sapiens (Omo I and Omo II). Journal of Human Evolution, 63, 577–585.

4 Bruner, E., & Pearson, O. (2013). Neurocranial evolution in modern humans: The case of Jebel Irhoud 1. Anthropological Science, 121, 31–41. Bruner, E., Athreya, S., de la Cuétara, J. M., & Marks, T. (2013). Geometric variation of the frontal squama in the genus Homo: Frontal bulging and the origin of modern human morphology. American Journal of Physical Anthropology, 150, 313–323. Chan, E. K. F., Hardie, R. A., Petersen, D. C., Beeson, K., Bornman, R. M. S., Smith, A. B., et al. (2015). Revised timeline and distribution of the earliest diverged human maternal lineages in southern Africa. PLoS ONE, 10(3), e0121223. Chavaillon, J. (1982). Position chronologique des hominidés fossile d’Ethiopie. In M. A. de Lumley (Ed.), Homo erectus et la place de l’Homme de Tautavel parmi les hominidés fossiles (pp. 766–797). Première congrès international de paléontologie humaine. UNESCO, Colloque International du Centre National de la Récherche Scientifique. Choudhury, A., Sengupta, D., Ramsay, M., & Schlebusch, C. (2020a). Bantu-speaker migration and admixture in southern Africa. Human Molecular Genetics, ddaa274. https://doi.org/10.1093/hmg/ddaa274 Choudhury, A., Aron, S., Botigué, L. R., Sengupta, D., Botha, G., Bensellak, T., et al. (2020b). High-depth African genomes inform human migration and health. Nature, 586, 741–748. Crevecoeur, I. (2008). Étude anthropologique du squelette du Paléolithique supérieur de Nazlet Khater 2 (Égypte). Leuven University Press. Crevecoeur, I. (2012). The Upper Palaeolithic human remains of Nazlet Khater 2 (Egypt) and past modern human diversity. In J. J. Hublin & S. McPherron (Eds.), Modern African origins: A North African perspective (pp. 205–221). Springer. Crevecoeur, I., Rougier, H., Grine, F. E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: Evidence from Nazlet Khater, Peştera cu Oase, and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Day, M. H. (1969). Early Homo sapiens remains from the Omo River region of southwest Ethiopia. Nature, 222, 1132–1138. Day, M. H., & Stringer, C. B. (1982). A reconsideration of the Omo Kibish remains and the erectus-sapiens transition. In M. A. de Lumley (Ed.), Homo erectus et la place de l’Homme de Tautavel parmi les hominidés fossils (pp. 814–846). Première congrès international de paléontologie humaine. UNESCO, Colloque International du Centre National de la Récherche Scientifique. Dirks, P. H. G. M., Roberts, E. M., Hilbert-Wolf, H., Kramers, J. D., Hawks, J., Dosseto, A., et al. (2017). The age of Homo naledi and associated sediments in the Rising Star cave, South Africa. eLife, 6, e24231. Grine, F. E. (2016). The Late Quaternary hominins of Africa: The skeletal evidence from MIS 6-2. In S. C. Jones & B. A. Stewart (Eds.), Africa from MIS 6-2: Population dynamics and paleoenvironments (pp. 323–381). Springer. Grine, F. E., Bailey, R. M., Harvati, K., Nathan, R. P., Morris, A. G., Henderson, et al. (2007). Late Pleistocene human skull from Hofmeyr, South Africa, and modern human origins. Science, 315, 226–229. Grine, F. E., Gunz, P., Betti-Nash, L., Neubauer, S., & Morris, A. G. (2010). Reconstruction of the late Pleistocene human skull from Hofmeyr, South Africa. Journal of Human Evolution, 59, 1–15. Gronau, I., Hubisz, M. J., Gulko, B., Danko, C. G., & Siepel, A. (2011). Bayesian inference of ancient human demography from individual genome sequences. Nature Genetics, 43, 1031–1034. Grün, R., & Stringer, C. B. (1991). Electron spin resonance dating and the evolution of modern humans. Archaeometry, 33, 153–199. Grün, R., Brink, J. S., Spooner, N. A., Taylor, L., Stringer, C. B., Franciscus, R. G., et al. (1996). Direct dating of Florisbad hominid. Nature, 382, 500–501.

F. E. Grine Hammer, M. F., Woerner, A. E., Mendez, F. L., Watkins, J. C., & Wall, J. D. (2011). Genetic evidence for archaic admixture in Africa. Proceedings of the National Academy of Sciences of the United States of America, 108, 15123–15128. Hawks, J., Elliott, M., Schmid, P., Churchill, S. E., de Ruiter, D. J., Garvin, et al. (2017). New fossil remains of Homo naledi from the Lesedi Chamber, South Africa. eLife, 6, e24232. Henn, B. M., Steele, T. E., & Weaver, T. D. (2018). Clarifying models of modern human origins in Africa. Current Opinion in Genetics & Development, 53, 148–156. Hollfelder, N., Breton, G., Sjödin, P., & Jakobsson, M. (2021). The deep population history in Africa. Human Molecular Genetics, ddab005. https://doi.org/10.1093/hmg/ddab005 Hublin, J. J. (1992). Recent human evolution in northwestern Africa. Philosophical Transactions of the Royal Society B, 337, 185–191. Hublin, J. J., Ben-Ncer, A., Bailey, S. E., Freidline, S. E., Neubauer, S., Skinner, M. M., et al. (2017). New fossils from Jebel Irhoud, Morocco and the pan-African origin of Homo sapiens. Nature, 546, 289–292. Ko, W. Y., Rajan, P., Gomez, F., Scheinfeldt, L., An, P., Winkler, C. A., et al. (2013). Identifying Darwinian selection acting on different human APOL1 variants among diverse African populations. American Journal of Human Genetics, 93, 54–66. Lahr, M., & Foley, R. A. (1998). Towards a theory of modern human origins: Geography, demography, and diversity in recent human evolution. American Journal of Physical Anthropology, 107, 137–176. Leakey, R. E. F. (1969). Early Homo sapiens remains from the Omo River region of south-west Ethiopia. Nature, 222, 1132–1133. Mallick, S., Li, H., Lipson, M., Mathieson, I., Gymrek, M., Racimo, F., et al. (2016). The Simons Genome Diversity Project: 300 genomes from 142 diverse populations. Nature, 538, 201–206. McDougall, I., Brown, F. H., & Fleagle, J. G. (2005). Stratigraphic placement and age of modern humans from Kibish, Ethiopia. Nature, 433, 733–736. McDougall, I., Brown, F. H., & Fleagle, J. G. (2008). Sapropels and the age of hominins Omo I and II, Kibish, Ethiopia. Journal of Human Evolution, 55, 409–420. Meyer, M., Arsuaga, J. L., de Filippo, C., Nagel, S., Aximu-Petri, A., Nickel, B., et al. (2016). Nuclear DNA sequences from the Middle Pleistocene Sima de los Huesos hominins. Nature, 531, 504–507. Pearson, O. M. (2013). Africa: The cradle of modern people. In F. H. Smith & J. C. M. Ahern (Eds.), The origins of modern humans: Biology reconsidered (pp. 1–44). Wiley. Pearson, O. M., Royer, D. F., Grine, F. E., & Fleagle, J. G. (2008). A description of the Omo I postcranial skeleton, including newly discovered fossils. Journal of Human Evolution, 55, 421–437. Richter, D., Grün, R., Joannes-Boyau, R., Steele, T. E., Amani, F., Rué, M., et al. (2017). The age of the hominin fossils from Jebel Irhoud, Morocco, and the origins of the Middle Stone Age. Nature, 546, 293–296. Rightmire, G. P. (1976). Relationships of Middle and Upper Pleistocene hominids from sub-Saharan Africa. Nature, 260, 238–240. Rightmire, G. P. (1978). Florisbad and human population succession in southern Africa. American Journal of Physical Anthropology, 48, 475–486. Rightmire, G. P. (2009). Middle and later Pleistocene hominins in Africa and Southwest Asia. Proceedings of the National Academy of Sciences of the United States of America, 106, 16046–16050. Scally, A., & Durbin, R. (2012). Revising the human mutation rate: Implications for understanding human evolution. Nature Reviews Genetics, 13, 745–753. Scerri, E. M. L., Thomas, M. G., Manica, A., Gunz, P., Stock, J. T., Stringer, C., et al. (2018). Did our species evolve in subdivided

1

Introduction

populations across Africa, and why does it matter? Trends in Ecology & Evolution, 33, 582–594. Schiffels, S., & Durbin, R. (2014). Inferring human population size and separation history from multiple genome sequences. Nature Genetics, 46, 919–925. Schlebusch, C. M., Skoglund, P., Sjödin, P., Gattepaille, L. M., Hernandez, D., Jay, F., et al. (2012). Genomic variation in seven Khoe-San groups reveals adaptation and complex African history. Science, 338, 374–379. Schlebusch, C. M., Malmström, H., Günther, T., Sjödin, P., Coutinho, A., Edlund, H., et al. (2017). Southern African ancient genomes estimate modern human divergence to 350,000 to 260,000 years ago. Science, 358, 652–655. Schlebusch, C. M., Sjödin, P., Breton, G., Günther, T., Naidoo, T., Hollfelder, N., et al. (2020). Khoe-San genomes reveal unique variation and confirm the deepest population divergence in Homo sapiens. Molecular Biology and Evolution, 37, 2944–2954. Sirugo, G., Williams, S. M., & Tishkoff, S. A. (2019). The missing diversity in human genetic studies. Cell, 177, 26–31. Smith, T. M., Tafforeau, P., Reid, D. J., Grün, R., Eggins, S., Boutakiout, M., et al. (2007). Earliest evidence of modern human life history in North African early Homo sapiens. Proceedings of the National Academy of Sciences of the United States of America, 104, 6128–6133. Song, S., Sliwerska, E., Emery, S., & Kidd, J. M. (2017). Modeling human population separation history using physically phased genomes. Genetics, 205, 385–395. Stringer, C. B. (1996). Current issues in modern human origins. In W. E. Meikle, F. C. Howell, & N. G. Jablonski (Eds.), Contemporary issues in human evolution (pp. 115–134). Memoir 21. California Academy of Sciences. Stringer, C. B. (2007). The origin and dispersal of Homo sapiens: Our current state of knowledge. In P. Mellars, K. Boyle, O. Bar-Yosef,

5 & C. Stringer (Eds.), Rethinking the human revolution (pp. 15–20). McDonald Institute for Archaeological Research Monograph Series. Stringer, C. (2016). The origin and evolution of Homo sapiens. Philosophical Transactions of the Royal Society B, 371, 20150237. https://doi.org/10.1098/rstb.2015.0237 Stringer, C. B., & Buck, L. T. (2014). Diagnosing Homo sapiens in the fossil record. Annals of Human Biology, 41, 312–322. Tishkoff, S. A., Reed, F. A., Friedlaender, F. R., Ehret, C., Ranciaro, A., Froment, A., et al. (2009). The genetic structure and history of Africans and African Americans. Science, 324, 1035–1044. Trinkaus, E. (2005). Early modern humans. Annual Review of Anthropology, 34, 207–230. Veeramah, K. R., & Hammer, M. F. (2014). The impact of whole-genome sequencing on the reconstruction of human population history. Nature Reviews Genetics, 15, 149–162. Veeramah, K. R., Wegmann, D., Woerner, A., Mendez, F. L., Watkins, J. C., Destro-Bisol, G., et al. (2012). An early divergence of KhoeSan ancestors from those of other modern humans is supported by an ABC-based analysis of autosomal resequencing data. Molecular Biology and Evolution, 29, 617–630. Vermeersch, P. M. (2002). Two Upper Palaeolithic burials at Nazlet Khater. In P. M. Vermeersch (Ed.), Palaeolithic quarrying sites in upper and middle Egypt (pp. 273–282). Leuven University Press. Vidal, C. M., Lane, C. S., Asrat, A., Barford, D. N., Mark, D. F., Tomlinson, E. L., et al. (2022). Age of the oldest known Homo sapiens from eastern Africa. Nature, 418, 579–583. Vincente, M., & Schlebusch, C. M. (2020). African population history: An ancient DNA perspective. Current Opinion in Genetics & Development, 62, 8–15. Weaver, T. D. (2012). Did a discrete event 200,000–100,000 years ago produce modern humans? Journal of Human Evolution, 63, 121–126. Williams, S. M., Sirugo, G., & Tishkoff, S. A. (2021). Embracing African genetic diversity. Med, 2, 19–20.

Part I Background and History

Chapter 2

Lost and Found: The Discovery and Rediscovery of the Hofmeyr Skull Alan G. Morris

Abstract The discovery of the Hofmeyr Skull reminds us that chance is the biggest factor in the uncovering of the fossil evidence for human evolution. Critical also is the presence of a network of interested amateurs who are, in turn, connected to local institutions where discoveries can be evaluated. The Hofmeyr skull was presented to the East London Museum in January 1956 where it triggered a series of investigations undertaken in the 1960s and 1970s. Unfortunately, without a date, these early studies did not lead to published results. During this time the skull suffered significant damage on three different occasions and was ultimately transferred from the East London Museum to the Port Elizabeth Museum, at which point it was miscataloged and effectively lost. The skull was rediscovered in the Port Elizabeth Museum in the 1980s and reunited with it original accession data. New methods of analysis have not only shown that the specimen is an important human fossil but have finally allowed the specimen to be successfully dated. The discovery and ‘rediscovery’ of the Hofmeyr skull shows how important careful museum curation is in the field of palaeoanthropology, and this is of special importance in the event that new analytical technologies can be applied in the future.

 

 



Keywords Eastern Cape Province East London Museum Marjorie Courtney-Latimer Raymond A Dart Boskop skull Optically-stimulated Luminescence

A. G. Morris (&) Department of Human Biology, University of Cape Town, Observatory, Cape Town, 7700, South Africa e-mail: [email protected]

Introduction Serendipity is perhaps the most powerful tool in the discovery of the fossil remains of human fossils. We tend to forget in these days of expeditions run by well-funded research institutions, that initial discoveries often relied on the chance finding of fossils by interested amateurs. Such was very much the case when a mineralised skull encased in red Karoo soil was brought to Marjorie Courtenay-Latimer at the East London Museum in January 1956. For generations, the stock farmers of the Karoo drylands of the Eastern Cape in South Africa had divided the land so that each of their sons could inherit a farm on his own. By the 1940s the land was stretched beyond its carrying capacity and the grazing animals had stripped the veld of its vegetation. The soil began to erode and soon the beds of the rivers became deep erosion channels as they funnelled the valuable topsoil south toward the Indian Ocean. The government of the day had to take drastic measures, and after expropriating all of the farms on the higher ground, it began a programme of construction of anti-erosion weirs in the riverbeds. It was in one of these deep erosion gullies that Chris Hattingh noticed the fossilised skull now known as the Hofmeyr cranium.

The Discovery Sometime in 1954, Mr. Hattingh was building an anti-erosion dam on the Vlekpoort River on his farm Klipdrift about 15 kms east of the village of Hofmeyr. He noticed the fossil skull lying near the riverbed at the bottom of an erosion channel at least six metres deep. Hattingh was not sure what to do with his discovery, but eventually he brought it to the attention of a friend, Mr. S. L. Moorcroft. It was Moorcroft who suggested the donation to Marjorie Courtney-Latimer of the East London Museum on the 6th of

© Springer Nature Switzerland AG 2022 F. E. Grine (ed.), Hofmeyr: A Late Pleistocene Human Skull from South Africa, Vertebrate Paleobiology and Paleoanthropology, https://doi.org/10.1007/978-3-031-07426-4_2

9

10

January 1956. There is some debate around the exact year of the discovery, as Hattingh stated in his letter to Courtenay-Latimer in early 1956 that he had found the skull about 18 months before,1 but in an oral statement to the archaeologist H. J. Deacon when Deacon visited the site in 1964, Hattingh indicated that the discovery was much earlier, circa 1952, and that the skull had spent some years in his garage because his wife would not allow him to bring it into the house (Kaye, 1965). The 1954/55 date perhaps seems more likely as it would have been fresher in his memory in 1956. Hattingh found no other bones or artefacts, but because the river was flowing strongly at the time, he knew that the skull had not travelled far from its original location.1 The skull was filled with sandy red alluvium. The left side of the braincase and some of the base of the skull had been broken away and lost, but the back part of the mandible and the first two cervical vertebrae were still attached in the position they would have been when the individual died and were now held in position by the hardened red soil. In the 1950s, the East London Museum and its Director, Courtenay-Latimer, had already gained a reputation for scientific discovery and it is no surprise that Moorcroft brought the skull there. The East London Museum was launched in 1931 with the young Courtenay-Latimer as its sole curatorial employee, a situation that would continue for 15 years (Bruton, 2019; Jewett, 2004). For the first few years the only collections were of some birds’ eggs, a few archaeological flints and some Xhosa cultural items originally collected by Courtenay-Latimer’s mother. This soon changed as Marjorie Courtenay-Latimer began to collect her own material. This included a range of specimens from a three-month sojourn on Bird Island off the coast of Port Elizabeth and a nearly complete skeleton of the mammallike reptile Kannemeyeria excavated with the help of two friends at Tarkastad in the Karoo in 1933 (Bursey, 2004). But the event which was to bring both the East London Museum and Marjorie Courtenay-Latimer to world fame was the discovery of Latimeria chalumnae in December 1938. The story of the identification of the first living coelacanth by Courtenay-Latimer is well known (Bruton, 2017, 2019; Smith, 1956; Thomson, 1991), but the key is the fact that Courtenay-Latimer was able to recognise the critical anatomical features of a living fish that linked it to a form previously seen only in fossil specimens. This was the same ability that she now used to recognise the antiquity and probable scientific importance of the human cranium from Hofmeyr. Courtenay-Latimer knew that if the skull was old, it would be important. She also recognised that its value would

1

The Archival Sources are located at the end of the text.

A. G. Morris

be limited unless there was more knowledge about the context of its discovery. To accomplish this, she travelled to Hofmeyr in August 1956 to see if she could find any more of the specimen. She was shown the donga where the skull had been found, but other than confirming the depth of erosion; she could find no other evidence.2 Despite her disappointment in failing to locate any more of the fossil, she decided that the specimen needed to be seen by a palaeontologist. While she was visiting Johannesburg for a museum conference, she brought the discovery of the skull to the attention of Professor Raymond Dart. Dart was keen to examine the skull, but he was unable to travel to East London and requested that the specimen be sent to him in Johannesburg. Courtenay-Latimer was very concerned about the possibility of the skull being damaged in transit and Dart had to reassure her that it was common practice to transfer specimens between museums.3 It was at this point that the skull suffered the first of several damaging events. During her absence from East London, the ornithologist at the museum “took it upon himself to chip away a lot of the matric [sic] and out into one of the frontal bones”.4 This resulted in significant damage to the lateral edges of the supraorbital margins on both sides of the frontal bone, but especially on the right side (Fig. 2.1). Courtenay-Latimer was furious and “dispensed with his services on account of this sort of thing and lack of cooperation.”4 Despite this setback, Dart encouraged her to send the specimen to him even though she continued to fear more damage to the specimen during transit. Dart emphasised that “if your specimen should turn out to be a near relative [of the newly discovered fossils from Makapansgat] it would be most exciting.”5 In the same letter he encouraged her to revisit the site to confirm the locality and provide pictures or diagrams of the site for a future publication. When the opportunity arose Courtenay-Latimer again travelled to the farm where the fossil had been recovered, but the success of the anti-erosion programme which had begun in the 1950s had buried the actual site of the discovery under meters of new sediment by then.6 The specimen was finally hand-delivered to Dart by M. D. W. Jeffreys in July 1961, and although Dart was willing to continue with the analysis, he was obviously disappointed that it was not an australopithecine. “It seems to belong to Homo sapiens, so do not expect anything extraordinary from its development which will be undertaken as soon as possible.”7

The First Analyses The skull’s arrival in Johannesburg coincided with a series of important discoveries at Makapansgat in the then northern Transvaal (Dart, 1962) and at Olduvai Gorge in Tanzania

2

Discovery and Rediscovery of Hofmeyr

11

Fig. 2.1 Detail of right supraorbital margin showing the damage caused during a presumed attempt to remove matrix from the orbit (Photo credit, F. E. Grine)

(Tobias, 1965), all dating from what was at the time considered the earliest era of human evolution at least two million years before present. Dart and his successor, Phillip Tobias, were focused on the study of these very early fossils and both thought the Hofmeyr specimen was too recent to be of exceptional importance. Despite this, Dart took the time to make some extensive notes on the skull including a set of measurements.8 His notes are undated, but it seems likely that they were made shortly after the skull arrived in Johannesburg, as a further letter he wrote to CourtenayLatimer in August 1961 reflects similar conclusions to what he recorded in his notes.9 Dart’s informal notes focussed on the relationship of Hofmeyr to specimens that had been accepted by Dart as “Boskopoid.” The concept of a large-headed extinct race was based on the cranium recovered from the site of Boskop in 1913 (Dart, 1923). While this concept was being seriously questioned by the early 1960s (Singer, 1958), Dart continued to focus on it as a reference point in his analysis of the Hofmeyr skull. The gist of his notes was a comparison of the Hofmeyr cranium to each of the possible ancient Boskop specimens, especially Tuinplaas (Springbok Flats), Border Cave and the original Boskop skull itself. He also compared the skull to Kabwe (“Rhodesian Man”) and Florisbad, both from contexts that were pre-Holocene and possibly very ancient. None of these specimens had been dated at the time

but all were accepted as “pre-Negro” or “pre-Bushman”, terms accepted by most authors at the time. The accepted model that Dart subscribed to in his writings assumed that the ancestors of both groups in South Africa had been robust and archaic in form and had undergone a progressive reduction in size and, in the case of the Khoesan, the acquisition of pedomorphic characters (Brothwell, 1963; Tobias, 1978). The Boskop type was viewed by him as being intermediate between the robust archaics and the modern Khoesan. Dart found that Hofmeyr did not fit very well with the Boskop specimens on his list, but he was determined to find some kind of anatomical association with earlier fossil specimens. In Dart’s opinion, the well-developed supra-orbital region and large mastoid were similar to Kabwe and Florisbad, but the skull did not seem to be as “primitive” (robust) as these latter specimens. Dart concluded his notes with the sentence: “There is no doubt from the dimensions of the cranium that we are dealing here with a large, massively built sapient human type of South African skull. The principal points at issue therefore are its racial relationships and its possible antiquity”.8 Although he shared his assessment with Courtenay-Latimer, Dart did not publish his analysis. Instead, the specimen was given to Mr. Keith Kaye for his B. Sc. (Med) Honours project under the supervision of Phillip Tobias, who was Dart’s replacement as Head of the

12

Department at the University of the Witwatersrand. Kaye’s project (Kaye, 1965) differed in approach from Dart’s analysis by not focussing on Boskop, and instead on looking at the blending types which were thought at the time to make up living South African aboriginal populations. This approach was very much influenced by Tobias’ thinking, especially in relation to the morphological origins of the Khoekhoe, or “Hottentots” as they were referred to at the time. In Tobias’ view, the physical composition of the living “Hottentots” was a mixture of four ancestral strains: the Boskop type, the Bush type, the Gerontomorphic (sometimes called Australoid) type and the Kakamas (East African Hamitic) type (Tobias, 1955). This perspective was echoed by Kaye in his project. Kaye looked at a wide range of crania from living and archaeological populations, including many of the same specimens referred to by Dart, but he also included specimens from North and East Africa. His analysis, consistent with the methods of the time, examined specific physical features as markers of hybridising types, and not surprisingly his study found signs of each of the four elements which comprised the “Hottentot” physical type. His conclusion was “that in spite of the fact that the skull does not belong to any population group living today, it appears to be of a Hottentot stock, belonging most likely to the Gonaqua who became extinct in the early Nineteenth century” (Kaye, 1965: 37). Dart had been pressing to have a professional archaeologist visit the site ever since his initial correspondence with Courtenay-Latimer. He asked Revil Mason, his colleague at Wits, to find someone who could visit the site.10 Dart’s efforts, through Mason, were finally successful and in September 1964, Hilary Deacon of the Albany Museum in Grahamstown, visited the site and provided Kaye with a thorough report on what he found. Sadly, what he found was relatively little. The location of the discovery was silted up and was “just mud and reeds now” (Kaye, 1965: 3). Deacon did find an eland horn core of similar mineralisation about a hundred metres upstream of the discovery site and at the depth of “4 or 5 feet”. He donated this to the Department of Anatomy for possible future dating attempts. The academic plan was for Kaye to continue this project at the Masters level, but he decided to return to his medical studies in 1966 and did no further work on the topic. Dart still felt that there was more to learn from a detailed study of the Hofmeyr specimen and early in 1966 he approached Hertha de Villiers, a lecturer in the Department of Anatomy, to continue the work. de Villiers wrote to Courtenay-Latimer asking permission to continue the study and also to request consent to remove a sample of bone for dating.11 With Courtenay-Latimer’s agreement, she sent the eland horn core and a human cervical vertebra to Oakley at the British Museum of Natural History for 14C analysis.12 She had

A. G. Morris

carefully removed a vertebral fragment from the adhering matrix for analysis but did not sample the skull itself. de Villiers waited three years for the laboratory in London to analyse the samples, but on discovering in early 1970 that the samples had not yet been processed, she requested their return from London. She then submitted them instead to the CSIR radiocarbon laboratory in Pretoria.13 While she waited for the results from Pretoria, de Villiers felt that it would be best if the skull was sent back to its repository in East London as she had completed her measurements and observations. She acted on this on the 2nd of June 1970. Courtenay-Latimer responded two days later telling de Villiers that the skull had been damaged in transit and that “only half” was left and “the maxilla bones were broken up completely”.14 A flurry of correspondence between them confirmed that the packing in Johannesburg had been inadequate, but that whoever unpacked the specimen in East London did not attempt to save any fragments for reconstruction. The damage not only included the anterior of the maxilla with all of its teeth mesial to the first molars, but also the squamous occipital, the tip of the mastoid and the gonial angle of attached right half of the mandibular ramus. At the same time as the heated correspondence was occurring between de Villiers and Courtenay-Latimer, Hilary Deacon wrote to confirm that he was unable to provide any further information about the site. He wrote that he understood “the circumstances of the find and its value, [but] I don’t think there is any particular associated problem that would be answered by radiometric dating.15 The results from Pretoria on the dating of the eland horn core came back later in the year giving a date of 3,020 ± 90 (Pta 261) years before present (Vogel & Marais, 1971). The vertebral fragment did not produce a date. At best, if the association between the horn core and skull could be accepted, the skull was indeed relatively recent, but if the association was not accepted, then the skull itself remained undated. Deacon did agree that the skull had a study value and that his division at the Albany Museum would be prepared to store the specimen for the long term if this was a problem for the East London Museum. With no date and no archaeological association, the now damaged specimen was no longer of interest to South African scientists. Courtenay-Latimer retired from the museum in 1973 and in February 1975, the collection of human archaeological skeletons at the East London Museum was exchanged for ornithological material from the collection of the Port Elizabeth Museum. The East London Museum had no one on staff with archaeological experience but did have a strength in ornithology, while the reverse was true in Port Elizabeth. In the process the Hofmeyr skull was miscataloged (R. M. Tietz, pers. comm. to A. G. Morris, 2 July 1990). The records at Port Elizabeth incorrectly labelled

2

Discovery and Rediscovery of Hofmeyr

13

the skull as a coastal find from the Qulu River Mouth, about 20 kms west of East London.

The Rediscovery The Hofmeyr skull, now mislabelled as Qulu, was seen by the German paleoanthropologist Günter Bräuer in 1980. Bräuer was intrigued by the “very robust and heavily mineralized skull” but could find no further information about the site. He arranged for a section of the parietal bone to be forwarded to Reiner Protsch in Frankfurt in order to assess whether there was sufficient nitrogen content for radiocarbon dating. The bone fragment was sent to Frankfurt in September 1980, but Protsch did not follow up with a date and Bräuer assumed that the sample possessed too little collagen for analysis (G. Bräuer, pers. comm to A. G. Morris, 9 March 1999). This was now the third sequence of damage to the cranium adding the loss of a significant piece of the left parietal to the previous damage to the frontal processes, the maxillary tooth row, and the cranial base (Fig. 2.2). Courtenay-Latimer, who was now retired and living in Port Elizabeth, visited the museum in July 1987 and complained to Nancy Tietz, her replacement as director of the East London Museum, that the Hofmeyr skull was now incorrectly labelled. Tietz mentioned this to the Director of the Port Elizabeth Museum who asked the anthropologist Francis Thackeray to have a look at the specimen. Although Thackeray could not confirm its identity, he noted that it had similarities to the ancient Florisbad specimen on which he had worked previously (J. F. Thackeray, pers. comm. to A. G. Morris, 27 January 1989). In his correspondence, Thackeray provided a brief report on the skull in which he noted that “the matrix has been analysed at the Geology Department of the University of Port Elizabeth, at the request of Dr. Graham Ross … of the Port Elizabeth Museum …. The matrix has been described as a riverine sand.” Thackeray then notified the current author of the location of the specimen and its questionable provenance. The skull’s identity as Hofmeyr rather than Qulu was later recognized from photographs by Morris, and this was confirmed from the original transfer records at the East London Museum by Tietz (R. M. Tietz, pers. comm. to A. G. Morris, 2 July 1990). The rediscovery of the Hofmeyr skull did not immediately solve any of the issues raised by Dart and Kaye in their analyses. Ultimately the problem was that the specimen remained undated. Morris recognised both the robust morphology and the impression of antiquity from the state of mineralisation (as Bräuer and Thackeray had done independently) and asked Tietz if it would be possible to revisit

Fig. 2.2 Condition of the Hofmeyr skull in 1968 before it was sent back to East London from Johannesburg compared to its condition in 1993 (Reproduced from Grine et al., 2007: Fig. 1)

the original excavation site. Tietz brought the request to the attention of Courtney-Latimer who immediately volunteered to join any expedition to locate the original site (Tietz, 2004). The original Hofmeyr location on the farm Klipdrift was visited by Morris, his student Jonathan Kovacs, Carl Vernon (the East London Museum ornithologist), Nancy Tietz (the East London Museum director) and Marjorie CourtenayLatimer, from the 26th to the 29th of October 1992. The team was hosted by the new owner of Klipdrift, Mr. J. Moolman, and housed in the farm homestead. Activities over the three days concentrated on examining the riverbanks to identify any archaeological exposures, but especially to identify the possible discovery location. Five exposures were identified (Fig. 2.3), ranging from Later Stone Age rock art to Middle Stone Age flakes, but none could be associated with the original discovery. The construction of the anti-erosion weir in 1954 had completely silted up the river by 1962 and the site had become overgrown with dense reed-beds by 1992 (Fig. 2.4), a situation

14

unchanged as seen in more recent visits to the site. A series of deep erosion gullies was still visible downstream of the discovery site and these gave an impression of the appearance of the site at the time of the skull’s discovery (Fig. 2.5). Some 100 m downstream of the discovery site was an extensive deflation along the edge of the riverbank (Site 4 on Fig. 3). A heavily mineralised metapodial of a kudu-sized bovid projected from the eroding surface, and a single MSA core and LSA bored stone were seen lying on the erosion surface (Fig. 2.6). There was no association between core, stone or bone. As such, other than confirming the discovery site location, the 1992 field visit did not add any further information about the skull’s context. The Hofmeyr skull was transferred to the Department of Human Biology at the University of Cape Town on the return from the site visit. Morris began a new analysis that compared the cranium to a range of modern populations (rather than single individuals) and it was confirmed that the skull fell within the range of modern populations, but on the edge of their variation. Fred Grine of Stony Brook University was invited to help work on the specimen at this stage and the collaboration resulted in the first publication of the

A. G. Morris

specimen (Morris & Grine, 1999). With the addition of Grine to the team, the techniques of analysis were expanded not only in terms of comparative samples (now including Upper Palaeolithic Europeans), but also in methodology (Grine et al., 2007). Multivariate analysis of linear cranial measurements was explored by Isabelle Ribot and this was followed by 3-D geometric morphometric analysis led by Katarina Harvati. A small fragment of bone detached from the broken left parietal was sent by Grine to the Oxford Radiocarbon Accelerator Unit for possible accelerator mass spectrometry (AMS) dating, but the lab result identified that too little collagen was preserved for this method to succeed (Grine et al., 2007). Grine therefore proposed that the matrix infill of the cranial vault be used as a dating resource as opposed to using methods that would continue to damage the bone itself. In 2003, he carefully cleared out the vault infill under ultraviolet light in the photographic dark room in the Department of Vertebrate Palaeontology at the Iziko Museum in Cape Town. The samples obtained under these conditions were submitted to the Research Laboratory for Archaeology and the History of Art Luminescence Dating

Fig. 2.3 Map of locations visited during the October 1992 field trip to the Hofmeyr area

2

Discovery and Rediscovery of Hofmeyr

15

Fig. 2.4 The anti-erosion weir at the Hofmeyr discovery site in January, 2005 (Photo credit, A. G. Morris)

Laboratory at the University of Oxford in the hope of obtaining an Optically Stimulated Luminescence (OSL) date. This method effectively dates the length of time the infill sediment has been resident in the skull. The method provided the breakthrough which finally provided an estimated date of just over 36,000 years and has allowed for more meaningful interpretations of its morphology (Grine et al., 2007).

Hofmeyr and Changing Methods in the Study of Human Origins Dart’s analysis in the 1960s was strongly coloured by the 19th and early twentieth century concept of racial typology (Morris, 2012). His first description of the skull was consequently couched in terms of racial types and ancestral strains in which statistical data and ranges of variation were missing. Kaye’s approach (under the influence of Tobias)

was only marginally better because a wider range of specimens was compared to the single skull from Hofmeyr, but his study was also hobbled by the need to base the analysis on imaginary types rather than on analyses of statistical data. The advent of multivariate statistics has allowed individual specimens to be compared in terms of morphological distance. The technique relates the appearance of the Hofmeyr skull to the wide range of variation seen in comparable Upper Palaeolithic specimens and has enabled a more holistic analysis of shape (Grine et al., 2007). The most recent analyses have used a morphometric approach that employs three-dimensional locations of standardised biometric points which produce a more nuanced interpretation of morphological differences. There has also been an expansion of new dating techniques in parallel to changes to the techniques of morphological analysis. Hofmeyr was discovered just when the first radiocarbon analyses were starting to produce reliable results, but the lack of preservation of collagen in the Hofmeyr bone prevented a successful application of this

16

A. G. Morris

Fig. 2.5 Erosion of the Vlekpoort riverbank downstream of the Hofmeyr site in October, 1992 (Photo credit, A. G. Morris)

technique to the specimen. In particular, Dart compared Hofmeyr to the crania from Kabwe and Florisbad, but these specimens are now considered to be far older than Hofmeyr. Kabwe’s faunal association links it to a period between 150,000 and 300,000 years ago (Klein, 1973), and Grün et al. (2020) have recently obtained an average U-series date of 299 ± 25 ka for the cranium, but an age of only 216 ± 25 ka for dentine from teeth embedded in it. Florisbad appears to be around 250,000 years old (Grün et al., 1996b). Both are related in some manner to Homo sapiens but are too ancient to be directly linked to Hofmeyr. Specimens from the Middle Stone Age which fall near the divergence of modern from archaic African forms are now dated to circa 100,000 to 150,000 years ago (Lombard et al., 2013). Of these, only the Border Cave cranium is in a similar state of preservation to Hofmeyr, but confirmed dates are only available for the fragmentary mandibles from the site (Grün & Beaumont, 2001; Grün et al., 1996a) and there is serious doubt about the cranium’s association with the Middle Stone Ages layers. It is certainly possible that the skull from Border Cave is a Later Stone Age intrusion (Sillen & Morris, 1996). The robust Tuinplaas (Springbok Flats) specimen is now thought to be less than 20,000 and perhaps even less than 11,000 years old (Pike et al., 2004). Fish Hoek, which was thought to be at least 12,000 years old

is now known to be only about 6,000 (Stynder et al., 2009). The original Boskop skull has no clear archaeological association and remains undated. All of this reminds us that robust ‘non-Khoesan’ morphology is not enough to confer antiquity by itself and that the morphological patterning of these Late Pleistocene (and earliest Holocene) specimens does not reflect a simple path from archaic to derived anatomy over time. Evolutionary interpretation remains at best speculative for fossil specimens without a secure date. The similarity of the Hofmeyr morphology to the early modern populations of Europe (Grine et al., 2007) has highlighted the disjuncture between morphological and genetic studies of regional southern African peoples. Living Khoesan populations demonstrate very ancient roots (Schlebusch et al., 2017) and there have been at least two genetic divisions earlier than 35,000 years ago that suggest that the major linguistic groups in Khoesan languages reflect divergent evolutionary events (Schlebusch et al., 2012). Yet the Hofmeyr cranium shows no specific similarity to any living Khoesan population. Instead, it presents a more “basal” morphology that extends from southern Africa through to North Africa (Crevecoeur et al., 2009) even though genetic evidence suggests that these groups were already diverging by the time of Hofmeyr at 36,000 years ago (Lombard et al., 2013; Schlebusch et al., 2017). The implication is that

2

Discovery and Rediscovery of Hofmeyr

Fig. 2.6 Site #4 on the Vlekpoort River as identified in Fig. 3. Exposure is in a deflation. recorded during the October 1992 field trip to the Hofmeyr area. a Core and bored stone, b heavily mineralised bovid metapodial in situ (Photo credit, A. G. Morris)

lineages as tracked by genetic markers in living peoples have been moving at a different tempo from the adaptive morphology seen in cranial features (Morris, 2005). This lesson of variation in tempo seen in different aspects of evolutionary events is critical to our understanding of human evolution as neither genetics nor morphology alone gives a definitive interpretation of evolutionary events. Perhaps the most important observation that can be drawn from the post-discovery history of the Hofmeyr skull is the value of careful curation of old museum specimens even when their provenance is uncertain. New methodologies unknown at the time of discovery can, like with Hofmeyr, not only provide information about time and context but can provide the key to expanding our broader knowledge about the evolution of the human species. Source Notes: 1. Letter from C. Hattingh to M. Courtenay-Latimer, 25 January, 1956. East London Museum Correspondence File (hereafter ELMCF): 8.2.3/H7: Hofmeyr Skull.

17

2. M. Courtenay-Latimer Diary, 28 August 1956. East London Museum: MCL Collection: No. 16. 3. Letter from R.A. Dart to M. Courtenay-Latimer, 22 June 1960. ELMCF: 8.2.3/H7: Hofmeyr Skull. 4. Letter from M. Courtenay-Latimer to R.A. Dart, 21 March 1961. ELMCF: 8.2.3/H7: Hofmeyr Skull. 5. Letter from R.A. Dart to M. Courtenay-Latimer, 2 June 1961. ELMCF: 8.2.3/H7: Hofmeyr Skull. 6. Letter from M. Courtenay-Latimer to R.A. Dart, 5 June 1962. ELMCF: 8.2.3/H7: Hofmeyr Skull. 7. Letter from R.A Dart to M. Courtenay-Latimer, 12 July 1961. ELMCF: 8.2.3/H7: Hofmeyr Skull. 8. Notes on the Hofmeyr skull in R.A. Dart’s handwriting. No date. ELMCF: 8.2.3/H7: Hofmeyr Skull. 9. Letter from R.A. Dart to M. Courtenay-Latimer, 5 August 1962. ELMCF: 8.2.3/H7: Hofmeyr Skull. 10. Letter from R. Mason to R.A. Dart, 25 November 1963. ELMCF: 8.2.3/H7: Hofmeyr Skull. 11. Letter from H. de Villiers to M. Courtenay-Latimer, 23 March 1966. ELMCF: 8.2.3/H7: Hofmeyr Skull. 12. Letter from H. de Villiers to M. Courtenay-Latimer, 22 March 1967. ELMCF: 8.2.3/H7: Hofmeyr Skull. 13. Letter from H. de Villiers to M. Courtenay-Latimer, 11 May 1970. ELMCF: 8.2.3/H7: Hofmeyr Skull. 14. Letter from M. Courtenay-Latimer to H. de Villiers, 4 June 1970. ELMCF: 8.2.3/H7: Hofmeyr Skull. 15. Letter from H. Deacon to M. Courtenay-Latimer, 7 June 1970, ELMCF: 8.2.3/H7: Hofmeyr Skull. Acknowledgements I thank Christa Kuljian and Jason Hemingway for their thorough and thoughtful reviews of the manuscript, and Geraldine Morcom for assistance with referencing the original documents in the East London Museum archives.

References Brothwell, D. R. (1963). Evidence of early population change in central and southern Africa: Doubts and problems. Man, 63, 101–104. Bruton, M. (2017). The annotated old fourlegs: The story of the coelacanth. Struik, Cape Town. Bruton, M. (2019). Curator and crusader: The life and work of Marjorie Courtney-Latimer. Footprint Press Southern Africa. Bursey, M. (2004). Obituary. The Coelacanth, 42, 12–16. Crevecoeur, I., Rougier, H., Grine, F., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: Evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Dart, R. A. (1923). Boskop remains from the South East African coast. Nature, 112, 623–625. Dart, R. A. (1962). The Makapansgat pink breccia australopithecine skull. American Journal of Physical Anthropology, 20(2), 119–126. Grine, F. E., Bailey, R. M., Harvati, K., Nathan, R. P., Morris, A. G., Henderson, G. M., et al. (2007). Late Pleistocene human skull from Hofmeyr, South Africa, and modern human origins. Science, 315, 226–229.

18 Grün, R. W., & Beaumont, P. B. (2001). Border Cave revisited: A revised ESR chronology. Journal of Human Evolution, 40, 467– 482. Grün, R., Beaumont, P., Tobias, P. V., & Eggins, S. (1996a). On the age of the Border Cave 5 mandible. Journal of Human Evolution, 45, 155–167. Grün, R., Brink, J. S., Spooner, N. A., Tayor, L., Stringer, C. B., Franciscus, R. G., et al. (1996b). Direct dating of Florisbad hominid. Nature, 382, 500–501. Grün, R., Pike, A., McDermott, F., Eggins, S., Mortimer, G., Aubert, M., et al. (2020). Dating the skull from Broken Hill, Zambia, and its position in human evolution. Nature, 580, 372–375. Jewett, S. L. (2004). Marjorie Courtenay-Latimer: More than the coelacanth. The Coelacanth, 42, 17–28. Kaye, K. W. (1965). Report on a fossil human skull from Hofmeyr, Eastern Cape. B.Sc.(Med) honours thesis, University of the Witwatersrand. Klein, R. G. (1973). Geological antiquity of Rhodesian Man. Nature, 244, 311–312. Lombard, M., Schlebusch, C., & Soodyall, H. (2013). Bridging disciplines to better elucidate the evolution of early Homo sapiens in southern Africa. South African Journal of Science, 109, 1–8. Morris, A. G. (2005). Prehistory in blood and bone: An essay on the reconstruction of the past from genetics and morphology. Transactions of the Royal Society of South Africa, 60, 111–114. Morris, A. G. (2012). Biological anthropology at the southern tip of Africa: Carrying European baggage in an African context. Current Anthropology, 53, S152–S160. Morris, A. G., & Grine, F. E. (1999). Hofmeyr and the origin of anatomically modern South Africans. Journal of Human Evolution, 36, A15. Pike, A. W. G., Eggins, S., Grün, R., & Thackeray, F. (2004). U-series dating of TP1, an almost complete human skeleton from Tuinplaas

A. G. Morris (Springbok Flats), South Africa. South African Journal of Science, 100, 381–383. Schlebusch, C. M., Skoglund, P., Sjödin, P., Gattepaille, L. M., Hernandez, D., Jay, F., et al. (2012). Genomic variation in seven Khoe-San groups reveals adaptation and complex African history. Science, 338, 374–379. Schlebusch, C. M., Malmström, H., Günther, T., Sjödin, P., Coutinho, A., Edlund, H., et al. (2017). Southern African ancient genomes estimate modern human divergence to 350 000 to 260 000 years ago. Science, 358, 652–655. Sillen, A., & Morris, A. G. (1996). Diagenesis of bone from Border Cave: Implications for the age of the Border Cave hominids. Journal of Human Evolution, 31, 499–506. Singer, R. (1958). The Boskop “race” problem. Man, 58, 173–178. Smith, J. L. B. (1956). Old fourlegs: The story of the coelacanth. Longman Green. Stynder, D. D., Brock, F., Sealy, J., Wurz, S., Morris, A. G., & Volman, T. (2009). A mid-Holocene AMS 14C date for the presumed Late Pleistocene Peers Cave human skeleton from South Africa. Journal of Human Evolution, 56, 431–434. Thomson, K. S. (1991). Living fossil: The story of the coelacanth. Norton. Tietz, R. M. (2004). Field trips and friendship: A memoire of Marjorie Courtenay-Latimer. The Coelacanth, 42, 37–47. Tobias, P. V. (1955). Physical anthropology and somatic origins of the Hottentots. African Studies, 14, 1–22. Tobias, P. V. (1965). Early man in East Africa. Nature, 149, 22–33. Tobias, P. V. (1978). The San: An evolutionary perspective. In P. V. Tobias (Ed.), The Bushmen: San hunters and herders of southern Africa (pp. 16–32). Human & Rousseau. Vogel, J. C., & Marais, M. (1971). Pretoria radiocarbon dates I. Radiocarbon, 13, 378–394.

Chapter 3

Genetic Divergence Within Southern Africa During the Later Stone Age Dana R. Al-Hindi, Austin W. Reynolds, and Brenna M. Henn

Abstract Southern Africa is currently home to a wide array of diverse ethnic groups and ancestries resulting from agriculturalist and pastoralist expansions, colonialism, slave trade and modern global trade networks. The vast majority of this change occurred in the last 2,000 years, however. Prior to these migrations, southern Africa was populated by hundreds of hunter-gatherer populations which can be loosely grouped as “Khoe-San”. Many publications refer these groups as Khoisan, San, Basawara, Khoekhoe, Bushmen, Coloured, etc., terms which we unpack in greater detail. Here, our goal is to situate the Hofmeyr skull in the context of the genetic diversity present in southern Africa during the Later Stone Age. The primary source of information for this review is DNA samples from contemporary populations distributed from Angola to South Africa. Over the past decade, we have seen a remarkable surge in DNA data generated from genome-wide autosomal loci, i.e. “genomic” data. These new genomic datasets have been used to estimate the earliest population divergence between Khoe-San groups and other populations. Linguistic data is often used to inform genetic sampling projects; thus, we provide a brief introduction to linguistic diversity across southern Africa. We describe how geneticists have identified population structure among Khoe-San groups, how this ancestry is structured across the landscape, the time depth of population structure, and changes in population size through time. We aim to elucidate common misconceptions of genetic data (such as the idea that population diverge at the same time as genetic lineages) and how/if genetic dates can be related to archaeological dates.

D. R. Al-Hindi (&)  A. W. Reynolds  B. M. Henn (&) Department of Anthropology, Davis Genome Center, University of California, Davis, CA 95616, USA e-mail: [email protected] A. W. Reynolds Department of Anthropology, Baylor University, Waco, TX 76706, USA





Keywords Khoe-San Divergence mtDNA Y-chromosome



Coalescence



Introduction The Cradle of Modern Humans South Africa is famously home to the “Cradle of Humankind” World Heritage site, referring to the Sterkfontein cave system which has produced numerous Australopithecine fossils and, more recently, Homo naledi skeletons from the Rising Star Cave. The prevailing view among many geneticists is that southern Africa (broadly speaking) is also the cradle of modern humans. By this we mean that the majority of the genetic ancestry in modern human populations has its source in an ancestral population which lived in southern Africa. This ancestral population is the last time that all present-day humans can trace back their common ancestors which contributed genetic material to future generations, and the deepest human lineages are among indigenous southern African populations (Bergström et al., 2021; Fan et al., 2019; Lorente-Galdos et al., 2019; Mallick et al., 2016; Schlebusch et al., 2012). Consider an example for contrast: the 1.5% of Neandertal DNA identified in Eurasian/Oceanian populations is neither the majority of their genomic ancestry nor is it shared by all human populations. It is important to consider the implications of this definition because it does not a priori exclude other regions of Africa or even other hominin species from contributing to subsequent modern human evolution. Based on the primacy of fossil evidence, eastern Africa has typically been considered the locus of origin for humans (Campbell & Tishkoff, 2008). However, in just the past decade, genomic evidence is nearly unequivocal in demonstrating that the deepest population divergence is between the Khoe-San and all other human groups. Eastern Africans simply do not carry the highest levels of genetic diversity

© Springer Nature Switzerland AG 2022 F. E. Grine (ed.), Hofmeyr: A Late Pleistocene Human Skull from South Africa, Vertebrate Paleobiology and Paleoanthropology, https://doi.org/10.1007/978-3-031-07426-4_3

19

20

nor the deepest divergent lineages among sampled populations. The absolute date of this Khoe-San divergence varies widely depending on the types of loci included, the mutation rate, and the method of analysis. Almost all dates range from 200 to 100 ka with a mode around 150 ka-120 ka, which corresponds roughly the marine isotope stage (MIS) 5 to MIS 6 transition (Knight et al., 2003; Poznik et al., 2013; Henn et al., 2018). The assumption in most tested genetic models is that human populations approximate an “isolationmigration” model of bifurcating population divergence which led to the modern genetic landscape. However, we note that more complex models of population reticulation or isolation by distance have not been widely tested (Chan et al., 2019; Henn et al., 2011, 2018; Scerri et al., 2018; Schlebusch et al., 2017; Skoglund et al., 2017). With nearly a dozen genetic publications having examined the date of divergence with broadly congruent results, the focus has turned to how assumptions about how our methods influence the results (i.e., model misspecification, reference bias, ghost populations, isolation by distance) and how calibrated rates accord with paleoanthropological discoveries.

Click-Speaking Hunter-Gatherers Within Africa While Africa encompasses a considerable amount of genetic diversity, it also harbors great deal of linguistic diversity, exhibiting a similar pattern of (phoneme) diversity decreasing with distance away from central and southern Africa under a founder effect model (Atkinson, 2011). This is particularly useful when unraveling population demographic histories. Both biological and social interaction between populations is necessary for linguistic and genetic change to arise, leading language and genetic trees to have a strong correlation (Cavalli-Sforza et al., 1992). This correlation has become a useful tool to genetic anthropologists. The linguistic classification of an individual (or their parents/grandparents) is often used as a proxy to help identify populations to sample. There are four major linguistic phyla within Africa – Nilo-Saharan, Niger-Kordofanian, Afroasiatic, and Khoisan (Greenberg, 1963). It is worth mentioning that there have been exchanges of views regarding the nomenclature of the Khoisan language family and the populations, where “Khoisan” is often used interchangeably. Schultze (1928) combined Khoi (now Khoe), meaning ‘people’ among Nama pastoralists, and “San” meaning ‘forager’, to collectively reference hunter-gatherer and pastoralist groups that reside within southern Africa. Since then, it has been borrowed by linguists to refer to their languages. The combined ethnonym has left some contemporary groups uncomfortable and

D. R. Al-Hindi et al.

culturally indistinguishable when referred to jointly (Schlebusch, 2010). Mainly, members of the South African San community have asked to remain distinct from their Khoespeaking neighbors and to be referred to by their specific ethnic identities or generally as the San (Schlebusch, 2010). To preserve their request, this paper will reference the populations cumulatively as Khoe-San, separate than that of the Khoisan languages, and to acknowledge their difference in identity through hyphenation. The Khoisan linguistic phylum is unique for its many and various click phonemes (Greenberg, 1963; Hutchison et al., 2001; Güldemann, 2014). These click-incorporating languages are only found within eastern and southern Africa, with the exception of Damin, a language that has independently established click-consonants in northern Australia (Güldemann & Stoneking, 2008). Five independent language families have been historically classified under Khoisan: Tuu, Kx’a, and Khoe-Kwadi from southern Africa, and Hadza and Sandawe from Tanzania (Güldemann, 2014; Güldemann & Stoneking, 2008). The Khoisan language map is illustrated in Fig. 3.1. The Dahalo people in Kenya use click consonants but speak a Cushitic language that likely adopted click phonemes through past exposure with click-speaking populations (although an independent innovation of click consonants has not been ruled out) (Güldemann & Stoneking, 2008). While the Khoe-San are closely related groups that reside relatively near one another, language families vary across populations and are not necessarily consistent with subsistence strategies. For example, the Taa and G||ana are primarily hunter-gatherers but speak Tuu and Khoe-Kwadi languages, respectively. The Nama prove to be another example, as they speak a KhoeKwadi language despite being genetically more similar to their hunter-gatherer Tuu-speaking 6¼ Khomani San and the Karretjie people (Pickrell et al., 2012; Schlebusch et al., 2012; Uren et al., 2016) (Fig. 3.1).

Substructure Among the Khoe-San The dispersal of the various subsistence and language practices have led researchers to question what factors have driven the distribution exhibited today. Investigating the genetic relationships between the Khoe-San groups has provided an interesting explanation for group distribution. Previous studies using genotyped single nucleotide polymorphisms (SNPs) have found that the Khoe-San populations separate into two clusters that correlate with geography, stretching in a northwest and southeast fashion through the Kalahari (Pickrell et al., 2012; Schlebusch et al., 2012). Non-genetic variables such as language and subsistence strategies do not correlate as well with the ancestry distribution (Pickrell et al., 2012), albeit genetic component

3

Genetic Divergence Within Southern Africa During …

21

Fig. 3.1 Khoisan language map. Languages are in bold, black text and language families are defined by color (Khoeid in red, !Ui-Tuu in blue, Kx’a in orange, Hadza in green, and the Sandawe in pink). Sizes of data points correspond to the estimated number of speakers in the past generation (including partial speakers), based on interview data and linguistic surveys. The map was originally created by Monika Feinen using data collected by Matthias Brenzinger at the Centre for African Language Diversity (CALDi) (Swingler, 2017). It has since then been updated by M. Feinen using new ethnographic data of reported speakers stretching from one generation to the present

predictions have suggested language plays a complicated role in the divide (Schlebusch et al., 2012). More recent studies found the Khoe-San populations can be further divided into three ancestries (Montinaro et al., 2017; Uren et al., 2016), complicating the geographical correlation previously observed (Pickrell et al., 2012; Schlebush et al., 2012) beyond a northeast and southwest pattern. The authors proposed that the population substructure was a result of the ecological boundaries (Uren et al., 2016). While an isolation-by-distance model can explain a 2-dimensional cline (Montinaro et al., 2017; Pickrell et al., 2012; Schlebusch et al., 2012; Vicente et al., 2019), it does not explain the circum-Kalahari signature (Fig. 3.2). There, the three ancestry signals are referred to as central Kalahari, northern Kalahari, and circum-Kalahari – corresponding to the rim of the Kalahari (Uren et al., 2016) (Fig. 3.2). The circum-Kalahari ancestry is comprised of all regional Khoisan language families (Uren et al., 2016). Though it’s

prominently present in the Shua, Khwe, !Xun, and Hai||om, its highest frequency is found in the Nama, and 6¼ Khomani (N|uu-speakers). Communities with majority central Kalahari ancestry include the Taa (North, East, and West), = Hoan, Naro, Shua and Kua, G|ui, and G||ana (mixture of Tuu and Kx’a speakers). The remaining populations are highly composed of the northern Kalahari ancestry – Ju/’hoansi (North and South), Khwe, and !Xun. An uneven distribution of autochthonous mtDNA haplogroups L0d and L0k reinforce the ancestry clusters. While both haplogroups are commonly found among the northern Kalahari ancestry, L0k is generally absent from the circum-Kalahari ancestry (such as the Nama, Karretjie, 6¼ Khomani) and is present at low frequencies in the central Kalahari ancestry (Schlebusch et al., 2011, 2013; Uren et al., 2016). This is within the same timeframe many of the mtDNA and Y-chromosome Khoe-San haplogroups coalesce. Divergence within the autochthonous mtDNA L0d and L0k

22

D. R. Al-Hindi et al.

Fig. 3.2 Five spatially distinct ancestries indicate deep population structure in southern Africa. The five most common ancestries in contemporary southern Africa populations are shown separately in A–E. Black dots represent the sampling location of populations in southern Africa. The third dimension in each map (depth of color) represents the mean ancestry proportion for each group for a given k ancestry, calculated from ADMIXTURE using unrelated individuals, and indicated in the color key as 0–100% for each k ancestries. Reprinted with permission from Uren et al. (2016)

haplogroups and Y-chromosome A2, A3b1, and B2b lineages have been estimated to take place during the MIS 3–4 (Barbieri et al., 2013, 2016; Chan et al., 2019; Pickrell et al., 2012; Schlebusch et al., 2011; Tishkoff et al., 2007; Wood et al., 2005). The L0d1 branch is the most frequent KhoeSan mtDNA haplogroup (Barbieri et al., 2013). L0d1 is found in all Khoe-San populations and splits into two branches (L0d1a and L0d1c) that have been reported to coalesce by Barbieri et al. (2013) and Chan et al. (2019) at ~55 kya and ~47 kya, respectively. Similarly, L0d3a and L0d3b have varying coalescent dates to L0d3, previously reported at ~45 kya (Barbieri et al., 2013) and more recently ~60 kya (Chan et al., 2019). Within L0d, L0d2 is more common than L0d3 and is found in frequencies higher than 10% in the Naro (25%), G||ana (13%), Hai||om (27%),

and Nama (34%). In Tuu-speakers from eastern Botswana, L0d2 is found as high as ~53% in the population. Furthermore, the coalescence dates of Y chromosomal haplogroup B2b (~45–60 kya) and those of the most common Khoe-San haplogroup A lineages (A2 at ~30kya and A3b1 at ~55kya) are among the oldest divergences in the human Y chromosome tree (Knight et al., 2003; Scozzari et al., 2012; Poznik et al., 2013; Barbieri et al., 2016). Low frequencies of the L0d branch are also found in eastern Africa (Behar et al., 2008; Chan et al., 2019). Previous studies hypothesized that Khoe-San populations inhabited eastern Africa prior to southern Africa, and the present-day distribution of L0 lineages could be explained by loss in drift and recent migrations into the region (Gonder et al., 2006). More recent work using 2,330-year-old aDNA

3

Genetic Divergence Within Southern Africa During …

samples predating the introduction of pastoralism in southern Africa explains the various L0 haplogroups in southern and eastern Africa as a result of an ancient population’s migration event out of modern-day Botswana starting ~130 kya (Chan et al., 2019). In this scenario, harsh climatic conditions during MIS 6 restricted an ancestral population to a refuge around palaeo-lake Makgadikgadi, until green corridors opened toward the northeast and southwest. Population migration in both directions resulted in the mtDNA distribution of L0 haplogroups observed today. Additional ancient genomewide analyses including southern African samples ranging between 2,000 and 1,200-years-old have suggested an ancestral forager population wide-spread between eastern and southern Africa, based on signals ascribed as Khoe-San ancestry in ancient samples from eastern Africa (Skoglund et al., 2017; Wang et al., 2020). However, the linguistic affinity between eastern and southern Khoisan speakers is still debated (Sands 1998; Güldemann & Vossen, 2000). It should be stressed that genetic divergence predates population divergence. Gene genealogies diverge when a new mutation occurs in a lineage, such as a new mutation in the mitochondrial genome. Both the ancestral and the derived mutation may continue to segregate in a population for some time. Whereas, population divergence occurs when gene flow between two subpopulations ends, ceasing to homogenize allele frequencies. In the presence of minimal gene flow, subpopulations are subject to genetic drift separately from one another, allowing for allele frequencies to fluctuate such that, given enough time, distinguish ancestry clusters. Notice in Fig. 3.3, although genetic divergence (DG) is observed in the ancestral population, population divergence (DP) does not take place until later. This concept is important when synthesizing dates across disciplines, specifically between geneticists and paleoanthropologists. Of particular importance is building context around genetic divergence estimates within the KhoeSan alongside the available fossil record from this time period (i.e., the Hofmeyr skull) (Fig. 3.3). The Khoe-San ancestry clusters and estimated divergence time have often been based on Wright’s FST, a commonly used statistic that describes genetic differentiation, or structure, between populations under the assumption of neutrality (Box 3.1, Eq. 3.2) (Holsinger & Weir, 2009). The timing of the divergence within Khoe-San using autosomal FST has been estimated to date back to ~25kya – 43kya, during MIS 3 (Pickrell et al., 2012; Schlebusch et al., 2012). Genetic differentiation between the Khoe-San ancestry clusters are

23

Fig. 3.3 Genetic divergence versus population divergence Adapted from Kimura (2015: Fig. 3.2). Colored circles (red, purple, blue, and green) represent mutations on each haplotype. Each split along the lineage marks a point of genetic divergence (DG). Population divergence (DP) after genetic divergence, highlighted by the dashed line. While the populations share mutations (red and purple), private mutations (green and blue) along haplotypes differentiate them from one another. The four haplotypes coalesce and share a most recent common ancestor (MRCA) marked by the star. To clearly visualize divergence and the MRCA, the star is positioned slightly over the initial split in the lineage

moderate (Wright’s FST ~0.012–0.034) (Schlebusch et al., 2012), though these values can vary widely (~0.011 – 0.055) (Uren et al., 2016). FST values less than 0.05 and between 0.05 to 0.15 are usually considered low and moderate, respectively. These values can be seen between populations such as the Tuscans from Italy and South Asians (0.03) or the Han Chinese and South Asians (0.07) (International HapMap 3 Consortium et al., 2010). Although, given that Khoe-San who reside within the same region and have similar subsistence, these values are substantial (Box 3.1).

24

D. R. Al-Hindi et al.

Box 3.1 Mean pairwise difference, denoted by p, is calculated based on single nucleotide variation between haplotypes randomly selected from a single population. This equation uses the frequencies of haplotypes i and j in the population, the nucleotide differences between i and j, and the summation of these three values multiplied together for all pairs of haplotypes to describe genetic diversity. X p¼ xi xj pij ð3:1Þ ij

FST, a measure of genetic differentiation, is frequently used to describe the genetic structure between subpopulations (Eq. 3.2). Where pt is the average pairwise difference among all subpopulations, and ps is the average pairwise difference within a single subpopulation. FST ¼ ðpt  ps Þ = pt

ð3:2Þ

Various factors can increase or decrease the nucleotide diversity of a population, including mutation, genetic drift and gene flow. Populations that are genetically more similar to one another will have an FST value closer to 0, while more variation will result in a value closer to 1. In practice, values of 1 are not observed, as no human population carries the exact opposite set of alleles compared to another. The highest FST values observed between human populations would at most be around 0.30. The !Kung San and Karitiana from Brazil, for instance, are about as geographically separate as two populations can be and have an FST of about 0.25 (Ramachandran et al., 2005). Measurements of FST have a strong relationship with geography, where FST increases with increased distance between the populations (Ramachandran et al., 2005). Much of this stems from continental or insular barriers that cause population divergence and limit gene flow. While FST is useful to describe genetic structure between or within populations, it can also be used to calculate genetic divergence times (Eq. 3.3). This equation is related to coalescent theory, a method that geneticists use to calculate the probability two alleles of the same gene coalesce one or more generations back in time. When doing so, it is necessary to take into account the effective population size, Ne, because genetic drift has a stronger effect on smaller populations than it does with larger populations. The effective population size is the number of individuals in an

idealized (Wright-Fisher) population that has the same magnitude of genetic drift as the actual population. This number is usually smaller than that of the census population size because not every individual will bare offspring. In diploid organisms, both sets of chromosomes need to be accounted for when estimating time of coalescence, thus the effective population is doubled, 2Ne.   1 t ð3:3Þ 1  FST ¼ 1  2Ne

The northern Kalahari Ju/’hoansi have a relatively high FST score between the central Kalahari ancestry groups Shua (0.054), Kua (0.044), and G||ana (0.036) (Uren et al., 2016). Interestingly, the Khwe cluster with the Ju/’hoansi within the northern Kalahari ancestry (Schlebusch et al., 2012; Uren et al., 2016) but share an FST score (0.055) comparable to that of the Ju’/hoansi and the Shua. Their distinction within the northern Kalahari cluster has been suggested to be a result of gene flow with Bantu-speaking populations within the region (Schlebusch et al., 2012). By contrast, the Nama have the greatest genetic differentiation with the Ju/’hoansi (0.038) and east Taa (0.025) who have high northern Kalahari and central-Kalahari ancestries, respectively. The Nama, who previously clustered with the southern San (Schlebusch et al., 2012), are central Khoisan language speakers and pastoralists who are genetically more similar to the 6¼ Khomani (0.003) despite the difference in subsistence strategies. While the Damara are a Khoisan-speaking population, they are genetically distinct from other Khoe-San groups, and it is assumed that their language was recently adopted from the Nama (Pickrell et al., 2012; Schlebusch et al., 2012; Uren et al., 2016). The divergence patterns and distribution of ancestries within the Khoe-San has previously been suggested to be a result of climate change within southern Africa between the Middle to Later Stone Age (MSA and LSA) (Behar et al., 2008; Henn et al., 2011). The diversification of archaeological artifacts during the MSA and LSA also suggests population fragmentation (Robbins et al., 2016; Mackay et al., 2014). Mackay and colleagues (2014) hypothesized that wetter and warmer climates (i.e., MIS 5 and MIS 3) promoted population dispersal into more environments, causing an increase in diversification within the archaeological record (Fig. 3.4). Conversely, cold and arid climatic conditions (i.e., MIS 4 and MIS 2) are envisioned as having caused populations to coalesce leading to shared attributes between artifacts. Though initially proposed to describe the fragmentation and coalescence patterns observed in the

3

Genetic Divergence Within Southern Africa During …

25

Fig. 3.4 Southern Africa’s population coalescence and fragmentation conceptually depicted from MIS 4 to MIS 2. This figure illustrates a possible scenario in which population fragmentation leads to strong population differentiation among southern African Khoe-San. During MIS 4, populations (in teal) may have been sequestered along the southwestern coastline of southern Africa. Climate change and relative warming of MIS 3 promoted populations to expand and disperse inland. As the climate gradually cooled into the next glacial period, MIS 2, populations become isolated, restricted by ecological boundaries such as an arid Kalahari. Depiction here does not include the Paleo-Agulhas Plain off the Cape which would have extended out an additional 90 km and was a productive zone for coastal foraging (Marean, 2010). Artistic contribution: Anna Goldfield

archaeological data of southern Africa, this hypothesis could be applied to the Khoe-San genetic data (Fig. 3.3). Given that the divergence among the Khoe-San ancestries date back to generally ~30 kya, it is plausible the warm climate during MIS 3 encouraged population dispersal into environments that may not have been previously as favorable; although contradicting hypotheses regarding the influence interglacial periods have on the social landscape have been proposed (see Marean et al., 2020). During the harsher climate conditions of MIS 2, populations may have been geographically constricted by difficult to traverse ecological boundaries, thus limiting gene flow and leading to the varying ancestries found within the Khoe-San today (see Fig. 3.3). While the Khoe-San ancestry distributions across the landscape have commonalities that may support the hypothesis put forward by Mackay et al. (2014), further collaborative research and modeling will be necessary to examine how uniformly applicable this model is. The Hofmeyr Skull falls directly in the time frame of increasing genetic diversity. This fossil dates back to ~36kya (Grine et al., 2007) and has different morphological characteristics than those found in regional contemporary populations (Crevecoeur et al., 2009). As it is currently the most complete human cranium from the sub-Saharan Africa during this time frame (Grine, 2016; McBrearty & Brooks, 2000),

understanding its behavioral, cultural, and environmental context will paint a more thorough image of southern Africa’s prehistory.

Late Holocene Gene Flow Other lineages commonly found in contemporary Khoe-San groups are likely due to recent demographic processes such as the Bantu expansion and European colonization. Haplogroup B2a (M150) and many of the lineages in haplogroup E, for instance, are associated with the Bantu migration (Berniell-Lee et al., 2009). However, one branch of haplogroup E (E3b1f-M293) makes up a significant proportion of the E lineages found Khoe-San groups in southern and eastern Africa and is consistent with the spread of pastoralism into southern Africa ~2,000 years ago (Henn et al., 2008). These recent demographic processes have resulted in large variation in the Y chromosome haplogroup frequencies in Khoe-San populations. The Ju/’hoansi have the highest frequency (~90%) of the characteristic Y-chromosomal haplogroups A and B, followed by the Nama (~70%), 6¼ Khomani San (~60%) and !Xuun (~60%) (Henn et al., 2011; Soodyall et al., 2008; Wood et al., 2005). The lowest

26

levels of KhoeSan Y-chromosomal haplogroups are found in the Damara, Khwe, and the Karretjie people (10–15% each, Cruciani et al., 2004; Tishkoff 2007).

Future Directions Future interdisciplinary research is necessary to further unravel southern Africa’s prehistory, though this is not without challenges. Different mediums, methods, and varying definitions across disciplines have given rise to opposing conclusions, further complicating interdisciplinary discussions (Pargeter et al., 2016; Stewart et al., 2020). While in the field of population genetics, populations are defined by differentiation in allele frequencies, in archaeology they are often defined by site and location. This can entail a difference in the null hypotheses related to population movement versus continuity during the Late Pleistocene. Archaeologists may perceive hunter-gathers as mobile, non-sedentary groups. The extent of that mobility can vary, but the null hypothesis still presumes that cultural continuity does not imply population continuity. Rarely are archaeologists confident that the assemblage at hand is from a single episode, but rather a time-averaged palimpsest. That is, an observed archaeological site typically depicts an average behavioral trend over generations rather than single groups. On the other hand, geneticists generally begin with a null hypothesis that, while people may migrate (at varying rates), meta-populations as a whole are not moving across large landscapes during prehistoric conditions unless they are expanding into unpopulated environments. That being said, a common critique of geneticists is the scarcity of hypothesis testing. The trans-disciplinary contrast in definitions seems natural given the different mediums and scales used between the fields. For example, mtDNA demographic modeling by Powell et al. (2009) showed patterns indicative of population growth within southern Africa during the late Pleistocene, though the archaeological record came to contrasting conclusions (Klein & Steele, 2013). The paleoanthropological field is also subject to specific constraints. The acidity of southern Africa’s soil has left the paleontological record slim, handing paleoanthropologists and archaeologists little preserved organic material. The Hofmeyr skull serves as a glimpse of past modern human morphology but is the only nearly complete sub-Saharan cranium from its time period and region that can do so (Grine, 2016). Reconciling differences across disciplines is crucial if we wish to elevate our anthropological understanding of modern humans in Africa during the Later Stone Age, and how the Hofmeyr skull fits within the archaeological and genetic data. More broadly, it is imperative to address these

D. R. Al-Hindi et al.

differences to build a more in-depth comprehension of human evolution. With gaps in the tangible record and the various resources and methods used across disciplines, complimentary and collaborations approaches tackling similar questions is a necessity. Acknowledgements We extend our appreciation to all the African participants that have contributed to the genetic studies reported in this paper. We thank Fred Grine for the invitation to be a part of this volume. Furthermore, we are grateful to Dr. Fransesco Montinaro, Dr. Teresa Steele, Sara Watson, and an anonymous reviewer for their comments. Lastly, we would like to acknowledge and thank Dr. Monika Feinen for her efforts in updating the Khoisan language map, and Dr. Anna Goldfield for creating our population coalescence and fragmentation figure. We thank the Genetics Society of America for granting permission to re-print Fig. 3.2.

References Atkinson, Q. D. (2011). Phonemic diversity supports a serial founder effect model of language expansion from Africa. Science, 332, 346– 349. https://doi.org/10.1126/science.1199295 Barbieri, C., Hübner, A., Macholdt, E., Ni, S., Lippold, S., Schröder, R., et al. (2016). Refining the Y chromosome phylogeny with southern African sequences. Human Genetics, 135, 541–553. https://doi.org/10.1007/s00439-016-1651-0 Barbieri, C., Vicente, M., Rocha, J., Mpoloka, S. W., Stoneking, M., & Pakendorf, B. (2013). Ancient substructure in early mtDNA lineages of southern Africa. American Journal of Human Genetics, 92, 285–292. https://doi.org/10.1016/j.ajhg.2012.12.010 Behar, D. M., Villems, R., Soodyall, H., Blue-Smith, J., Pereira, L., Metspalu, E., et al. (2008). The dawn of human matrilineal diversity. American Journal of Human Genetics, 82, 1130–1140. https://doi.org/10.1016/j.ajhg.2008.04.002 Bergström, A., Stringer, C., Hajdinjak, M., Scerri, E. M. L., & Skoglund, P. (2021). Origins of modern human ancestry. Nature, 590, 229–237. https://doi.org/10.1038/s41586-021-03244-5 Campbell, M. C., & Tishkoff, S. A. (2008). African genetic diversity: Implications for human demographic history, modern human origins, and complex disease mapping. Annual Review of Genomics and Human Genetics, 9, 403–433. https://doi.org/10.1146/annurev. genom.9.081307.164258 Cavalli-Sforza, L. L., Minch, E., & Mountain, J. L. (1992). Coevolution of genes and languages revisited. Proceedings of the National Academy of Sciences of the United States of America, 89, 5620– 5624. https://doi.org/10.1073/pnas.89.12.5620 Chan, E. K. F., Timmermann, A., Baldi, B. F., Moore, A. E., Lyons, R. J., Lee, S. S., et al. (2019). Human origins in a southern African palaeo-wetland and first migrations. Nature, 575, 185–189. https:// doi.org/10.1038/s41586-019-1714-1 Crevecoeur, I., Rougier, H., Grine, F., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: Evidence from Nazlet Khater, Peştera cu Oase, and Hofmeyr. American Journal of Physical Anthropology, 140, 347– 358. https://doi.org/10.1002/ajpa.21080 Cruciani, F., Fratta, R. L., Santolamazza, P., Sellitto, D., Pascone, R., Moral, P., et al. (2004). Phylogeographic analysis of haplogroup E3b (E-M215) Y chromosomes reveals multiple migratory events within and out of Africa. American Journal of Human Genetics, 74, 1014–1022. https://doi.org/10.1086/386294 Fan, S., Kelly, D. E., Beltrame, M. H., Hansen, M. E. B., Mallick, S., Ranciaro, A., et al. (2019). African evolutionary history inferred

3

Genetic Divergence Within Southern Africa During …

from whole genome sequence data of 44 indigenous African populations. Genome Biology, 20, 82. https://doi.org/10.1186/ s13059-019-1679-2 Gonder, M. K., Mortensen, H. M., Reed, F. A., de Sousa, A., & Tishkoff, S. A. (2006). Whole mtDNA genome sequence analysis of ancient African lineages. Molecular Biology and Evolution, 24, 757–768. https://doi.org/10.1093/molbev/msl209 Greenberg, J. H. (1963). The languages of Africa. Indiana University Press. Grine, F. E., Bailey, R. M., Harvati, K., Nathan, R. P., Morris, A. G., Henderson, G. M., et al. (2007). Late Pleistocene human Skull from Hofmeyr, South Africa, and modern human origins. Science, 315, 226–229. https://doi.org/10.1126/science.1136294 Grine, F. E. (2016). The late quaternary hominins of Africa: The skeletal evidence from MIS 6–2. In S. C. Jones & B. A. Stewart (Eds.), Africa from MIS 6–2: Population dynamics and paleoenvironments (pp. 323–381). Springer. Güldemann, T., & Fehn, A. M. (Eds.). (2014). Beyond ‘Khoisan.’ Historical relations in the Kalahari Basin. John Benjamins, Amsterdam. https://doi.org/10.1075/cilt.330 Güldemann, T., & Stoneking, M. (2008). A historical appraisal of clicks: A linguistic and genetic population perspective. Annual Review of Anthropology, 37, 93–109. https://doi.org/10.1146/ annurev.anthro.37.081407.085109 Güldemann, T., & Vossen, R. (2000). Khoisan. In B. Heine & D. Nurse (Eds.), African languages: An introduction (pp. 99–122). Cambridge University Press. Henn, B. M., Gignoux, C. R., Jobin, M., Granka, J. M., Macpherson, J. M., Kidd, J. M., et al. (2011). Hunter-gatherer genomic diversity suggests a southern African origin for modern humans. Proceedings of the National Academy of Sciences of the United States of America, 108, 5154–5162. https://doi.org/10.1073/pnas. 1017511108 Henn, B. M., Steele, T. E., & Weaver, T. D. (2018). Clarifying distinct models of modern human origins in Africa. Current Opinion in Genetics & Development, 53, 148–156. https://doi.org/10.1016/j. gde.2018.10.003 Holsinger, K. E., & Weir, B. S. (2009). Genetics in geographically structured populations: Defining, estimating and interpreting FST. Nature Reviews Genetics, 10, 639–650. https://doi.org/10.1038/ nrg2611 Hutchison, J., Heine, B., & Nurse, D. (Eds.) (2001). African languages: An introduction. International Journal of African Historical Studies, 34, 240–241. https://doi.org/10.2307/3097354 International HapMap 3 Consortium, Altshuler, D. M., Gibbs, R. A., Peltonen, L., Altshuler, D. M., Gibbs, R. A., et al. (2010) Integrating common and rare genetic variation in diverse human populations. Nature 467, 52–58. https://doi.org/10.1038/nature09298 Kimura, R. (2015). Human migrations and adaptations in Asia inferred from genomic diversity. In Y. Kaifu, M. Izuho, T. Goebel, H. Sato, & A. Ono (Eds.), Emergence and diversity of modern human behavior in Paleolithic Asia (pp. 34–50). Texas A&M University Press. Klein, R. G., & Steele, T. E. (2013). Archaeological shellfish size and later human evolution in Africa. Proceedings of the National Academy of Sciences of the United States of America, 110, 10910– 10915. https://doi.org/10.1073/pnas.1304750110 Knight, A., Underhill, P. A., Mortensen, H. M., Zhivotovsky, L. A., Lin, A. A., Henn, B. M., et al. (2003). African Y chromosome and mtDNA divergence provides insight into the history of click languages. Current Biology, 13, 464–473. https://doi.org/10.1016/ s0960-9822(03)00130-1 Lorente-Galdos, B., Lao, O., Serra-Vidal, G., Santpere, G., Kuderna, L. F. K., Arauna, L. R., et al. (2019). Whole-genome sequence analysis of a Pan African set of samples reveals archaic gene flow from an

27 extinct basal population of modern humans into sub-Saharan populations. Genome Biology, 20, 77. https://doi.org/10.1186/ s13059-019-1684-5 Mackay, A., Stewart, B. A., & Chase, B. M. (2014). Coalescence and fragmentation in the late Pleistocene archaeology of southernmost Africa. Journal of Human Evolution, 72, 26–51. https://doi.org/10. 1016/j.jhevol.2014.03.003 Mallick, S., Li, H., Lipson, M., Mathieson, I., Gymrek, M., Racimo, F., et al. (2016). The simons genome diversity project: 300 genomes from 142 diverse populations. Nature, 538, 201–206. https://doi. org/10.1038/nature18964 Marean, C. W. (2010). Pinnacle point cave 13B (Western Cape Province, South Africa) in context: The Cape Floral Kingdom, shellfish, and modern human origins. Journal of Human Evolution, 59, 425–443. https://doi.org/10.1016/j.jhevol.2010.07.011 Marean, C. W., Cowling, R. M., & Franklin, J. (2020). The PalaeoAgulhas Plain: Temporal and spatial variation in an extraordinary extinct ecosystem of the Pleistocene of the Cape Floristic Region. Quaternary Science Reviews, 235, 106161. https://doi.org/10.1016/ j.quascirev.2019.106161 Mcbrearty, S., & Brooks, A. S. (2000). The revolution that wasn’t: A new interpretation of the origin of modern human behavior. Journal of Human Evolution, 39, 453–563. https://doi.org/10.1006/jhev. 2000.0435 Montinaro, F., Busby, G. B. J., Gonzalez-Santos, M., Oosthuitzen, O., Oosthuitzen, E., Anagnostou, P., et al. (2017). Complex ancient genetic structure and cultural transitions in southern African populations. Genetics, 205, 303–316. https://doi.org/10.1534/ genetics.116.189209 Pargeter, J., MacKay, A., Mitchell, P., Shea, J., & Stewart, B. A. (2016). Primordialism and the ‘Pleistocene San’ of southern Africa. Antiquity, 90, 1072–1079. https://doi.org/10.15184/aqy.2016.100 Pickrell, J. K., Patterson, N., Barbieri, C., Berthold, F., Gerlach, L., Güldemann, T., et al. (2012). The genetic prehistory of southern Africa. Nature Communications, 3, 1143. https://doi.org/10.1038/ ncomms2140 Powell, A., Shennan, S., & Thomas, M. G. (2009). Late Pleistocene demography and the appearance of modern human behavior. Science, 324, 1298–1301. https://doi.org/10.1126/science.1170165 Poznik, G. D., Henn, B. M., Yee, M. C., Sliwerska, E., Euskirchen, G. M., Lin, A. A., et al. (2013). Sequencing Y chromosomes resolves discrepancy in time to common ancestor of males versus females. Science, 341, 562–565. https://doi.org/10.1126/science.1237619 Ramachandran, S., Deshpande, O., Roseman, C. C., Rosenberg, N. A., Feldman, M. W., & Cavalli-Sforza, L. L. (2005). Support from the relationship of genetic and geographic distance in human populations for a serial founder effect originating in Africa. Proceedings of the National Academy of Sciences of the United States of America, 102, 15942–15947. https://doi.org/10.1073/pnas.0507611102 Scerri, E. M. L., Thomas, M. G., Manica, A., Gunz, P., Stock, J. T., Stringer, C., et al. (2018). Did our species evolve in subdivided populations across Africa, and why does it matter? Trends in Ecology & Evolution, 33, 582–594. https://doi.org/10.1016/j.tree. 2018.05.005 Schlebusch, C. (2010). Issues raised by use of ethnic-group names in genome study. Nature, 464, 487–487. https://doi.org/10.1038/ 464487a Schlebusch, C. M., de Jongh, M., & Soodyall, H. (2011). Different contributions of ancient mitochondrial and Y-chromosomal lineages in ‘Karretjie people’ of the Great Karoo in South Africa. Journal of Human Genetics, 569, 623–630. https://doi.org/10.1038/jhg.2011. 71 Schlebusch, C. M., Malmström, H., Günther, T., Sjödin, P., Coutinho, A., Edlund, H., et al. (2017). Southern African ancient genomes estimate modern human divergence to 350,000 to 260,000 years

28 ago. Science, 358, 652–655. https://doi.org/10.1126/science. aao6266 Schlebusch, C. M., Skoglund, P., Sjödin, P., Gattepaille, L. M., Hernandez, D., Jay, F., et al. (2012). Genomic variation in seven Khoe-San groups reveals adaptation and complex African history. Science, 338, 374–379. https://doi.org/10.1126/science.1227721 Schultze, L. (1928). Zur Kenntnis des Körpers der Hottentotten und Buschmänner. Zoologische und anthropologische Ergebnisse einer Forschungsreise im westlichen und zentralen Sudafrika, 5, 147– 227. Scozzari, R., Massaia, A., D’Atanasio, E., Myres, N. M., Perego, U. A., Trombetta, B., et al. (2012). Molecular dissection of the basal clades in the human Y chromosome phylogenetic tree. PLoS One, 7, e49170. https://doi.org/10.1371/journal.pone.0049170 Skoglund, P., Thompson, J. C., Prendergast, M. E., Mittnik, A., Sirak, K., Hajdinjak, M., et al. (2017). Reconstructing prehistoric African population structure. Cell, 171, 59-71.e21. https://doi.org/10.1016/j. cell.2017.08.049 Soodyall, H., Makkan, H., Haycock, P., & Naidoo, T. (2008). The genetic prehistory of the Khoe and San. Southern African Humanities, 20, 37–48. Stewart, B. A., Zhao, Y., Mitchell, P. J., Dewar, G., Gleason, J. D., & Blum, J. D. (2020). Ostrich eggshell bead strontium isotopes reveal persistent macroscale social networking across late Quaternary southern Africa. Proceedings of the National Academy of Sciences of the United States of America, 117, 6453–6462. https://doi.org/10. 1073/pnas.1921037117

D. R. Al-Hindi et al. Swingler, H. (2017). Passion for languages, compassion for linguistic communities. University of Cape Town News. Retrieved May 08, 2017, from https://www.news.uct.ac.za/article/-2017-05-08passion-for-languages-compassion-for-linguistic-communities Tishkoff, S. A., Gonder, M. K., Henn, B. M., Mortensen, H., Knight, A., Gignoux, C., et al. (2007). History of click-speaking populations of Africa inferred from mtDNA and Y chromosome genetic variation. Molecular Biology and Evolution, 24, 2180–2195. https://doi.org/10.1093/molbev/msm155 Uren, C., Kim, M., Martin, A. R., Bobo, D., Gignoux, C. R., & van Helden, P. D. (2016). Fine-scale human population structure in southern Africa reflects ecogeographic boundaries. Genetics, 204, 303–314. https://doi.org/10.1534/genetics.116.187369 Vicente, M., Jakobsson, M., Ebbesen, P., & Schlebusch, C. M. (2019). Genetic affinities among southern Africa hunter-gatherers and the impact of admixing farmer and herder populations. Molecular Biology and Evolution, 36, 1849–1861. https://doi.org/10.1093/ molbev/msz089 Wang, K., Goldstein, S., Bleasdale, M., Clist, B., Bostoen, K., BakwaLufu, P., et al. (2020). Ancient genomes reveal complex patterns of population movement, interaction, and replacement in sub-Saharan Africa. Science Advances, 6, eaaz0183. https://doi.org/10.1126/ sciadv.aaz0183 Wood, E. T., Stover, D. A., Ehret, C., Destro-Bisol, G., Spedini, G., McLeod, H., et al. (2005). Contrasting patterns of Y chromosome and mtDNA variation in Africa: Evidence for sex-biased demographic processes. European Journal of Human Genetics, 13, 867– 876. https://doi.org/10.1038/sj.ejhg.5201408

Chapter 4

Geological Setting of the Hofmeyr Locality Johann Neveling

Abstract The Hofmeyr skull was discovered from the banks of the Vlekpoort River, which traverses a broad plain northeast of the town of Hofmeyr, located in the hinterland of the Eastern Cape Province of South Africa. The fossil locality is directly underlain by the reddish siltstones of the Burgersdorp Formation, while the resistant sandstones of the overlying Molteno Formation form the slopes of the Bamboesberg mountains more than 10 km to the east. Both these formations accumulated through deposition by northwards-flowing rivers draining the Cape Fold Belt to the south, during the Triassic (251–201 Ma), filling the main Karoo Basin, which was at that time situated in southern Gondwana. The modern landscape, dominated by the Great Escarpment, was formed during a long, erosion-dominant period, which started right after Gondwana break-up, was accelerated during the predominant humid tropical climate associated with the Cretaceous, and continues to this day. Short, steep south-flowing river systems, such as the Great Fish River drainage, which encompass the Vlekpoort River, drains the landscape across and south of the Great Escarpment. Quaternary deposits drape the hill slopes and plain north of Hofmeyr. The Late Pleistocene age obtained from the Hofmeyr specimen overlaps with the fluvial record from central South Africa, which is a time when a dry climate is assumed to have predominated over central South Africa. Locally the Vlekpoort River valley is bounded by resistant Karoo bedrock, with a prominent dolerite dyke approximately 4 km south of the fossil locality, one of several to intersect the modern landscape, controlling the local channel profile. Over the past century the construction of numerous dams and weirs in the river channel has interrupted the

J. Neveling (&) Geoscience Mapping Unit, Council for Geoscience, Private Bag X112, Pretoria, 0001, South Africa e-mail: [email protected]

natural river evolution processes by raising the local base level, which has resulted in in the silting-up of the channel floor and the Hofmeyr fossil locality.











Keywords Hofmeyr Late Pleistocene Riverton ForVlekpoort River Great Fish River mation Bamboesberg

Introduction The Hofmeyr cranium was discovered in the early 1950s in the southeastern hinterland of South Africa (Fig. 4.1). No associated fauna or archaeological artefacts were reported with the cranium, but the recent application of modern dating techniques, specifically the combined application of optically stimulated luminescence and uranium-series dating methods, enabled a Late Pleistocene age to be assigned to the specimen (Grine et al., 2007). Currently, very little is known about its geological context and the silting of the local drainage system subsequent to the discovery of the cranium. The purpose of the present contribution is to provide a sound understanding of the geological context of both the Hofmeyr locality and the surrounding landscape.

Geography The Hofmeyr skull was discovered as it eroded from the banks of the Vlekpoort River where it traverses the farm Goedemoed 35, some 17 km east-northeast of the town of Hofmeyr in the Eastern Cape Province of South Africa (Fig. 4.2). This is an area of low topography, located on the eastern edge of a plain referred to as the Springbok Flats (note this is not a unique name in South Africa), but less than 10 km west of the foot of the Bamboesberg mountain range.

© Springer Nature Switzerland AG 2022, corrected publication 2023 F. E. Grine (ed.), Hofmeyr: A Late Pleistocene Human Skull from South Africa, Vertebrate Paleobiology and Paleoanthropology, https://doi.org/10.1007/978-3-031-07426-4_4

29

30

J. Neveling

Fig. 4.1 Satellite image of South Africa showing the position of the Hofmeyr cranium locality (star), and the Great Escarpment (white stippled line) relative to key towns and cities in South Africa. Satellite photo courtesy of NASA (https://visibleearth.nasa.gov)

The lower plains of the Springbok Flats support an upper Karoo vegetation that rapidly transitions to the dry highland grassland of the Bamboesberg range (Mucina & Rutherford, 2011). The latter is a southern outlier of the greater Drakensberg Range and reaches altitudes of 1800– 2000 m, separating the Springbok Flats around Hofmeyr from a high plateau (~1750 m), which in this part of the Eastern Cape Province extends approximately 30 km eastwards. The Hofmeyr locality has an elevation of ~1295 m, which is approximately 60 m lower than the edge of the plain at the foot of the Bamboesberg mountains, 8 km to the east. The intervening plain is drained by numerous small dongas (gulleys) and rivulets that feed into the Vlekpoort River.

Basement Geology The landscape that hosts the Hofmeyr skull locality is controlled by the underlying geology, which has formed over a period of more than 300 million years. The region surrounding the fossil locality is underlain by rocks assigned to

the Karoo Supergroup, a largely sedimentary sequence that covers the central and eastern parts of South Africa’s surface area. At Hofmeyr and the surrounding area, the Karoo Supergroup rocks are the source of all sediments transported and deposited by the drainage systems traversing the modern landscape. The Karoo sequence accumulated in a basin that was one of several that developed throughout the southern part of the supercontinent Gondwana during the late Paleozoic, as the result of the unique combination of tectonic stresses associated with shortening and accretion along the southern margin of the continent, together with extensional stresses propagating southwards from the divergent Tethyan margin (Wopfner, 2002). The largest and most extensive of these was the main Karoo Basin, which is named after the semidesert region in the southern hinterland of South Africa. It formed to the north of the Cape Fold Belt, which borders the southern rim of the continent and is considered to be the source of much of the basin-fill’s sedimentary rocks (Rubidge, 2005). Most workers interpret the Karoo Basin to be a foreland basin (Catuneanu et al., 1998, 2002; Cole, 1992; Johnson, 1991; Johnson et al., 2006; Viglietti et al., 2017a), but alternative interpretations have been proposed

4

Geological Setting of Hofmeyr Locality

31

Fig. 4.2 Hillshade image of the area in the vicinity of the Hofmeyr cranium on the farm Goedemoed 35, northeast of the town of Hofmeyr. The cranium was discovered on the southern bank of the Vlekpoort River which traverses a wide plain and drains the Bamboesberg mountains to the east. Altitude data of key topographic features are based on 1:50,000 topographic maps and Google Earth

(Lindeque et al., 2011; Tankard et al., 2009, 2012; Turner, 1999; Turner & Thomson, 1998). Sediment derived from highlands to the north, and increasingly so with time, from the mountains of the Cape Fold Belt to the south (Fig. 4.3), filled the basin between the Late Carboniferous to Middle Jurassic (Johnson et al., 2006), forming the rock sequence of the Karoo Supergroup, over a period of approximately 120 million years. The current consensus is that the orogenic processes that formed the Cape Fold Belt (interpreted to have formed part of a much larger mountain belt across Gondwana) also provided the loading that formed the main Karoo Basin. As a result, the basin-fill is highly asymmetrical, with a total cumulative thickness of ~12 km proposed for the entire basin-fill in the southeast (Johnson et al., 2006; but see Lindeque et al., 2011; Scheiber-Enslin, 2016) thinning to the north. This trend is accentuated by the underlying crustal structure with thicker stratigraphy recorded where the Karoo Basin is underlain by the Namaqua-Natal Metamorphic Belt in the south, thinning rapidly where the basin crosses the southern edge of the more rigid Kaapvaal Craton in the northern half of the basin (Fig. 4.3). Hofmeyr is located in the deeper, southern part of the basin, but the impact of the crustal geology is best evidenced by the rapid thinning of the

Katberg and Molteno formations near the southern boundary of the Free State Province, approximately 100 km to the north of Hofmeyr.

Karoo Supergroup Differences in sedimentary process and distribution allowed for the recognition of four main sedimentary lithostratigraphic units – the Dwyka, Ecca, Beaufort and Stormberg groups – capped by the Middle Jurassic basalts of the Drakensberg Group (Fig. 4.3). The sedimentary units represent a general shift from deep water deposition under cold conditions during the Late Carboniferous–earliest Permian (Catuneanu et al., 2002, 2005), to fully continental deposition under warmer and arid conditions in the Jurassic (Bordy & Head, 2018). Massive diamictite and mudstones characterize the basal Dwyka Group, which accumulated under glacial conditions. The overlying Ecca Group represents the thickest and aerially most extensive unit of the Karoo Supergroup and demonstrates substantial lateral variation, as reflected by the fact that it hosts 16 formations. In the south it is dominated

32

J. Neveling

Fig. 4.3 Distribution of the key lithostratigraphic units of the Karoo Supergroup. Stippled line represents the southern margin of the Kaapvaal Craton. The red star represents the Hofmeyr locality

by a thick (3000–5000 m) sequence of sandstones and siltstones that accumulated under deep water conditions (Baiyegunhi & Gwavava, 2016; Johnson et al., 2006). Both the Dwyka and Ecca groups underlie the Hofmeyr locality, but the closest outcrops of these rocks occur more than 150 km to the south. The study area is primarily underlain by the rocks of the Beaufort Group, which was formed by fluvial processes draining the Cape Fold Belt to the south (Catuneanu et al., 1998, 2005). It is subdivided into a lower Adelaide Subgroup and upper Tarkastad Subgroup (Johnson, 1976; SACS, 1980), with the former subdivided in the Eastern Cape Province into the Abrahamskraal, Middleton and Balfour formations (Cole et al., 2016; SACS, 1980). The Balfour Formation, which outcrops in the vicinity of the town of Cradock, is a predominantly argillaceous interval (Viglietti et al., 2017b), consisting of stacked second order upwards-fining cycles (Catuneanu & Elango, 2001) of sandstones and greenish and bluish grey mudstones (Johnson, 1976). The Balfour Formation achieves a thickness of approximately 500 m near Cradock (Viglietti et al., 2017b). Further north, the Tarkastad Subgroup underlies the plain around Hofmeyr (Fig. 4.4). The basal unit of the Subgroup,

the Katberg Formation, is defined as an arenaceous unit characterized by the occurrence of thick sandstone bodies. However, in this part of the Karoo Basin the distinction is qualitative and based on a relative increase in sandstone content compared to the underlying Balfour and overlying Burgersdorp formations (Groenewald, 1996; Johnson, 1976; Johnson et al., 2006). The succeeding Burgersdorp Formation is characterized by greyish red siltstones (Fig. 4.5), with subordinate light grey sandstones, arranged in upward-fining cycles a few meters to tens of meters thick (Groenewald, 1996; Johnson & Hiller, 1990). The sandstones are predominantly fine- to medium-grained and greenish to yellowish grey. Individual sandstones grade both vertically and laterally into siltstones, and clasts of silt- and mudstone may be dispersed throughout the sandstones. The siltstones are characterized by greyish red and greenish grey colors. Well-preserved, though non-abundant, remains of fossil reptiles and amphibians have been reported from the rocks of the Burgersdorp Formation exposed in the study area (Kitching, 1995). A minimum thicknesses of 250 m has been reported for the Burgersdorp Formation exposed near Steynsburg, 37 km to the northwest, and in excess of 350 m towards the town of

4

Geological Setting of Hofmeyr Locality

33

Fig. 4.4 Geological map of bedrock geology of the study area. The proposed position of the contact between the Katberg and Burgersdorp formations (dashed line) is based on Cole et al. (2004)

Fig. 4.5 Resistive dolerite dykes typically cap large hills and mountains on the Karoo landscape, such as at the Teebus hill, situated 37 km northwest of the fossil locality. Below the remnant dolerite sill capping the hill, the red siltstones of the Burgersdorp Formation dominate the upper two thirds of the hill with the uppermost part of the sandstone-dominated Katberg Formation exposed on the lower hill slopes

34

Tarkastad, 50 km to the southeast (Groenewald, 1996; Neveling, 2004). The eroding Burgersdorp Formation forms the lower slopes of the mountains and hills immediately to the east of the Hofmeyr locality. On the 3124 Middelburg 1: 250,000 geological map, the contact between the Katberg and Burgersdorp formations is shown to roughly coincide with the R391 road, which runs north-northeast from Hofmeyr. This is, however, difficult to verify in the field due to the gradational nature of the contact and the flat topography. Yet, the map suggests that the Hofmeyr locality is directly underlain by either the uppermost Katberg or lowermost Burgersdorp formations. The Bamboesberg (mountains) that rises ten km to the east of the locality are capped by the pale sandstone and interbedded grey siltstones and mudstones of the Molteno Formation (Cole et al., 2004), which is the lowermost unit of the Stormberg Group (Johnson, 1976; Turner, 1975). An unconformity estimated to represent a depositional hiatus of 4.5–6 million years, separates the uppermost Burgersdorp Formation from the overlying Molteno Formation in this part of the basin (Catuneanu et al., 1998, 2005; Hancox, 1998). It reflects the fact that the Stormberg Group accumulated during a phase of orogenic unloading (Catuneanu et al., 1998). In the field, Cole et al. (2004) equated the base of the Molteno Formation with the appearance of the first fine- to coarse-grained glittering sandstone in the succession, which has been associated with a distinct break in slope. A maximum thickness of 625 m has been reported for the Molteno Formation in the vicinity of the town of Indwe (Christie, 1981), but the formation thins to the west and a thickness of 250 m was reported for it 20–30 km to the north of the Hofmeyr locality, near the town of Steynsburg (Cole et al., 2004; Turner, 1975). Earlier workers subdivided the Molteno Formation into a lowermost Bamboesberg Member and a middle Indwe Sandstone Member, while the upper part of the Formation has been variously subdivided and named by previous authors (Bordy et al., 2005; Christie, 1981, 1986; Hancox, 1998; MacDonald, 1993; Turner, 1975). To date only the Bamboesberg and Indwe members have been formally accepted by the South African Committee for Stratigraphy (SACS, 1980). The Molteno Formation is generally considered to be a sandstone-dominated unit, but exposures of this formation in the field contain a significant proportion of finer grained sedimentary rocks. Near the town of Sterkstroom (50 km to the east), the basal Bamboesberg Member is composed of up to five stacked fining-upward sequences, each of which is composed of laterally extensive medium to fine-grained sandstones, fining into thin lenticular siltstones and, occasionally, coal (Hancox, 1998). These fining-upwards sequences progressively pinch-out to the north and west, with younger sequences offlapping older ones (Hancox,

J. Neveling

1998), so that the Bamboesberg Member thins in these directions (Christie, 1986; Hancox, 1998). Thus, the basal fining-upwards sequence is absent at Groot Doringhoek Pass, more than 20 km north of the Hofmeyr locality, where the Bamboesberg Member reaches a maximum thickness of only 85 m. The Bamboesberg Member is capped by the more resistant Indwe Member, which is characterized by coarse-grained, pebbly sandstone (Christie, 1986; Turner, 1975), a characteristic weathering pattern, and by the glittery appearance imparted by the presence of secondary quartz overgrowths (Hancox, 1998; Turner, 1975). An erosive surface separates it from the underlying Bamboesberg Member (Turner, 1975). The Indwe Member attains a maximum thickness of 60 m near Indwe (Christie, 1981) and thins rapidly from south to north, following the regional pattern of the Molteno Formation, in that the major thinning (Hancox, 1998) occurs in the region of Maletswai (Aliwal North) where it crosses over onto the Kaapvaal Craton. The Indwe Member is succeeded by the dominantly argillaceous (informal) “Mayiputi” Member, followed by the arenaceous “Qiba”, predominantly argillaceous “Tsomo” and coarse, arenaceous “Loskop” members (Christie, 1981, 1986, but see Bordy et al., 2005; Turner, 1975). Although these achieve great thickness, especially further east (Christie, 1986), they outcrop at the higher altitudes of the Bamboesberg plateau and do not contribute sediment to the Great Fish River drainage, of which the Vlekpoort River forms a part. A coarse-grained sandstone bed defines the top of the “Loskop Member” and Molteno Formation and is overlain by reddish-colored mudstones of the Elliot Formation. The upper contact with the Elliot Formation is not exposed in the map area, with the closest approach located ~55 km to the northeast (Fig. 4.4). The Elliot Formation is a red-bed succession of mudstones, siltstones and subordinate fine- to medium-grained sandstones, lacking widespread marker beds, with strong red–purple-maroon diagenetic coloration, mainly of the argillaceous lithologies (Bordy & Eriksson, 2015; Bordy et al., 2020). The maximum preserved thickness of 460– 480 m is reported in the type area near Khowa (Elliot), with thicknesses of 260–300 m reported northeast of the town of Molteno (Bordy et al., 2004b; Botha, 1968). Bordy et al., (2004a, 2004b) proposed an informal lithostratigraphic subdivision of the Elliot Formation, with the lower Elliot Formation characterized by thick, multi-storied sandstones separated by thick (up to 30 m), laterally extensive mudstone intervals. In contrast, the upper Elliot Formation is characterized by fine to very fine sandstone sheets several tens of meters wide (average >100 m), and up to 6 m thick, separated by mudstone units 0.5–10 m thick. The Elliot Formation grades into the Clarens Formation, which consists almost exclusively of massive, well-sorted, fine-grained

4

Geological Setting of Hofmeyr Locality

sandstone of eolian genesis (Bordy & Head, 2018; Johnson et al., 2006). The latter formation forms prominent outcrops in the mountains northeast of Molteno.

Karoo Bedrock Minerology The soils of the landscape in the immediate vicinity of the fossil locality derive from the underlying Katberg, Burgersdorp, Molteno and, to the south, also the Balfour, formations (Fig. 4.4). Little is known about the mineralogy of the argillaceous rocks of these formations, with research restricted to a thin interval at the top of the Balfour Formation (Coney et al., 2007; Li et al., 2017). Li et al. (2017) report quartz, feldspars, chlorite and illite as the only minerals present at consistently detectable levels. More is known about the mineralogy of the sandstones, with the Adelaide Subgroup sandstones having been described as lithic (Johnson, 1991) and feldspathic to lithic wackes (Gastaldo et al., 2018) as defined by Pettijohn (1975), although Viglietti et al. (2018) reported lower matrix and higher quartz content for these sandstones. In general, the quartz content of the sandstone increases progressively up in the stratigraphy. The wackes of the Katberg Formation demonstrates an upwards increase in quartz at the expense of feldspar, although it still plots (Fig. 4.6) as lithic arenite and lithic wacke (Hancox, 1998; Johnson, 1991; Pace et al., 2009, but see Viglietti et al., 2018). The sandstones of the Burgersdorp Formation have also been described as lithic arenite to sublithic arenites and lithic greywacke (Fig. 4.6) and exhibit an upwards increase in

Fig. 4.6 Framework mineralogy plot for the sandstones of the Balfour, Katberg, Burgersdorp formations and the Bamboesberg and Indwe members of the Molteno Formation as reported by Johnson (1991), Hancox (1998) and Viglietti et al. (2018). (Q = quartz, F = feldspar, L = Lithics)

35

quartz content (Hancox, 1998). The sandstones of the lower Burgersdorp Formation show low modal quartz and high rock fragment values, similar to the sandstones of the upper Katberg Formation. Sandstones from the middle of the Formation show higher quartz values, but are consistently below 60%, whereas those from the upper Burgersdorp Formation are consistently over 60% modal quartz. Feldspar grains show various stages of textural and mineralogical maturity, with sodium plagioclase (albite) predominant (Hancox, 1998). In general, feldspars from the finer grained overbank siltstone and sandstone splays show less evidence of alteration. Heavy minerals are rare in the middle and upper Burgersdorp Formation, being more abundant in the lowermost reaches of the Formation in the southwest, where they are represented by oxides and zircon (Hancox, 1998). Rare rock fragments include polycrystalline quartz grains of sedimentary, plutonic and metamorphic origin, with rare granite fragments (Du Toit, 1954; Hancox, 1998). Johnson (1991) interpreted the continued presence of feldspar and volcanic fragments in significant quantities in the Katberg and Burgersdorp formations as evidence of a dissected arc source. Hancox (1998), based on the presence of granite, reworked sedimentary and metamorphic rock fragments, as well as the heavy mineral suite, proposed a mixed plutonic, low grade metamorphic and sedimentary provenance for the Burgersdorp Formation, with the latter probably including the rocks of the underlying Dwyka, Ecca and Lower Beaufort groups. Quartz forms the predominant component of the sandstones of the two lower units in the Molteno Formation, with quartz overgrowths common (Hancox, 1998). The Bamboesberg Member sandstones have been classified as sublithic arenites, lithic arenites, arkosic arenites, subarkoses and arkoses, while the Indwe Member also include quartz arenites (Fig. 4.6). Feldspar grains vary from fresh to totally weathered grains, with sodium rich plagioclase predominating in the basal part of the Bamboesberg Member, while potassium feldspars increase in abundance up-section into the Indwe Member. Rock fragments occur relatively commonly in the coarser grained sandstones of the Bamboesberg and Indwe members and consist predominantly of polycrystalline quartz of sedimentary, igneous and metamorphic origin. Granite fragments and quartz arenite pebbles have also been recorded. Preserved detrital mica is rare, but when present, muscovite is the dominant form. Heavy minerals usually occur as accessory minerals in concentrations less than 1% (Hancox, 1998). Christie (1981) lists the four non-opaque heavy minerals present as garnet, zircon, rutile and tourmaline. Opaque heavy minerals include ilmenite and magnetite, with rare leucoxene (Christie, 1981). Based on petrographic evidence the provenance area for the Molteno Formation is interpreted to have included a

36

number of different source rock types. Hancox (1998) proposed that the provenance area included a mixture of low to high grade metamorphic and acid plutonic rocks. The granite component is believed to have been supplied from the eastern margin of the Falkland Plateau (Turner, 1975; Veevers et al., 1994), and the metamorphic material from within the southern Cape Fold Belt, with some contributions from the underlying formations of the Karoo Supergroup (Hancox, 1998).

Gondwana Breakup The Clarens Formation is capped by the Drakensberg Group, which is a dominantly basaltic succession up to 1.5 km thick (Duncan & Marsh, 2006). It simultaneously terminated sedimentary deposition in the Karoo Basin and heralded the break-up of Gondwana (Watkeys, 2006). Regional variation is reported for the lower lava units of the Drakensberg Group (De Wit et al., 2020; Lock et al., 1974; Marsh & Eales, 1984; Marsh et al., 1997). The thick, upper units of the Drakensberg Group (assigned to the Lesotho Formation) show more uniform thicknesses and geochemistry and have been interpreted to represent stacked lava flows that erupted via a widespread network of fissures onto a planar surface. Based on 40Ar/39Ar dating, Duncan et al. (1997) determined that the entire lava pile was extruded and crystalized within a very short period, at 183 ± 1 Ma (i.e. during the Lower Jurassic). The eruption of the Drakensberg basalts was accompanied by the emplacement of dolerite dykes and sills in the underlying beds which are interpreted as the shallow feeder system for these flood basalts (Coetzee & Kisters, 2018). The extensive presence of this dyke and sill network, which share a close geochemical signature with the magma described from the lava pile, in the sedimentary rocks of the Karoo indicate that the Karoo lavas preserved today are just the erosional remnants of a once far more extensive carapace of lava that appears to have covered much of southern Africa (Duncan & Marsh, 2006; Hanson et al., 2009). The Karoo dolerite suite occurs as an interconnected network of dykes, sills and saucer-shaped sheets (Chevallier & Woodford, 1999). The thicknesses of the sills and sheets range from a few meters to 200 m and these often form resistant ledges capping many of the hills and mountains of the modern Karoo landscape (Fig. 4.5e). Dykes are generally 2–10 m wide and 5–30 km long (Duncan & Marsh, 2006). A general increase in tectonic activity during the Lower Jurassic, associated with Karoo volcanism, was followed by the linkage of fracture systems across Gondwana (Watkeys, 2006). The supercontinent split into West and East Gondwana (the former consisting of Africa and South America)

J. Neveling

between 155 and 135 Ma (Fig. 4.7). The splitting of West Gondwana into two plates (South America and Africa) commenced in ~132 Ma, as marked by the onset of extraction of the Falkland Plateau from the southern Natal Valley (Goodlad et al., 1982). Break-up was complete (Watkeys, 2006) when the Falkland Plateau cleared the continental margin during the Upper Cretaceous (Martin & Hartnady, 1986; Partridge & Maud, 1987; Roberts et al., 2006). By about 90 Ma the present-day plate configuration of the Southern Hemisphere had emerged, and the continental margin and present coastal configuration of southern Africa has been relatively stable since.

Post-breakup Southern Africa At the time of continental separation, Africa stood high, with mean surface elevations of 2400 m predicted for Lesotho, decreasing to 1500 m for the western interior (Partridge, 1998; Partridge & Maud, 1987; Rust & Summerfield, 1990). This raises questions about the nature of the coastal topography at the time. There has been considerable debate on the position and rate of retreat of the Great Escarpment (Brown et al., 2002; Doucoré & De Wit, 2003; Partridge & Maud, 1987), which today is a seaward-facing, horseshoe-shaped scarp that trends shore-parallel to, and between 50 and 200 km inland of, the coast of Namibia, South Africa, and Lesotho (Fig. 4.1). Traditionally the escarpment has been considered to have been a shoreline feature following the breakup of Gondwana, which has since retreated episodically to its current position (Burke, 1996; King, 1953; Partridge & Maud, 1987). These models propose that southern Africa was subjected to episodes of rapid uplift (e.g., Partridge & Maud, 1987), with the modern landscape a relatively recent feature. This school of thought is best typified by the models of Partridge and Maud (1987, 2000; Partridge et al., 2006) who, based on the assumptions that surfaces originally graded to sea level and that the Great Escarpment was originally at the coast, developed a chronology for the geomorphic evolution of southern Africa that included three large-scale erosional phases, punctuated by asymmetrical uplift of the sub-continent during the Miocene and Pliocene respectively. More recently, a more static model has emerged that proposes the formation of the Great Escarpment near its present location, through differential erosion, during the Cretaceous, with slow to no uplift (Brown et al., 2002; Cockburn et al., 2000; Doucoré & De Wit, 2003; Fleming et al., 1999; Gurnis et al., 2000). In this model, the escarpment formed just east of its present position as a result of rapid river-incision after Gondwana breakup and largely

4

Geological Setting of Hofmeyr Locality

37

Fig. 4.7 Southern Africa at ~135 Ma. Western Gondwana is starting to split into South America and Africa, heralding a predominantly erosional environment for the southern African subcontinent. Arrows demonstrate predominant sediment transport patterns

maintained its position due to flexural isostatic rebound, with rapid fluvial erosion and dissection of the landscape during the Cretaceous. The high elevation of the southern African sub-continent (and possible high rates of uplift), combined with the humid tropical climate during the Cretaceous, initiated a major erosive phase in the evolution of southern African (Brown et al., 2000; Cockburn et al., 1999; Partridge, 1998; Partridge & Maud, 1987; Tinker et al., 2008). Much of the sediment resulting from this denudation, was deposited in a series of onshore and offshore graben and half-graben rift basins that formed as a result of extension during break-up, in the Cape Fold Belt and along the western and southern coast of South Africa (Bate & Malan, 1992; McMillan et al., 1997; Thomson, 1999). The resulting marine sequences of Cretaceous age are of considerable thickness (in excess of 8 km in places) and support the findings of fission track analyses onshore (Brown et al., 1994), which indicate that by the mid-Cretaceous, between 1 and 3 km of bedrock had been removed from the post-rifting surface of the subcontinent (Hanson et al., 2009; Partridge & Maud, 2000), including from the Karoo Supergroup.

Tinker et al. (2008), assessing offshore sediment accumulation, identified two pulses of denudation (140–120 Ma and 100–80 Ma) during the Cretaceous, with the earlier period having a greater impact on landscape degradation of the hinterland. In contrast, denudation rates decreased significantly during the Cenozoic (Tinker et al., 2008). Doucouré and De Wit (2003) reconstructed the paleotopography of the African continent using seismic tomography and geological data. Their reconstruction showed that the topography was already high during the Cretaceous, and that there is no need to invoke high rates of Cenozoic uplift to explain the modern topography. It implies that the southern African landscape to a large degree resembled the presentday topography by the end of the Mesozoic. Another outcome of these cycles of sediment removal was the emergence of a vast planation surface, developed at different elevations above and below the Great Escarpment, first referred to as the African Surface by King (1944, 1947) and Fair and King (1954). Today, a number of mountain massifs endure above this surface, including the Cape Fold Mountains and the Drakensberg range, which extends through the Eastern Cape Province, Lesotho and KwaZulu-Natal.

38

Cenozoic Geology An integrated drainage network had been established on the African Surface by the Late Cretaceous, with remnants found today in several high-lying fluvial terraces (Partridge et al., 2006). The rifting of West Gondwana into South America and Africa during the Cretaceous initiated the development of two river networks (the Karoo River and Kalahari River), which drained the larger area of South Africa (De Wit, 1993). The Kalahari River subsequently captured the middle reaches of the Karoo River to form the Orange River (De Wit et al., 2000), which still drains more than half of South Africa’s surface area, as well as large parts of Namibia. Most of the eastern and southern escarpment is drained by short east- and south-flowing rivers (De Wit et al., 2000). High relief along the escarpment produced relatively small drainage basins, with rivers that are predominantly degradational at present. Given the long term degradational setting of South Africa’s interior, older Cenozoic (i.e., Paleocene to Pliocene) deposits are rare (Partridge et al., 2006) and either physically constrained (e.g., crater-fills) or represented by high-lying erosional remnants of larger features (e.g., high-lying fluvial terraces). Hillslope deposits are widespread in the northern part of the Eastern Cape Province, as in other areas of topographic relief, usually occurring as thin veneers of talus, colluvial gravel or pedisediment (Partridge et al., 2006). Talus deposits, consisting of coarse angular rock fragments, drape the slopes of the mountains and hills of the study area, grading downhill into colluvium (Cole et al., 2004). Van Niekerk (1977) proposed that older talus deposits along the steep hill and mountain slopes, which are locally well cemented by carbonates, probably date from the Miocene. Thick (up to 25 m), unconsolidated colluvial deposits often drape bedrock on lower hillslopes in KwaZulu-Natal, the Eastern Cape and Mpumalanga provinces. These deposits accumulated during the Late Pleistocene and Holocene (Botha & Partridge, 2000; Botha et al., 1994; Clarke et al., 2003) and have been assigned to the Masotcheni Formation (Partridge et al., 2006). Colluvial sediments predominantly accumulated through sheetwash transport of eroded soils and talus (Botha & Partridge, 2000) and commonly show the textural and structural influence of post-depositional pedogenic weathering (Botha & Federof, 1995) so that the sequences exhibit stacked, buried paleosols. These paleosol units can be correlated between dongas over a wide area and reflects the impact of changing environments. The deposits are susceptible to donga erosion (Partridge et al., 2006) and record cutand-fill episodes that erode pre-weathered colluvium, soil and larger clasts (Botha, 1996). Working primarily in KwaZulu-Natal, Botha et al. (1994) recognized at least four

J. Neveling

geomorphic cycles, comprising gully cut-and-fill and paleosol formation during the past 120,000 years. Alluvium occurs widespread across the study area (Cole et al., 2004) and achieves thicknesses in excess of six meters (Holmes & Marker, 1995). It consists primarily of coarse- to fine-grained sand, with rare coarser pebble stringers, grading upwards into clay. Clays in the alluvium of the high-lying valleys close to the escarpment (which Cole et al., 2004 placed further north than Partridge & Maud, 2000) are often rich in organic material (Cole et al., 2004). Calcrete commonly occurs on the pediplains and is most commonly associated with alluvium in areas where the thickness of medium- to coarse-grained sand exceeds one meter (Cole et al., 2004). The formation of calcrete on the pediplains is mainly associated with a fluctuating water table under semiarid conditions. Partridge (1993) and Bousman et al. (1988) proposed that deep, in situ, bedrock weathering during the Holocene contributed significantly to the alluvial deposits of the valleys and plains of the interior. A more common view is that the alluvium resulted from cut-and-fill sedimentation by steep, swiftly flowing mountain streams or sheetwash deposition (Cole et al., 2004). Holmes and Marker (1995) reported Holocene ages for paleosols contained in the alluvium at three sites between Cradock and Middelburg and Cole et al. (2004) suggested that only Holocene sediments remain in the plain alluvium. They did agree, however, with De Wit (1993) who proposed an older age (e.g., Early Pleistocene or even Pliocene) for the sandy alluvial deposits draping the pediment bordering the proximal river valleys (such as the Sundays and Fish rivers), which is supported by the Late Pleistocene age reported for the Hofmeyr cranium (Grine et al., 2007). The Hofmeyr cranium was collected from the banks of Vlekpoort River, which forms part of the upper reaches of the Great Fish River drainage system (Fig. 4.8). Until recently, research on Cenozoic fluvial systems of South Africa tended to focus on the older, larger drainage systems of central and northern South Africa (De Wit et al., 2000). In the Eastern Cape Province Hattingh (1996) described flights of fluvial terraces associated with the Sundays River drainage south of Kirkwood. He assigned a Late Miocene age to the higher-lying terraces, whose abundant coarse sediments record deposition under warm, humid conditions. The lowerlying terraces have been interpreted to span much of the intervening period up to Holocene times and record increasing aridity. Hill (1993), working further north, reported the presence of calcretized terraces along the upper reaches of the Sundays River drainage system between Graaff-Reinet and Pearston, as well as the Great Fish River and some of its tributaries. To date no detailed investigation has been undertaken on the fluvial deposits associated with the upper reaches of either of these river systems.

4

Geological Setting of Hofmeyr Locality

39

Fig. 4.8 The Hofmeyr cranium was collected (yellow star) from the banks of the Vlekpoort River. The latter forms part of the Fish River drainage, which borders the Sundays River drainage to the west. The position of the Erfkroon locality (red pentagon) is shown on the inset map

The paleoenvironmental setting governing southern Africa’s fluvial systems during the Quaternary is still not well understood (Tooth, 2015). Most research has focused on drainage systems in Namibia (Eitel et al., 2006; Srivastava et al., 2006), the larger Vaal and Orange systems (Keen-Zebert et al., 2013; Lyons et al., 2014; Tooth & McCarthy, 2004) and rivers draining the eastern escarpment (Keen-Zebert et al., 2013; Tooth et al., 2007, 2013). Of these, the middle Modder River provides perhaps the closest analogue to the Vlekpoort River in terms of the prevailing climate, source and bedrock geology (although the Vlekpoort is much smaller and situated closer to the drainage headwaters). A multi-disciplinary investigation was undertaken at Erfkroon, a Florisian fossil locality situated along the middle Modder River (which forms part of the Orange-Vaal River catchment) in the Free State Province of South Africa (Churchill et al., 2000; Tooth et al., 2013). Claassen (2018) assigned these deposits to the Riverton Formation, which refers to a series of stepped alluvial terraces that occurs intermittently along the banks of the lower Harts, Vaal, Riet and middle Orange rivers. The Riverton Formation consists of unconsolidated to semi-consolidated alluvial and colluvial sands, silts and clays with subordinate gravel beds and lenses that erosionally overly the older Rietputs Formation strata or Karoo bedrock (De Wit et al., 2000). Younger deposits are stacked upon, cross-cut, or

onlap older deposits, with the youngest deposits typically located closest to the main channel. Nine depositional stages, spanning the past 163,000 years were recognized at Erfkroon (Churchill et al., 2000; Tooth et al., 2013) of which one overlaps the age of the Hofmeyr cranium. Lyons et al. (2014) proposed that a dry climate (e.g., rainfall of 200–400 mm/year) predominated in the western Free State between 46 and 32 ka, which is comparable or slightly lower than the modern annual rainfall (Schultze, 1997; Shultze & Kunz, 2011). Chase and Meadows (2007) postulated slightly drier conditions (compared to modern environments) for the Eastern Cape hinterland over the last glacial maximum (i.e., 26.5–19 ka; Clark et al., 2009), but little is known about the prevailing climate during the era that immediately preceded it, when the Hofmeyr individual lived. Since the same modern summer rainfall region (Chase & Meadows, 2007) encompass the Erfkroon and Hofmeyr localities, which also share similar modern annual rainfall figures (Shultze & Kunz, 2011), it is postulated that the Hofmeyr region exhibited a climate similar to, or slightly drier than, modern climate during the life of the Hofmeyr individual. At Erfkroon, the Modder River cut a narrow (  500 m wide) valley into bedrock that has been infilled by alluvium. Present-day deep channel incision has led to floodplain abandonment and the formation of a paired alluvial terrace. Similarly, the Vlekpoort River is bounded and constrained

40

J. Neveling

Fig. 4.9 The Vlekpoort River Valley. A Airphoto image of the Hofmeyr cranium locality. The yellow star represents the position of sampling locality H2 (Copeland et al., 2022). The eroded edge of the river valley forms a prominent ridge to the north of the Vlekpoort River. Low hills form another topographic high a few kilometers to the north. B Field photograph of the Vlekpoort River Valley (from the southern bank of the Vlekpoort River, facing to the north). The field party stands at the position of sampling locality H2 (Copeland et al., 2022). The river valley edge is visible *600 m behind them and the low hills bounding the valley in the distance

by resistant Karoo bedrock. At the Hofmeyr fossil locality, the river valley is up to a kilometer wide and bounded by a prominent bedrock ridge to the north (Fig. 4.9). Approximately 3 km downstream the river valley is constrained by low hills and narrows to less than 300 m (Fig. 4.10). In narrow, bedrock-confined river valleys, such as the Vlekpoort River northeast of Hofmeyr, lateral sediment distribution through meandering and avulsion is restricted and vertical aggradation dominates (Tooth et al., 2013). The lowest depth of channel incision (or lower profile buffer; Holbrook et al., 2006), is determined by the stability of any downstream barrier (Fig. 4.11). In large parts of South Africa (Tooth & McCarthy, 2004, 2013) this is often set by resistant dolerite sills and dikes that crop out in riverbeds, providing an anchor for upstream river longitudinal profiles.

The highest surface of aggradation, or upper buffer profile, is determined by channel–floodplain processes that occur within the constraints imposed by valley width and climatically driven aggradation phases (Holbrook et al., 2006; Tooth et al., 2013). Sediment deposition is controlled by climatically driven changes in flow regime, sediment supply and fluvial style. Thus, deposition can become decoupled from bedrock and thick alluvial packages are commonly characterized by cut-and-fill processes. Barriers (also referred to as buttresses) may be buried by alluvium during aggradational phases (Holbrook et al., 2006). When downstream barriers are breached upstream channel incision can take place over a wide vertical range that may exceed the thickness of accumulated alluvial sediments and continue into bedrock, as is currently the case at Erfkroon (Tooth et al., 2013).

4

Geological Setting of Hofmeyr Locality

41

Fig. 4.10 Google Earth image of the Vlekpoort River Valley, showing topographic control on valley width downstream of the Hofmeyr skull locality (yellow star). A prominent dolerite dyke (red line) forms a natural fluvial buttress approximately 5.5 km downstream. Artificial buttresses were introduced by the construction of weirs and dam wall in the Vlekpoort channel (yellow lines)

Resistant dolerite dykes crisscross the Springbok Flats, with a prominent dyke approximately 4 km to the south of the Hofmeyr cranium locality a good candidate for a natural downstream buttress. However, the natural evolution of the Vlekpoort River has been interrupted in the past century by the construction of numerous dams and weirs in the river channel. When the Hofmeyr skull was discovered in the early 1950s, waterflow was unconfined, but within ten years of the discovery a small anti-erosion weir had been constructed approximately 200 m downstream from the location at which the skull had been found. Today there are at least three weirs present in the Vlekpoort channel in the 5 km downstream of the Hofmeyr cranium locality, with another two above (Fig. 4.10). These man-made structures anchor, or in some cases raise, the local base level, resulting in increased deposition of recent sediments over older deposits, upstream.

Conclusions The high-lying Springbok Flats and Bamboesberg mountains which dominate the geography north of Hofmeyr, were formed by a protracted erosional-dominant phase which commenced with the break-up of Gondwana in the Jurassic. Many of the regional landscape features were most likely established during the Cretaceous when between 1 and 3 km of bedrock was removed from the South African interior. A key erosional landscape feature is the Great Escarpment, a shore-parallel scarp that separates a high interior (where Hofmeyr is located) from a lower coastal region 50–200 km wide. Locally, high topographical features, such as the isolated hills and mountains of the Karoo landscape north of Hofmeyr, and the Bamboesberg mountains to the east of the fossil locality, are erosional remnants protected by harder

42

J. Neveling

Fig. 4.11 Impact of buttress on fluvial profiles. With a stable buttress aggradation and down-cutting is restricted within a buffer zone. When buttress fall (down-cutting) the buffer zone also drops, resulting in upslope erosion. When buttress rise (such as the introduction of man-made structures) this creates additional preservation space above the buttress that is filled via depositional processes

and more weathering-resistant cap-rocks of dolerite and sandstone. Landscape denudation (and local sedimentation) continues to this day, albeit at much reduced rates, but it can be safely assumed that the Hofmeyr landscape at the end of the Pleistocene looked much as it does today. Short and steep south-flowing river systems, such as the Great Fish River, drain the landscape seaward across the Great Escarpment, with the Vlekpoort River as a small tributary in the uppermost reaches of this drainage. The sediments transported by, and deposited on the banks of, the Vlekpoort River, which also contribute to the alluvium and soils of the Springbok Flats, derive from the siltstones and sandstones of the underlying Burgersdorp Formation and the Molteno Formation, which forms the Bamboesberg mountains to the east. These local source rocks accumulated as fluvial deposits in the ancient Karoo Basin during the Triassic Period and, as of yet, appear to be geochemically indistinct. This suggests that regional differences in vegetation are the result of variation in local geomorphology, altitude and microclimate. The Vlekpoort River is hosted by a shallow, bedrock-confined valley that is up to a kilometer wide at

the fossil locality but narrows to ~300 m approximately 3 km downstream. Lateral sediment distribution through meandering and avulsion is generally limited in bedrock-confined valley settings, with vertical aggradation (through repeated cut-and-fill episodes) predominant within the constraints imposed by valley width. Sediment deposition in such settings is normally controlled by climatically driven changes in flow regime, sediment supply and fluvial style. Such a mechanism is broadly supported by the overlap in ages of the Hofmeyr cranium and depositional stage 8 reported from the Erfkroon locality in the western Free State. Today Hofmeyr and Erfkroon share a broadly similar, low summer rainfall pattern and it is therefore assumed that during the previous interglacial the Eastern Cape hinterland, similar to what has been reported for the western Free State, experienced a climate comparable to, or slightly drier than, modern times. Hence the Hofmeyr individual had to face a natural world not too dissimilar from that facing local inhabitants of this area today. Acknowledgements John Hancox is thanked for his cogent comments and suggestions which improved this manuscript.

4

Geological Setting of Hofmeyr Locality

References Baiyegunhi, C., & Gwavava, O. (2016). Variations in isochore thickness of the Ecca sediments in the Eastern Cape Province of South Africa, as deduced from gravity models. Acta Geologica Sinica, 90, 1699–1712. Bate, K. J., & Malan, J. A. (1992). Tectonostratigraphic evolution of the Algoa, Gamtoos and Pletmos Basins, offshore South Africa. In M. J. de Wit & I. G. D. Ransome (Eds.), Inversion tectonics of the Cape Fold Belt, Karoo and Cretaceous Basins of southern Africa (pp. 61–73). A.A. Balkema. Bordy, E., Abrahams, M., Sharman, G. R., Viglietti, P. A., Benson, R. B. J., McPhee, B. W., et al. (2020). A chronostratigraphic framework for the upper Stormberg Group: Implications for the Triassic-Jurassic boundary in southern Africa. Earth-Science Reviews, 203, 103120. Bordy, E. M., & Eriksson, P. (2015). Lithostratigraphy of the Elliot Formation (Karoo Supergroup), South Africa. South African Journal of Geology, 118, 311–316. Bordy, E. M., Hancox, P. J., & Rubidge, B. S. (2004a). A description of the sedimentology and palaeontology of the Late Triassic – Early Jurassic Elliot Formation in Lesotho. Palaeontologia Africana, 40, 43–58. Bordy, E. M., Hancox, P. J., & Rubidge, B. S. (2004b). Basin development during the deposition of the Elliot Formation (Late Triassic – Early Jurassic), Karoo Supergroup, South Africa. South African Journal of Geology, 107, 397–412. Bordy, E. M., Hancox, P. J., & Rubidge, B. S. (2005). The contact of the Molteno and Elliot formations throughout the main Karoo Basin, South Africa: A second-order sequence boundary. South African Journal of Geology, 108, 351–364. Bordy, E. M., & Head, H. V. (2018). Lithostratigraphy of the Clarens Formation (Stormberg Group, Karoo Supergroup), South Africa. South African Journal of Geology, 121, 119–130. Botha, B. J. V. (1968). The stratigraphy of the Red Beds Stage, Karoo System, at Elliot, C.P. Transactions of the Geological Society of South Africa, 71, 101–113. Botha, G. A. (1996). Cyclical colluvial accretion on bedrock pediments during the late Quaternary in northern Kwazulu-Natal, South Africa. Zeitschrift Für Geomorphologie, 103, 85–102. Botha, G. A., & Federof, N. (1995). Palaeosols in late Quaternary colluvium, northern KwaZulu–Natal, South Africa. Journal of African Earth Sciences, 21, 291–311. Botha, G. A., & Partridge, T. C. (2000). Colluvial deposits. In T. C. Partridge & R. R. Maud (Eds.), The Cenozoic of southern Africa (pp. 88–99). Oxford University Press. Botha, G. A., Wintle, A. G., & Vogel, J. C. (1994). Episodic late Quaternary palaeogully erosion in northern Kwazulu-Natal, South Africa. CATENA, 23, 327–340. Bousman, C. B., Partridge, T. C., Scott, L., Metcalfe, S. E., Vogel, J. C., Seaman, M., et al. (1988). Palaeoenvironmental implications of Late Pleistocene and Holocene valley fills in Blydefontein Basin, Noupoort, C.P. South Africa. Palaeoecology of Africa, 19, 43–67. Brown, R. W., Gallagher, K., Gleadow, A. J. W., & Summerfield, M. A. (2000). Morphotectonic evolution of the South Atlantic margins of Africa and South America. In M. A. Summerfield (Ed.), Geomorphology and global tectonics (pp. 255–280). Wiley. Brown, R. W., Summerfield, M. A., & Gleadow, A. J. W. (1994). Apatite fission track analysis: Its potential for estimation of denudation rates and implications for models of long-term landscape development. In M. J. Kirkby (Ed.), Process models and theoretical geomorphology (pp. 23–53). Wiley. Brown, R. W., Summerfield, M. A., & Gleadow, A. J. W. (2002). Denudational history along a transect across the Drakensberg

43

Escarpment of Southern Africa derived from apatite fission track thermochronology. Journal of Geophysical Research, 107, 2350. https://doi.org/10.1029/2001JB000745 Burke, K. (1996). The African plate. South African Journal of Geology, 99, 341–409. Catuneanu, O., & Elango, H. N. (2001). Tectonic control on fluvial styles: The Balfour Formation of the Karoo Basin, South Africa. Sedimentary Geology, 140, 291–313. Catuneanu, O., Hancox, P. J., Cairncross, B., & Rubidge, B. S. (2002). Foredeep submarine fans and forebulge deltas: Orogenic off-loading in the underfilled Karoo Basin. Journal of African Earth Sciences, 35, 489–502. Catuneanu, O., Hancox, P. J., & Rubidge, B. S. (1998). Reciprocal flexural behaviour and contrasting stratigraphies: A new basin development model for the Karoo retroarc foreland system, South Africa. Basin Research, 10, 417–439. Catuneanu, O., Wopfner, H., Eriksson, P. G., Cairncross, B., Rubidge, B. S., Smith, R. M. H., et al. (2005). The Karoo basins of southcentral Africa. Journal of African Earth Sciences, 43, 211–253. Chase, B. M., & Meadows, M. E. (2007). Quaternary dynamics of southern Africa’s winter rainfall zone. Earth-Science Reviews, 84, 103–138. Chevallier, L., & Woodford, A. C. (1999). Morpho-tectonics and mechanism of emplacement of the dolerite rings and sills of the western Karoo, South Africa. South African Journal of Geology, 102, 43–54. Christie, A. D. M. (1981). Stratigraphy and sedimentology of the Molteno Formation in the Elliot and Indwe area, Cape Province. M.Sc. thesis, University of Natal, Durban. Christie, A. D. M. (1986). Molteno Coalfield. In C. R. Anhaeusser & S. Maske (Eds.), Mineral deposits of southern Africa (Vol. II, pp. 2063–2070). Geological Society of South Africa. Churchill, S. E., Brink, J. S., Berger, L. R., Hutchinson, R. A., Rossouw, L., Stynder, D., et al. (2000). Erfkroon: A new Florisian fossil locality from fluvial contexts in the western Free State, South Africa. South African Journal of Science, 96, 161–163. Claassen, D. (2018). The lithostratigraphy of the Riverton Formation, South Africa. South African Journal of Geology, 121, 327–335. Clark, P. U., Dyke, A. S., Shakum, J. D., Carlson, A. E., Clark, J., Wohlfarth, B., et al. (2009). The Last Glacial Maximum. Science, 325, 710–714. Clarke, M. L., Vogel, J. C., Botha, G. A., & Wintle, A. G. (2003). Late Quaternary hillslope evolution recorded in eastern South African colluvial Badlands. Palaeogeography, Palaeoclimatology, Palaeoecology, 197, 199–212. Cockburn, H. A. P., Brown, R. W., Summerfield, M. A., & Seidl, M. A. (2000). Quantifying passive margin denudation and landscape development using a combined fission-track thermochronology and cosmogenic isotope analysis approach. Earth and Planetary Science Letters, 179, 429–435. Cockburn, H. A. P., Seidl, M. A., & Summerfield, M. A. (1999). Quantifying denudation rates on inselbergs in the central Namib Desert using in situ–produced cosmogenic 10Be and 26Al. Geology Boulder, 27, 399–402. Coetzee, A., & Kisters, A. F. M. (2018). The elusive feeders of the Karoo Large Igneous Province and their structural controls. Tectonophysics, 747–748, 146–162. Cole, D. I. (1992). Evolution and development of the Karoo Basin. In M. J. de Wit & I. G. D. Ransome (Eds.), Inversion tectonics of the Cape Fold Belt, Karoo and Cretaceous Basins of Southern Africa (pp. 87–99). Balkema. Cole, D. I., Johnson, M. J., & Day, M. O. (2016). Lithostratigraphy of the Abrahamskraal Formation (Karoo Supergroup), South Africa. South African Journal of Geology, 119, 415–424.

44 Cole, D. I., Neveling, J., Hattingh, J., Reddering, J. S. V., Chevallier, L. P., & Bender, P. A. (2004). The geology of the Middelburg area. Explanation of Sheet 3124 (scale 1:250,000) (43 pp., 39 figs, 2 tables). Council for Geoscience. Coney, L., Reimold, W. U., Hancox, J., Mader, D., Koeberl, C., McDonald, I., et al. (2007). Geochemical and mineralogical investigation of the Permian-Triassic boundary in the continental realm of the southern Karoo Basin, South Africa. Paleoworld, 16, 67–104. Copeland, S. R., Grimes, V., Neveling, J., Lee-Thorp, J. A., Grine, F. E., Yang, Z., et al. (2022). Isotopic evidence for the geographic origin, movement and diet of the Hofmeyr individual. In F. E. Grine (Ed.), Hofmeyr – A Late Pleistocene human skull from South Africa (pp. 47–68). Springer. de Wit, M. C. J. (1993). Cainozoic evolution of drainage systems in the North-western Cape. Ph.D. thesis, University of Cape Town. de Wit, M. C. J., Linol, B., Furnes, H., Muedi, T., & Valashiya, K. (2020). Pillow talk: Volcanic rocks of the Karoo that formed many leagues under the Gondwana Sea. South African Journal of Geology, 123, 297–330. de Wit, M. C. J., Marshall, T. R., & Partridge, T. C. (2000). Fluvial deposits and drainage evolution. In T. C. Partridge & R. R. Maud (Eds.), The Cenozoic of southern Africa (pp. 55–72). Oxford University Press. Doucouré, C. M., & de Wit, M. J. (2003). Old inherited origin for the present near bimodal topography of Africa. Journal of African Earth Sciences, 36, 371–388. Duncan, R. A., Hooper, P. R., Rehacek, J., Marsh, J. S., & Duncan, A. R. (1997). The timing and duration of the Karoo igneous event, southern Gondwana. Journal of Geophysical Research, 102, 18127–18138. Duncan, A. R., & Marsh, J. S. (2006). The Karoo Igneous Province. In M. R. Johnson, C. R. Anhaeusser, & R. J. Thomas (Eds.), The Geology of South Africa (pp. 501–520). Geological Society of South Africa. Du Toit, A. L. (1954). Geology of South Africa (3rd ed.). Oliver and Boyd. Eitel, B., Kadereit, A., Blümel, W. D., Hüser, K., Lomax, J., & Hilgers, A. (2006). Environmental changes at the eastern Namib Desert margin before and after the Last Glacial Maximum: New evidence from fluvial deposits in the upper Hoanib River catchment, northwestern Namibia. Palaeogeography, Palaeoclimatology, Palaeoecology, 234, 201–222. Fair, T. J. D., & King, L. C. (1954). Erosional land-surfaces in the eastern marginal areas of South Africa. Transactions of the Geological Society of Southern Africa, 62, 19–26. Fleming, A., Summerfield, M. A., Stone, J. O., Fifield, L. K., & Cresswell, R. G. (1999). Denudation rates for the southern Drakensberg Escarpment, SE Africa, derived from in-situ–produced cosmogenic (super 36) Cl: Initial results. Journal of the Geological Society of London, 156, 209–212. Gastaldo, R. A., Neveling, J., Geissman, J. W., & Kamo, S. (2018). A lithostratigraphic and magnetostratigraphic framework in a geochronologic context for a purported Permian-Triassic boundary section at Old (West) Lootsberg Pass, Karoo Basin, South Africa. Geological Society of America Bulletin, 130, 1411–1439. Goodlad, S., Martin, A. K., & Hartnady, C. J. H. (1982). Mesozoic magnetic anomalies in the southern Natal Valley. Nature, 295, 686– 688. Grine, F. E., Bailey, R. M., Harvati, K., Nathan, R. P., Morris, A. G., Henderson, G. M., et al. (2007). Late Pleistocene human skull from Hofmeyr, South Africa, and modern human origins. Science, 315, 226–229. Groenewald, G. H. (1996). Stratigraphy of the Tarkastad Subgroup, Karoo Supergroup, South Africa. Ph.D. thesis, University of Port Elizabeth, South Africa.

J. Neveling Gurnis, M., Mitrovica, J. X., Ritsema, J., & van Heijst, H. J. (2000). Constraining mantle density structure using geological evidence of surface uplift rates: The case of the African superplume. Geochemistry, Geophysics, Geosystems, 1, Paper No. 1999GC000035. Hancox, P. J. (1998). A stratigraphic, sedimentological and palaeoenvironmental synthesis of the Beaufort–Molteno contact in the Karoo Basin. Ph.D. thesis, University of the Witwatersrand, Johannesburg. Hanson, E. K., Moore, J. M., Bordy, E. M., Marsh, J. S., Howarth, G., & Robey, J. V. A. (2009). Cretaceous erosion in central South Africa: Evidence for upper-crustal xenolits in kimberlite diatremes. South African Journal of Geology, 112, 125–140. Hattingh, J. (1996). Late Cenozoic drainage evolution in the Algoa Basin with special reference to the Sundays River Valley. Council for Geoscience Bulletin, 128, 1–103. Hill, R. S. (1993). The geology of the Graaff-Reinet area. Explanation of Sheet 3224 (scale 1:250,000) (31 pp.). Geological Survey of South Africa, Government Printer. Holbrook, J., Scott, R. W., & Oboh-Ikuenobe, F. E. (2006). Base-level buffers and buttresses: A model for upstream versus downstream control on fluvial geometry and architecture within sequences. Journal of Sedimentary Research, 76, 162–174. Holmes, P. J., & Marker, M. E. (1995). Evidence for environmental change from Holocene valley fills from three central Great Karoo upland sites. South African Journal of Science, 91, 617–620. Johnson, M. R. (1976). Stratigraphy and Sedimentology of the Cape and Karoo Sequences in the Eastern Cape Province. Ph.D. thesis, Rhodes University, Grahamstown. Johnson, M. R. (1991). Sandstone petrography, provenance and plate tectonic setting in Gondwana context of the southeastern CapeKaroo Basin. South African Journal of Geology, 94, 137–154. Johnson, M. R., & Hiller, N. (1990). Burgersdorp Formation. In M. R. Johnson (Ed.), Catalogue of South African lithostratigraphical units (Vol. 2, pp. 9–10). South African Committee for Stratigraphy. Johnson, M. R., Van Vuuren, C. J., Visser, J. N. J., Cole, D. I., Wickens, H., Christie, A. D. M., et al. (2006). Sedimentary rocks of the Karoo Supergroup. In M. R. Johnson, C. R. Anhaeusser, & R. J. Thomas (Eds.), The geology of South Africa (pp. 461–499). Geological Society of South Africa. Keen-Zebert, A., Tooth, S., Rodnight, H., Duller, G. A. T., Roberts, H. M., & Grenfell, M. (2013). Late Quaternary floodplain reworking and the preservation of alluvial sedimentary archives in unconfined and confined river valleys in the eastern interior of South Africa. Geomorphology, 185, 54–66. King, L. C. (1944). Geomorphology of the Natal Drakensberg. Transactions of the Geological Society of South Africa, 47, 255. King, L. C. (1947). Landscape study in southern Africa. Proceedings of the Geological Society of South Africa, 50, 23. King, L. C. (1953). Canons of landscape evolution. Geological Society of America Bulletin, 64, 721–752. Kitching, J. W. (1995). Biostratigraphy of the Cynognathus Assemblage Zone. In B. S. Rubidge (Ed.), Reptilian biostratigraphy of the Permian-Triassic Beaufort Group (Karoo Supergroup) (pp. 40–45). SACS Biostratigraphic Series. Li, J., Gastaldo, R. A., Neveling, J., & Geissman, J. W. (2017). Siltstones across the Daptocephalus (Dicynodon) and Lystrosaurus Assemblage zones, Karoo Basin, South Africa, show no evidence for aridification. Journal of Sedimentary Research, 87, 653–671. Lindeque, A., de Wit, M. J., Ryberg, T., Weber, M., & Chevallier, I. (2011). Deep crustal profile across the southern Karoo Basin and Beattie magnetic anomaly, South Africa: An integrated interpretation with tectonic implications. South African Journal of Geology, 114, 265–292. Lock, B. E., Paverd, A. L., & Broderick, T. J. (1974). Stratigraphy of the Karroo volcanic rocks of the Barkley East District. Transactions of the Geological Society of South Africa, 77, 117–129.

4

Geological Setting of Hofmeyr Locality

Lyons, R., Tooth, S., & Duller, G. A. T. (2014). Late Quaternary climatic changes revealed by luminescence dating, mineral magnetism and diffuse reflectance spectroscopy of river terrace palaeosols: A new form of geoproxy data for the southern African interior. Quaternary Science Reviews, 95, 43–59. MacDonald, A. J. (1993). A reassessment of coal resources in the western part of the Molteno coal province. Geological Survey of South Africa Bulletin, 116, 1–54. Marsh, J. S., & Eales, H. V. (1984). The chemistry and petrogenesis of igneous rocks of the Karoo Central Area, southem Africa. In A. J. Erlank (Ed.), Petrogenesis of the volcanic rocks of the Karoo Province (Vol. 13, pp. 27–67). Geological Society of South Africa, Special Publication. Marsh, J. S., Hooper, P. R., Rehacek, J., Duncan, R. A., & Duncan, A. R. (1997). Stratigraphy and age of Karoo basalts of Lesotho and implications for correlations within the Karoo Igneous Province. In J. J. Mahoney & M. F. Coffin (Eds.), Large igneous provinces (Vol. 100, pp. 247–272). Geophysical Monographs, American Geophysical Union. Martin, A. K., & Hartnady, C. J. H. (1986). Plate tectonic development of the southwest Indian Ocean: A revised reconstruction of East Antarctica and Africa. Journal of Geophysical Research, 91, 4767– 4786. McMillan, I. K., Brink, G. J., Broad, D. S., & Maier, J. J. (1997). Late Mesozoic sedimentary basins off the south coast of South Africa. In R. C. Selley (Ed.), Sedimentary basins of the world, 3: African basins (pp. 319–376). Elsevier. Mucina, L., & Rutherford, M. C. (Eds.). (2011). The vegetation of South Africa, Lesotho and Swaziland. Strelitzia, 19. South African Biodiversity Institute. ISBN: 978-1919976-21-1. Neveling, J. (2004). Stratigraphic and sedimentological investigation of the contact between the Lystrosaurus and the Cynognathus Assemblage Zones (Beaufort Group: Karoo Supergroup). Council for Geoscience Bulletin, 137, 1–164. Pace, D. W., Gastaldo, R. A., & Neveling, J. (2009). Aggradational and degradational landscapes in the Early Triassic of the Karoo Basin and evidence for dramatic climate shifts following the P/Tr Event. Journal of Sedimentary Research, 79, 276–291. Partridge, T. C. (1993). Warming phases in southern Africa during the last 150 000 years: An overview. Palaeogeography, Palaeoclimatology, Palaeoecology, 101, 237–244. Partridge, T. C. (1998). Of diamonds, dinosaurs and diastrophism: 150 million years of landscape evolution in southern Africa. South African Journal of Geology, 101, 167–184. Partridge, T. C., Botha, G. A., & Haddon, I. G. (2006). Cenozoic deposits of the interior. In M. R. Johnson, C. R. Anhaeusser, & R. J. Thomas (Eds.), The geology of South Africa (pp. 585–604). Geological Society of South Africa. Partridge, T. C., & Maud, R. R. (1987). Geomorphic evolution of southern Africa since the Mesozoic. South African Journal of Geology, 90, 179–208. Partridge, T. C., & Maud, R. R. (2000). Macro-scale geomorphic evolution of southern Africa. In T. C. Partridge & R. R. Maud (Eds.), The Cenozoic of southern Africa (pp. 3–18). Oxford University Press. Pettijohn, F. J. (1975). Sedimentary rocks. Harper and Row. Roberts, D. L., Botha, G. A., Maud, R. R., & Pether, J. (2006). Coastal Cenozoic deposits. In M. R. Johnson, C. R. Anhaeusser, & R. J. Thomas (Eds.), The Geology of South Africa (pp. 605–628). Geological Society of South Africa. Rubidge, B. S. (2005). Re-uniting lost continents-fossil reptiles from the ancient Karoo and their wanderlust. South African Journal of Geology, 108, 135–172.

45 Rust, D. J., & Summerfield, M. A. (1990). Isopach and borehole data as indicators of rifted margin evolution in southwestern Africa. Marine and Petroleum Geology, 17, 277–287. Scheiber-Enslin, S. E. (2016). Integrated geophysical investigation of the Karoo Basin, South Africa. Ph.D. thesis, University of Witwatersrand, Johannesburg. Schultze, R. E. (1997). South African atlas of agrohydrology and climatology (pp. 1–176). Water Research Commission Report TT82/96. Shultze, R. E., & Kunz, R. P. (2011). Climate change and annual precipitation statistics: A 2011 perspective. In R. E. Shultze (Ed.), A perspective on climate change and the South African water sector (pp. 91–96). Water Research Commission Report No TT 518/12. South African Committee for Stratigraphy (SACS). (1980). Stratigraphy of South Africa. Part 1. (Compiler L. E. Kent). Lithostratigraphy of the Republic of South Africa, South West Africa/Namibia, and the Republics of Bophuthatswana, Transkei and Venda. Handbook 8, Geological Survey of South Africa, Pretoria. Srivastava, P., Brook, G. A., Marais, E., Morthekai, P., & Singhvi, A. K. (2006). Depositional environment and OSL chronology of the Homeb silt deposits, Kuiseb River, Namibia. Quaternary Research, 65, 478–491. Tankard, A., Welsink, H., Aukes, P., Newton, R., & Stettler, E. (2009). Tectonic evolution of the Cape and Karoo basins of South Africa. Marine and Petroleum Geology, 26, 1379–1412. Tankard, A., Welsink, H., Aukes, P., Newton, R., & Stettler, E. (2012). Geodynamic interpretation of the Cape and Karoo basins, South Africa. In D. G. Roberts & A. W. Bally (Eds.), Regional geology and tectonics: Phanerozoic passive margins, cratonic basins and global tectonic maps (pp. 869–945). Elsevier. Thomson, K. (1999). Role of continental break-up, mantle plume development and fault reactivation in the evolution of the Gamtoos Basin, South Africa. Marine and Petroleum Geology, 16, 409–429. Tinker, J., de Wit, M., & Brown, R. (2008). Mesozoic exhumation of the southern Cape, South Africa, quantified using apatite fissiontrack thermochronology. Tectonophysics, 455, 77–93. Tooth, S. (2015). Changes in fluvial systems during the Quaternary. In J. Knight & S. W. Grab (Eds.), Quaternary environmental change in southern Africa (pp. 170–187). Cambridge University Press. Tooth, S., Hancox, P. J., Brandt, D., McCarthy, T. S., Jacobs, Z., & Woodborne, S. (2013). Controls on the genesis, sedimentary architecture, and preservation potential of dryland alluvial succession in stable continental interiors: Insights from the incising Modder River, South Africa. Journal of Sedimentary Research, 83, 541–561. Tooth, S., & McCarthy, T. S. (2004). Anabranching in mixed bedrockalluvial rivers: The example of the Orange River above Augrabies Falls, Northern Cape Province, South Africa. Geomorphology, 57, 235–262. Tooth, S., Rodnight, H., Duller, G. A. T., McCarthy, T. S., Marren, P. M., & Brandt, D. (2007). Chronology and controls of avulsion along a mixed bedrock-alluvial river. Geological Society of America Bulletin, 119, 452–461. Turner, B. R. (1975). The stratigraphy and sedimentary history of the Molteno Formation in the Main Karoo Basin of South Africa and Lesotho. Ph.D. thesis, University of Witwatersrand, Johannesburg. Turner, B. R. (1999). Tectonostratigraphical development of the Upper Karoo foreland basin: Orogenic unloading versus thermally induced Gondwana rifting. Journal of African Earth Sciences, 28, 215–238. Turner, B. R., & Thomson, K. (1998). Tectonostratigraphical development of the Upper Karoo foreland basin: Orogenic unloading versus Gondwana rifting. Implications for the Mesozoic evolution of the Falklands plateau and west Gondwana. Journal African Earth Science, 27, 201–202.

46 Van Niekerk, C. D. (1977). Die geologie noord-oos van Graaff-Reinet. M.Sc. thesis, University of the Orange Free State, Bloemfontein. Veevers, J. J., Cole, D. I., & Cowan, E. J. (1994). Southern Africa: Karoo Basin and Cape Fold Belt. Memoir of the Geological Society of America, 184, 223–279. Viglietti, P., Frei, D., Rubidge, B. S., & Smith, R. M. H. (2018). U-Pb detrital zircon dates and provenance data from the Beaufort Group (Karoo supergroup) reflect sedimentary recycling and air-fall tuff deposition in the Permo-Triassic Karoo Foreland Basin. Journal of African Earth Sciences, 143, 59–66. Viglietti, P., Rubidge, B. S., & Smith, R. M. H. (2017a). New Late Permian tectonic model for South Africa’s Karoo Basin: Foreland

J. Neveling tectonics and climate change before the end-Permian crisis. Nature Scientific Reports, 7, 10861. Viglietti, P., Rubidge, B. S., & Smith, R. M. H. (2017b). Revised lithostratigraphy of the Upper Permian Balfour and Teekloof Formations of the Main Karoo Basin, South Africa. South African Journal of Geology, 120, 45–60. Watkeys, M. K. (2006). Gondwana breakup: A South African perspective. In M. R. Johnson, C. R. Anhaeusser, & R. J. Thomas (Eds.), The geology of South Africa (pp. 531–539). Geological Society of South Africa. Wopfner, H. (2002). Tectonic and climatic events controlling deposition in Tanzanian Karoo basins. Journal of African Earth Science, 34, 167–177.

Chapter 5

Isotopic Evidence for the Geographic Origin, Movement and Diet of the Hofmeyr Individual Sandi R. Copeland, Vaughan Grimes, Johann Neveling, Julia A. Lee-Thorp, Frederick E. Grine, Zhaoping Yang, Christopher Dean, and Michael P. Richards

Abstract The Hofmeyr skull is a singular and important fossil find, dating to a period in the late Pleistocene when the human fossil record is extremely poorly represented in southern Africa. However, its lack of contextual evidence is a serious impediment to a complete appreciation of the

S. R. Copeland (&) Environmental Stewardship Group, Los Alamos National Laboratory, P.O. Box 1663 Los Alamos, NM 87545, USA e-mail: [email protected] V. Grimes Department of Archaeology, Memorial University, St. John’s, NL A1C 5S7, Canada Department of Earth Sciences, Memorial University, St. John’s, NL A1C 5S7, Canada J. Neveling Geoscience Mapping Unit, Council for Geoscience, Private Bag X112, Pretoria, 0001, South Africa J. A. Lee-Thorp School of Archaeology, University of Oxford, Oxford, OX1 3TG, UK F. E. Grine Department of Anthropology, Stony Brook University, Stony Brook, NY 11794, USA Department of Anatomical Sciences, Stony Brook University School of Medicine, Stony Brook, NY 11794, USA Z. Yang Geological Survey of Canada, 601 Booth St, Ottawa, ON K1A 0E8, Canada C. Dean Centre for Human Evolution Research, Natural History Museum, Cromwell Road, London, SW7 5BD, UK Department of Cell and Developmental Biology, University College London, Gower Street, London, WC1E 6BT, UK M. P. Richards Department of Archaeology, Simon Fraser University, Burnaby, BC V5A 1S6, Canada

specimen, requiring that every bit of information possible be extracted from the fossil itself. Here we investigate the mobility and dietary ecology of the Hofmeyr individual by analyzing the strontium (87Sr/86Sr), carbon (d13C), and oxygen (d18O) isotopic composition from molar tooth enamel. We compared the former against 87Sr/86Sr in the parietal bone, associated endocranial matrix, and bioavailable strontium isotopes from an 80 km radius of the find location. The strontium isotope data are consistent with a scenario in which the Hofmeyr individual lived in the study area as a youth. The d18O value is consistent with expectations for an individual from the Karoo in a cooler Pleistocene climate. The d13C value suggests that most dietary carbon was from C3 sources, with c. 15–20% from C4 plants (grasses or sedges) and/or the animals that consumed those plants.





Keywords Bioavailable strontium Oxygen isotopes Enamel structure

Carbon isotopes Karoo



Introduction When the Hofmeyr skull was discovered in the early 1950s eroding from a dry riverbed, no associated fauna or artifacts were noted or collected with it. Within ten years subsequent to finding the skull, a small anti-erosion weir was constructed just downstream from the location at which the skull had been found. The weir caused the streambed to fill up with sediment and thereby cover up the sediment layers that had yielded the skull. Without the ability to conduct further study of the skull’s original context, it is a challenge to understand the environmental and ecological context in which the individual lived. Despite the lack of directly associated fauna and sedimentary context, we can potentially unearth aspects of the Hofmeyr individual’s diet and geographic movements using

© Springer Nature Switzerland AG 2022 F. E. Grine (ed.), Hofmeyr: A Late Pleistocene Human Skull from South Africa, Vertebrate Paleobiology and Paleoanthropology, https://doi.org/10.1007/978-3-031-07426-4_5

47

48

isotopic analyses of tooth enamel, with reference to background isotopic studies from the region. Strontium and oxygen isotopes can potentially help to inform the locations the individual lived during youth (Ericson, 1985), and oxygen isotopes can also shed light on paleoclimatic conditions. Carbon isotopes can indicate the relative proportions of C3 and C4 foods consumed and may help elucidate the individual’s consumption of certain plant and animal foods. Nitrogen isotopes could potentially identify trophic level – the amounts of high protein animal food in the diet (O’Connell, 2017). We sampled parietal bone, dentine and enamel of the Hofmeyr mandibular third molar tooth (RM3) as well as associated matrix for isotopic composition. We undertook a study of the bioavailable strontium isotopes across the region by analyzing predominantly plants, as well as naturally occurring water, soils, and animals. Strontium isotopes from the Hofmeyr individual are interpreted in light of the bioavailable strontium results. Oxygen isotopes are interpreted with reference to precipitation patterns of the region and considerations of the glacial conditions during the Late Pleistocene. The carbon isotope results are interpreted in light of our understanding of the distributions of C3 and C4 plants in the Karoo region of southern Africa (Vogel et al., 1978). Unfortunately, the attempt to extract dentine collagen in order to determine the nitrogen isotope composition failed due to lack of preserved organic material in the specimen.

Background The town of Hofmeyr is in the Eastern Karoo of South Africa, a semi desert region lying north of the escarpment at 1250 m above sea level, with low rainfall, and extremes of heat and cold in the austral summer and winter, respectively (Schulze, 1972) (Figs. 5.1 and 5.2). The skull discovery location is about 15 km northeast of Hofmeyr, where the eastern edge of the Nama-Karoo biome and the western edge of the Grassland biome intersect (Rutherford et al., 2006). The bioregions in this vicinity, which are defined by plant diversity, are Upper Karoo, Dry Highveld Grassland, and Sub-Escarpment Grassland (Rutherford et al., 2006). The region has been used extensively for commercial sheep farming for more than a century, and prior to that was used by Khoi pastoralists for hundreds of years. Several of us visited the location at which the Hofmeyr skull was found in January 2008 to assess the local environment and to conduct the bioavailable strontium isotope study (Fig. 5.3). The site of recovery has been obscured by sediment that has backed up behind an anti-erosion weir that was constructed a little downstream of the site not long after its discovery (Fig. 5.4).

S. R. Copeland et al.

Two of the sampling localities that represent different geological settings are illustrated in Figs. 5.5 and 5.6. The Hofmeyr individual was an adult at the time of death. The skull’s relatively good state of preservation indicates that the length of time between death and incorporation of sediment within the skull was likely to have been short, and that it was not transported any distance between the location at death and its discovery site (Grine et al., 2007).

Strontium Isotopes The ratio of two naturally occurring isotopes of strontium, 87 Sr/86Sr, varies in bedrock as a result of differences in initial 87 Rb content and age. 87Rb decays into 87Sr with a half-life of 49 billion years, while 86Sr is stable. Therefore, rocks that are older and/or had greater initial concentrations of Rb initially will have higher 87Sr/86Sr ratios (Faure & Powell, 1972). The particular 87Sr/86Sr of bedrock mineral components are passed into soils derived from the bedrock, and groundwater. Additional trace amounts of strontium, potentially with different 87Sr/86Sr ratios, may be introduced to the soils and biosphere from precipitation and dust (Grousset et al., 1992; Vitousek et al., 1999). The resulting 87 Sr/86Sr ratio in the biosphere that is available to plants and animals is known as the bioavailable strontium. Local bioavailable 87Sr/86Sr may vary across the landscape with bedrock type and water transport – so, for instance, river borne or flood sediments would represent a mixture of influences. Since strontium is chemically similar to calcium, it is incorporated into plants and then in trace quantities into animal tissues including tooth enamel, dentine, and bone. Once tooth enamel is fully mineralized during the individual’s youth, its strontium isotope composition is fixed, and reflects the bioavailable Sr isotope ratios from the location of the animal’s youth.

Oxygen Isotopes The oxygen isotope composition of human tooth enamel (expressed as d18O)1 is determined mainly by the oxygen ingested in liquid water, and to a lesser extent from carbohydrates or lipids in solid foods (Luz et al., 1984; Podlesak et al., 2008). Oxygen isotope ratios in the environment are Stable isotope ratios are by convention expressed in the d notation, in parts per thousand (per mille or ‰) relative to an international standard, as d xZ = (Rs/Rref – 1)  1000, where R = isotope ratio (13C/12C, 18 16 O/ O). For carbon isotopes the standard is the marine limestone PDB (now VPDB). Oxygen isotopes may be expressed relative to VPDB or to Standard Mean Ocean Water (VSMOW). Negative values denote that the sample has lower abundances of the heavier isotope than does the standard. 1

5

Isotopic Evidence for Movement and Diet

49

Fig. 5.1 Overview of the region near the discovery location of the Hofmeyr skull. The foreground geology is Quaternary alluvium, and the mountains are capped with Molteno Formation sandstones and dolerite sills

Fig. 5.2 Landscape view near the Hofmeyr skull discovery location. The foreground is Quaternary alluvium, and the mountains are capped with Molteno Formation sandstone and dolerite sills

50

S. R. Copeland et al.

Fig. 5.3 The weir (sampling location H1) that was built just downstream of the location of the Hofmeyr skull discovery

strongly influenced by precipitation, which varies with distance from primary moisture source, altitude, temperature, and amount of precipitation (Rozanski et al., 1993). Evaporation from standing water and soils, and evapotranspiration in plants under low humidity conditions leads to subsequent 18O-enrichment in local water sources and in plants (Barbour, 2007). The primary moisture source for precipitation in the Eastern Cape is from the adjacent SW Indian Ocean, with rainfall occurring in the Austral summer (Tyson & Preston-Whyte, 2000). The d18O composition of carbonate and phosphate in tooth enamel reflects that of the blood bicarbonate from which it is formed during enamel formation (Passey et al., 2005). In the case of bulk enamel samples of human teeth, d18O values reflect annual averages from water and food. In the vicinity of Hofmeyr and the Eastern Karoo, modern drinking water d18O ranges from about −5.2 to −0.5‰ relative to the VSMOW standard (West et al., 2014).

Carbon Isotopes The primary distinctions in carbon isotopes at the base of all terrestrial foodwebs reside in plant photosynthesis. Plants reliant on the ancient C3 pathway – comprising all trees, most shrubs, herbs, grasses and sedges in temperate regions with cool growing seasons – are strongly depleted in 13C compared to plants, such as grasses and sedges, that follow the C4 pathway. In an open African woodland, the modern range and global average for C3 vegetation of d13C = −27.0 ± 1.1‰, does not overlap with that of the C4 plants, at −12.5 ± 1.0‰ (Codron et al., 2013). In many African environments where the growing (i.e., rainfall) season is warm, the grassy vegetation is dominated by C4 taxa, but in the Eastern Cape, cooler south-facing slopes may also include C3 grasses (Cowling, 1983). These values are not immutable, as modern plant values are shifted by +2‰

5

Isotopic Evidence for Movement and Diet

51

Fig. 5.4 Standing at the approximate location of the Hofmeyr skull discovery site (sampling location H4), which is now covered with infilled sediment. From left, Fred Grine, the sons of J. Moolman, Michael Richards, Sandi Copeland, and Johann Neveling

(current estimate, cf. Keeling et al., 2017) due to the Suess effect on d13CCO2, and they should be adjusted for application in ancient settings. An additional adjustment of 0.5‰ accounts for the effect of lower pCO2 in Glacial settings on C3 plants (Hare et al., 2018). These considerations change the d13C means for C3 and C4 plants in a Glacial setting to −24.5‰ and −10.5‰ respectively, but do not alter the distinction between the two pathways by much. The distinctions in carbon isotopes in plant types are incorporated into the tissues of consumers including humans, with an offset that has been estimated as 13‰ (Passey et al., 2005). As for oxygen isotopes, d13C in tooth enamel carbonate derives from blood bicarbonate during enamel formation and therefore reflects an integrated dietary signal from all food sources (Ambrose & Norr, 1993; Passey et al., 2005) at the time of tooth mineralization.

Geology The bedrock geology of this region (Table 5.1) is comprised of Permian (299–251 Ma) to Triassic (251–199 Ma) rocks of the Karoo Supergroup, intrusive Jurassic (201–145 Ma) dolerite

dykes and sills (Johnson et al., 2006), and alluvial Quaternary sediments (2.58 Ma – present, following Gibbard et al., 2010). The Karoo Supergroup formed from sediments filling the inland Karoo Basin over approximately 120 million of years. The progressively younger Balfour, Katberg, and Burgersdorp Formations that underlie the Hofmeyr locality form part of the Beaufort Group, and comprise sandstones, siltstones, and mudstones that accumulated in this basin as fluvial deposits. The mountains 15 km to the east are largely made up of the fluvial sandstones and siltstone of the Molteno Formation, which forms part of the informal Stormberg Group (Table 5.1). The Balfour, Katberg, Burgersdorp, and Molteno Formations derive primarily from the Paleozoic rocks forming the Cape Fold Belt to the south (e.g., Garzanti et al., 2014; Hanson, 2003). The sedimentary sequence forming the Cape Fold Belt was in turn derived from the much older Archaean and Proterozoic rocks to the north. Those ancient rocks have had more time for 87Rb to decay into 87Sr and therefore the source rocks for the majority of Karoo Basin sediments are expected to have higher strontium isotope values compared to modern seawater (0.7092) (Faure & Powell, 1972). The intrusive dolerite dykes that are scattered across the study area and throughout the vast Karoo are associated with

52

S. R. Copeland et al.

the Jurassic basaltic flows that cap the Drakensberg Mountains (Duncan & Marsh, 2006) (Fig. 5.7). In general, dolerites and basalts are expected to have relatively low 87 Sr/86Sr values (Faure & Powell, 1972). The modern Karoo landscape is the product of extensive erosional events that occurred during the Cretaceous and Cenozoic (Partridge & Maud, 2000). As a result, the landscape in the vicinity of Hofmeyr to Cradock is speckled with dolerite dykes and scattered with local concentrations of alluvial sediments that accumulated from the surrounding bedrock during the Quaternary. Overlying this is a shallow soil and fluvial deposits of small drainage networks draining the highlands to the east.

Methods Hofmeyr Skull Samples

Fig. 5.5 Collecting plant samples at a localized outcrop of dolerite (sampling location H5)

We sampled the lower right permanent third molar (RM3) of the Hofmeyr skull carbon and oxygen isotope composition in the enamel, and for strontium isotope composition in the enamel and dentine. For this purpose, the crown and root were

Fig. 5.6 Collecting plant samples at a road-cut outcrop of the Balfour Formation (sampling location H8)

5

Isotopic Evidence for Movement and Diet

53

Table 5.1 Ages and descriptions of the major geological units near Hofmeyr

extracted without damage from the alveolar wall. Unlike dentine and bone, enamel is notably non-susceptible to diagenetic uptake of strontium and carbon owing to its essentially non-porous structure with relatively large crystals and a low organic component (Budd et al., 2000; Hoppe et al., 2003; Lee-Thorp & Sponheimer, 2003). Because of this, comparison of enamel strontium isotope values with those of dentine, bone, and the burial environment (e.g., the endocranial matrix) can help to assess potential diagenetic alteration and to ascertain bioavailable 87Sr/86Sr of the burial environment. The carbon and oxygen isotope sampling and analysis of the RM3 enamel was performed in the Department of Archaeological Sciences, Bradford University, U.K. Approximately 5 mg of enamel powder was abraded from the exposed inner enamel of the enamel wedge, close to the enamel-dentine junction, using a diamond-tipped ball burr drill-bit. The powder was pretreated in 1.8 ml centrifuge tubes first with 1.5% sodium hypochlorite solution for 30 min to remove organic contaminants, and then with 0.1 M acetic acid for 10 min following standard methods (Lee-Thorp & Sponheimer, 2003; Sponheimer, 1999). After thorough rinsing in distilled water, the sample was freeze dried and placed in individual reaction vessels for reaction with phosphoric acid at 70 °C in a Gasbench, purified by cryogenic distillation and the resultant CO2 analyzed for

13

C⁄12C ratios using a Thermo Delta V Advantage mass spectrometer. Carbon isotope ratios (13C⁄12C) are expressed as d13C values in parts per thousand (‰) relative to the PDB standard. Oxygen isotope ratios (18O⁄16O) are also generated from this method, and are expressed as d18O values in parts per thousand relative to the SMOW standard. Analytical error as determined by multiple measurements of standards was less than 0.2‰ for both d13C and d18O. For the strontium isotope analysis of the molar, we took two sets of samples from enamel and dentine. In the first instance, strontium isotope values were measured on the distal surface of the enamel cap (external enamel) and the mesiolingual root surface immediately adjacent to the sampled region of enamel (Fig. 5.8). These measurements were obtained using laser ablation multicollector inductively coupled plasma mass spectrometry (LA-MC-ICP-MS) in 2007 at the Department of Human Evolution, Max Planck Institute for Evolutionary Anthropology, in Leipzig, Germany. The operating parameters for the LA-MC-ICP-MS are provided in Table 5.2. The tooth was placed into the laser ablation chamber of the UP213 deep-UV YAG laser ablation system (New Wave Research, Inc.) and ablated in four lines on the external enamel surface, and in two lines on the root for dentine samples. The resultant ablated material was entrained in a stream of helium (He) and transferred to the inlet system of the Thermo Fisher Neptune

54

S. R. Copeland et al.

Fig. 5.7 Map of study area showing geology and sampling sites and results of the bioavailable strontium isotope study based on plants. At this scale, the underlying bedrock is depicted and not the Quaternary alluvium

multicollector inductively coupled plasma mass spectrometer (MC-ICP-MS). The locations of the laser scans are shown in Fig. 5.8. The raw data from the LA-MC-ICP-MS Sr isotope analysis of the first set of Hofmeyr tooth samples were re-processed using the data reduction procedure by Yang et al. (2011) and a Ca argide and Ca dimer interference correction procedure from Woodhead et al. (2005). A detailed description of the re-processing methods and results are described in the Notes section at the end of the chapter. In the second instance, a small wedge of enamel and dentine was removed from the distal end of the tooth

(Fig. 5.8B) and using the same LA-MC-ICP-MS set-up, four samples were taken from the internal enamel immediately adjacent to the enamel-dentine junction (EDJ), and two samples were taken from the dentine (Fig. 5.9). The enamel samples covered the length of the preserved EDJ which, as discussed below, entailed about 47 incremental lines (the Brown Striae of Retzius, or BSR) that manifest on the outer enamel surface as perikymata. Enamel adjacent to the EDJ preserves the initial environmental strontium present during tooth development more reliably than that of external enamel (Müller et al., 2019). These data were therefore used to verify the 87Sr/86Sr and variability of the first set. The

5

Isotopic Evidence for Movement and Diet

55

Institute for Human Evolution using the same Sr extraction method described above for the bone sample.

Bioavailable Strontium Isotope Study

Fig. 5.8 Occlusal views of the lower right third molar before (A) and after (B) a small wedge was removed from the distal end. On the lingual view of the M3 (C) black lines (added in Photoshop) show the approximate locations of the sampling of the external enamel (SEVA5835_2 through SEVA5835_5) and dentine (root) (SEVA5836_2 and SEVA5836_3)

strontium isotope data from the wedge were not subjected to re-processing for Ca argide and Ca dimer interference. Detailed observations on aspects of the enamel crown formation are presented below in the Results and Discussion. A small fragment of bone from the left parietal was analyzed for strontium isotopes using the column/anion exchange solution method in the MC-ICP-MS at the Department of Human Evolution, Max Planck Institute for Evolutionary Anthropology, in Leipzig, Germany, following the method of Copeland et al. (2008). A piece of bone weighing 23.6 mg was dissolved in 14.3 M nitric acid, dried down and re-dissolved in 1 ml 3 M HNO3. This solution was added to pre-cleaned 2 ml columns containing Eichrom Sr specific resin and subjected to multiple washes of 3 M HNO3 before the purified Sr was eluted using 2 ml of milliQ water (>18 Mega Ohm). The Sr containing solutions were acidified with HNO3 to a final concentration of 2% and analyzed on the MC-ICP-MS. Finally, a diamond drill bit was used to powder and collect some of the hardened endocranial matrix (breccia) preserved inside the cranium. The powder was divided into two 30 mg aliquots, one was leached in water and one in 0.1 M acetic acid for three hours. Both leachates were analyzed using solution MC-ICP-MS at the Max Planck

There are eight major geological units within an 80 km radius of the Hofmeyr skull’s discovery location according to the 1:250,000 geological maps produced by the Council for Geoscience (previously the Geological Survey) of the Republic of South Africa (3124 Middelburg, 3126 Queenstown, and 3224 Graaff-Reinet). Six of these geological units were sampled in this study: the Balfour, Katberg, Burgersdorp, and Molteno Formations, the dolerite dykes/sills, and the Quaternary alluvium (Table 5.1). The Elliot and Clarens Formations and Drakensberg basalts in the far northeastern corner of the radius were not sampled. In order to characterize the biologically available 87 Sr/86Sr across the region, we sampled mainly plants from 19 sampling localities, designated H1–H19, within this 80 km radius (Table 5.3). We chose our sampling localities to ensure that each geological unit is represented by at least two localities, and in places as far away from human development and agriculture as possible, so as to avoid extraneous strontium inputs such as those from fertilizers (Thomsen & Andreasen, 2019). Plants are a practical medium for establishing a bioavailable strontium isoscape because they are easy to collect and are typically the main source of strontium in humans and other mammals (Evans et al., 2009; Snoeck et al., 2020). The variation that may exist between or within plants due to factors such as rooting depth (Hartman & Richards, 2014; Poszwa et al., 2004; Reynolds et al., 2012) can be overcome by combining several plants from each sampling location into a single, averaged sample. Samples from local mammals with small home range sizes (such as rodents) would be advantageous since the animals average all sources of bioavailable strontium (Price et al., 2002), but the practical element of collecting specimens, such as by trapping rodents, can be prohibitive (but see Radloff et al., 2010). Samples of small mammal fossils that were contemporaneous with the Hofmeyr individual would be even more advantageous for confirming ancient local bioavailable 87Sr/86Sr, but no such fossils were recovered with the skull, and the site is now buried due to construction of the weir. At each sampling locality except for H1, we collected the leaves and stems from up to three individual plants (Table 5.4). Plants were snipped and placed into paper bags to allow them to dry. We also sampled water, millipede exoskeletons, and soils from some of the localities on an opportunistic basis. At sampling locality H1, the small weir, the only sample collected was the surface water behind the weir, which was analyzed for

56

S. R. Copeland et al.

Table 5.2 Operating parameters for the LA-MC-ICP-MS MC-ICP-MS operation parameters RF power Cooling gas Aux gas Sample gas Interface cones Mass resolution Nebulizer Cup configuration

1200 W 15 L/min 0.7 L/min 0.9–1.16 L/min Ni Low (400) PFA microflow (50 ul/min) L4(82Kr); L3(83Kr); L2(84Sr); L1(85Rb); Center(86Sr); H1(87Sr); H2(88Sr); H3(89Y)

UP213 UV laser operation parameters Wavelength Ar flow He flow Frequency Spot diameter Pre-ablation Translation rate Raster length Ablation Translation rate Energy density Fluence Raster length Data collection Data block Gas blank time Ablation

213 nm 0.56–0.62 L/min 0.77 L/min 20 Hz 160 µm 150 µm/s 800 µm 2 µm/s ~1.5 mJ 7–9 J/cm2 800 µm 1 block; 150 cycles; 2 s integration 50 cycles 100 cycles

strontium isotope ratios. We also collected and analyzed standing water from the mountain 19 km to the northeast on sediments from the Molteno Formation (sampling locality H6). From three of the sampling localities (H3, H6, and H8), we collected soil samples to analyze for bioavailable strontium. From two sampling localities (H4 and H8) we collected and analyzed 87Sr/86Sr in dried millipede exoskeletons, which are rich in calcium and strontium. The dried plant material collected from sample localities H2–H19 was ashed at 500 °C for 8 h in a muffle oven in the stable isotope preparation laboratory of the Department of Archaeology, University of Cape Town. At the Max Planck Institute for Evolutionary Anthropology, in Leipzig, Germany, about 30–40 mg of plant ash was prepared for Sr extraction, purification, and analysis using the same column/ anion exchange solution method and MC-ICP-MS described for bone, following the method of Copeland et al. (2008). The millipede exoskeleton rings were washed and ultrasonicated in milli-Q water, rinsed in pure acetone, and dried overnight. About 6–10 mg for each sample was prepared for Sr analysis using the same column/anion exchange solution method described above at the Max Planck Institute for Evolutionary Anthropology.

Each soil sample was analyzed by a water leachate and a weak acid leachate. The soil sample was sieved to collect the 12,240 31,520–26,540 31,520–26,540 12,870 15,560 34,160 34,930 31,500–30,680 31,500–30,680 31,500–30,680 13,315–12,918 25,000–22,000 37,750 12,180 11,570 23,500–22,500 26,170 28,510

Douka et al. (2020) Pettitt et al. (2003) Villotte and Henry-Gambier (2010) Henri-Gambier (2001) Formicola et al. (2004) Formicola et al. (2004) Pettit and Trinkaus (2000) Pettit and Trinkaus (2000) Pettit and Trinkaus (2000) Henry-Gambier et al. (2000) Billy (1992) Henry-Gambier (2002) Henry-Gambier (2002) Henry-Gambier (2002) Henry-Gambier (2002) Saos et al. (2020) Svoboda et al. (2002) Svoboda et al. (2002) Svoboda et al. (2002) Svoboda et al. (2002) Svoboda et al. (2002) Arensburg and Bar-Yosef (1973) Mallegni et al. (1999) Mussi (2001); Henry-Gambier (2001) Mussi (2001); Henry-Gambier (2001) Vandermeersch et al. (2013) Vandermeersch et al. (2013) Svoboda et al. (2002) Henry-Gambier et al. (2000) Svoboda et al. (2002) Svoboda et al. (2002) Svoboda et al. (2002) Svoboda et al. (2002) Svoboda et al. (2002) Šefčáková et al. (2011) Belfer-Cohen et al. (2004) Vermeersch (2002) Orschiedt (2000) Orschiedt (2000) Nadel et al. (2000) Svoboda et al. (2002) Soficaru et al. (2007) (continued)

108

F. E. Grine

Table 6.A1 (continued) Site and specimen Peştera Muerii Peştera cu Oase 1 2 3 Prĕdmostí 1 3 4 9 10 18 Sunghir 1 5 Wadi Kubbaniya 1 * The 14C dates reported here are taken from

Age*

Geochronology source

30,150–29,930 Soficaru et al. (2006) 35,200 Trinkaus et al. (2003); Rougier et al. (2007) 34,950 Trinkaus et al. (2003); Rougier et al. (2007) 34,950 Trinkaus et al. (2003); Rougier et al. (2007) 26,780–24,340 Svoboda (2008) 26,780–24,340 Svoboda (2008) 26,780–24,340 Svoboda (2008) 26,870–24,340 Svoboda (2008) 26,780–24,340 Svoboda (2008) 26,780–24,340 Svoboda (2008) 38,900–33,590 Nalawade-Chavan et al. (2014) 38,900–33,590 Nalawade-Chavan et al. (2014) 18,500–17,620 Wendorf et al. (1988); Pazdur et al. (1994) the cited sources and are not calibrated except for those for La Crouzade 6

Sources for Context, Geochronology and Craniometric Data Abri Pataud Bricker, H., & Mellars, P. (1987). Datations 14C de l’Abri Pataud (Les Eyzies, Dordogne) par le rocede “accelerateur-spectometre de masse”. L’Anthropologie, 91, 227–234. Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Douka, K., Chiotti, L., Nespoulet, R., & Higham, T. (2020). A refined chronology for the Gravettian sequence of Abri Pataud. Journal of Human Evolution, 141, 102730. https://doi.org/10.1016/j.jhevol.2019.102730. El Mansouri, M., El Fouikar, A., & Saint-Martin, B. (1996). Correlation between 14C ages and aspartic acid racemization at the Upper Palaeolithic site of the Abri Pataud (Dordogne, France). Journal of Archaeological Science, 23, 803–809. Movius, H., & Vallois, H. (1959). Crane protoMagdalenien et venus du Perigordien final trouves dans l’Abri Pataud. L’Anthropologie, 63, 213–232. Arene Candide Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358.

Pettitt, P.B., Richards, M., Maggi, R., & Formicola, V. (2003). The Gravettian burial known as the Prince (“Il Principe”): new evidence for his age and diet. Antiquity, 77, 15–19. Baousso da Torre Mussi, M. (1986). On the chronology of the burials found in the Grimaldi Caves. Antropologia Contemporanea, 9, 95–104. Mussi, M. (2001). Earliest Italy: An Overview of the Italian Paleolithic and Mesolithic. New York: Kluwer Academic/Plenum. Verneau, R. (1906). Les Grottes de Grimaldi, Monaco. Anthropologie, 2, 57–124. Villotte, S., & Henry-Gambier, D. (2010). The rediscovery of two Upper Palaeolithic skeletons from Baousso da Torre cave (Liguria-Italy). American Journal of Physical Anthropology, 141, 3–6. Villotte, S., Samsel, M., & Sparacello, V. (2017). The paleobiology of two adult skeletons from Baousso da Torre (Bausu da Ture) (Liguria, Italy): implications for Gravettian lifestyle. Comptes Rendus Palevol,16, 462– 473. Barma del Caviglione Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Henry-Gambier, D. (2001). La sépulture des enfants de Grimaldi (Baoussé-Roussè, Italie). Réunion des Musées Nationaux, Paris: CTHS.

6

Description of Skull Morphology

Henry-Gambier, D. (2002). Les fossiles de Cro Magnon (Les Eyzies-de-Tayac, Dordogne): nouvelles données sur leur position chronologique et leur attribution culturelle. Bulletins et Mémoires de la Société d'Anthropologie de Paris, 14, 89–112. Barma Grande Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Formicola, V., Pettit, P.B., & del Lucchese, A. (2004). A direct AMS radiocarbon date on the Barma Grande 6 Upper Paleolithic skeleton. Current Anthropology, 45, 114–118. Graziosi, P. (1942). L’uomo fossile della Barma Grande ai Balzi Rossi nel Museo di Mentone. Archivio per l’Antropologia e l’Etnologia, 72, 22–50. Brno Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Jelínek, J. (1959). Der fossile mensch Brno II: bestattung und skelettmaterial. Anthropos (Brno), 9, 17–22. Matiegka, J. (1929). The skull of the fossil man Brno III, and the cast of its interior. Anthropologie (Prague), 7, 90–107. Oliva, M. (2000). The Brno II Upper Palaeolithic burial. In: W. Roebroeks, M. Mussi, J. Svoboda, & K. Fennema (Eds.), Hunters of the Golden Age (pp. 143–153). Leiden: Leiden University Press. Pettitt, P.B., & Trinkaus, E. (2000). Direct radiocarbon dating of the Brno 2 Gravettian human remains. Anthropos (Brno), 38, 149–150. Bruniquel Henry-Gambier, D., Valladas, H., Tisnérat-Laborde, N., Arnold, M., & Bresson, F. (2000). Datation de vestiges humains présumés du Paléolithique supérieur par la méthode du carbone 14 en spectrométrie de masse par accélerateur. Paleo, 12, 201–212. Guerret, M. (1952). Les squelettes magdaléniens du Musée de Montauban. Recueil Acadamie de Montauban, 58, 48–66.

109

Chancelade Billy, G. (1992). La morphologie de l’homme de Chancelade. un siècle de controversies. In: J.P. Rigaud, H. Laville, & B. Vandermeersch (Eds), Le Peuplement Magdalénien. Paleogéographie, Physique et Humaine (pp. 71–77). Paris: Editions du C.T.H.S. Vallois, H.V. (1941–1946). Nouvelles recherches sur le squelette de Chancelade. L’Anthropologie, 50, 165–202. Cro Magnon Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Henry-Gambier, D. (2002). Les fossiles de Cro Magnon (Les Eyzies-de-Tayac, Dordogne): nouvelles données sur leur position chronologique et leur attribution culturelle. Bulletins et Mémoires de la Société d'Anthropologie de Paris, 14, 89–112. Henry-Gambier, D., Nespoulet, R., & Chiotti, L. (2013). An Early Gravettian cultural attribution for the human fossils from the Cro-Magnon rock shelter (Les EyziesdeTayac, Dordogne). Paleo, 24, 121–138. Vallois, H.V., & Billy, G. (1965). Nouvelles recherches sur les Hommes fossiles de l’abri de Cro-Magnon. Anthropologie (Paris), 69: 47–74, 249–272. La Crouzade Henry-Gambier, D., & Sacchi, D. (1991). Sur quelques restes humains leptolithiques de la grotte de la Crouzade, Aude. L'Anthropologie, 95, 155–179. Henry-Gambier, D., & Sacchi, D. (2008). La Crouzade V-VI (Aude, France): un des plus anciens fossiles d’anatomie moderne en Europe occidentale. Bulletins et mémoires de la Société d’Anthropologie de Paris, 20, 79–104. Saos, T., Grégoire, S., Bahain, J.J., Higham, T., Moigne, A.M., Testu, A., et al. (2020). The Middle and Upper Palaeolithic at La Crouzade Cave (Gruissan, Aude, France): new excavations and a chronostratigraphic framework. Quaternary International, 551, 85–104. Dolní Vĕstonice Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet

110

F. E. Grine

Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Franciscus, R.G., & Vlček, E. (2006). The cranial remains. In: E. Trinkaus, & J. Svoboda (Eds), Early modern human evolution in Central Europe: the people of Dolní Vĕstonice and Pavlov (pp. 63–152). Oxford: Oxford University Press. Svoboda, J. (1995). L’art gravettien en Moravie: contexte dates et styles. L’Anthropologie, 99, 258–272. Svoboda, J., Lŏzek, V. & Vlček, E. (1996). Hunters between East and West. New York: Plenum. Svoboda, J., van der Plicht, J., & Kuželka, V. (2002). Upper Palaeolithic and Mesolithic human fossils from Moravia and Bohemia (Czech Republic): some new 14C dates. Antiquity 76, 957–962. Trinkaus, E., & Svoboda, J. (2006). Early modern human evolution in Central Europe: the people of Dolní Vĕstonice and Pavlov. Oxford: Oxford University Press.

Condemi, S., Capecchi, G., Monti, L., Voisin, J.L., Mounier, A., Ricci, S., et al. (2014). I resti umani rinvenuti a Paglicci (Rignano Garganico e FG): nota preliminare. In: C. Peretto, M. Arzarello, & J. Arnaud (Eds.), Annali del l’Universita’ di Ferrara e Museologia Scientifica e Naturalistica, 10/2, 233–238. Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Mallegni, F. (1992). Squelette de femme d’une sépolture des couches gravettiennes de la Grotta Paglicci près de Rignano Garganico (Pouilles, Italie): Paglicci 25. Rivista di Antropologia, 70, 209–216. Mallegni, F., Bertoldi, F., & Manolis, S. (1999). The Gravettian female human skeleton of Paglicci cave: Paglicci 25 (Rignano Garganico, Puglia, southern Italy). Homo, 50, 127–148.

Ein Gev Arensburg, B., & Bar-Yosef, O. (1973). Human remains from Ein Gev 1, Jordan Valley, Israël. Paléorient, 1, 201–206. Grotte des Enfants Alciati, G., Ascenzi, A., Borgognini, S., Canci, A., Formicola, V., Giacobini, G., et al. (2005). Catalogue of Italian fossil human remains from the Palaeolithic to the Mesolithic. Journal of Anthropological Sciences (Supplemento Monografico), 84, 1–184. Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Henry-Gambier, D. (2001). La sépulture des enfants de Grimaldi (Baoussé-Roussè, Italie). Réunion des Musées Nationaux, Paris: CTHS. Mussi, M. (1986). On the chronology of the burials found in the Grimaldi Caves. Antropologia Contemporanea, 9, 95–104. Mussi, M. (2001). Earliest Italy: an overview of the Italian Paleolithic and Mesolithic. New York: Kluver Academic/Plenum. Grotta Paglicci Borgognini Tarli, S., Foranaciari, G., & Palma di Cesnola, A. (1980). Restes humains des niveaux gravettiens de la Grotte Paglicci. Bulletin et Mémoires de la Société d’Anthropologie de Paris, s. XIII, 7I, 125–152.

Jebel Qafzeh Bar-Yosef, O., & Belfer-Cohen, A. (2004). The Qafzeh Upper Paleolithic assemblages: 70 years later. Eurasian Prehistory, 2, 145–180. Köppel, R. (1935). Das alter der neuendeckten schädel von Nazareth. Biblica, 16, 58–73. Ronen A., & Vandermeersch, B. (1972). The Upper Paleolithic sequence in the cave of Qafza (Israel). Quaternaria, 16, 189–202. Vandermeersch, B., Arensburg, B., Bar-Yosef, O., & Belfer-Cohen, A. (2013). Upper Paleolithic human remains from Qafzeh Cave, Israel. Journal of the Israel Prehistoric Society, 43, 7–21. Yokoyama, Y., Falgueres, C., & De Lumley, M.A. (1997). Datation direct d'un crane proto-cro-magnon de Qafzeh par la spectrometrie gamma non destructive. Comptes Rendues de l’Academie de Science, Paris, Series 2, 324, 773–779. Konĕprusy (Zlatȳ kůň) Churchill, S.E., & Smith, F.H. (2000). Makers of the Early Aurignacian of Europe. Yearbook of Physical Anthropology, 43, 61–115. Prošek, F., Stárka, V., Hrdlička, L., Hokr, Z., Ložek, V., & Dohnal, Z. (1952). Výzkum jeskynĕ Zlatého konĕu Konĕprus. Československý Kras, 5,161–179 Svoboda, J. (2000). The depositional context of the early Upper Paleolithic human fossils from the Koneprusy (Zlaty kun) and Mladeč Caves, Czech Republic. Journal of Human Evolution, 38, 523–536. Svoboda, J., van der Plicht, J., & Kuželka, V. (2002). Upper Palaeolithic and Mesolithic human fossils from

6

Description of Skull Morphology

Moravia and Bohemia (Czech Republic): some new 14C dates. Antiquity, 76, 957–962. Vlček, E. (1957). Pleistocénní človĕk z jeskynĕ na Zlatém koni u Konĕprus. Anthropozoikum, 6, 283–311. Laugerie-Basse Henry-Gambier, D., Valladas, H., Tisnérat-Laborde, N., Arnold, M., & Bresson, F. (2000). Datation de vestiges humains présumés du Paléolithique supérieur par la méthode du carbone 14 en spectromé de masse par accélerateur. Paleo, 12, 201–212. Mladeč Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Frayer, D.W., Jelínek, J., Oliva, M., & Wolpoff, M.H. (2006). Aurignacian male crania, jaws and teeth from the Mladeč caves, Moravia, Czech Republic. In: M. Teschler-Nicola (Ed.), The Mladeč caves and their remains (pp. 185–272). Vienna: Springer. Svoboda, J., van der Plicht, J., & Kuželka, V. (2002). Upper Palaeolithic and Mesolithic human fossils from Moravia and Bohemia (Czech Republic): some new 14C dates. Antiquity, 76, 957–962. Wild, E.M., Teschler-Nicola, M., Kutschera, W, Steier, P., Trinkaus, E., & Wanek, W. (2005). Direct dating of Early Upper Palaeolithic human remains from Mladeč. Nature, 435, 332–335. Wolpoff, M.H., Frayer, D.W., & Jelínek, J. (2006). Aurignacian female crania, jaws and teeth from the Mladeč caves, Moravia, Czech Republic. In: M. Teschler-Nicola (Ed.), The Mladeč caves and their remains (pp. 273–340). Vienna: Springer. Moča Šefčáková, A., Katina, S., Mizera, I., Halouzka, R., Barta, P., & Thurzo, M. (2011): A Late Upper Palaeolithic skull from Moča (the Slovak Republic) in the context of Central Europe. Acta Musei Nationalis Pragae, Series B, Historia Naturalis, 67, 3–24. Nahal Ein Gev Arensburg, B. (1977). New Upper Palaeolithic human remains from Israel. Eretz Israel, 13, 208–215. Belfer-Cohen, A., Davidson, A., Goring-Morris, A.N., Lieberman, D., & Spiers, M. (2004). Nahal Ein Gev 1: a late Upper Palaeolithic site by the sea of Galilee, Israel. Paléorient, 30, 25–45.

111

Nazlet Khater Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Thoma, A. (1984). Morphology and affinities of the Nazlet Khater man. Journal of Human Evolution, 13, 287–296. Vermeersch, P.M. (2002). Two Upper Paleolithic burials at Nazlet Khater. In P.M. Vermeersch (Ed.), Palaeolithic quarrying sites in Upper and Middle Egypt (pp. 273– 282). Leuven: Leuven University Press. Oberkassel Henke, W. (1984). Vergleichend–morphologische Kennzeichnung der Jungpaläolithiker von Oberkassel bei Bonn. Zeitschrift für Morphologie und Anthropologie, 75, 27–44. Henke, W. (1986). Die magdalénienzeitlichen Menschenfunde von Oberkassel bei Bonn. Bonner Jahrbucher, 186, 317–366. Orschiedt, J. (2000). Germany. Hominid remains an update. Anthropologie et Préhistoire, Supplement 10, 74– 75. Ohalo II Hershkovitz, I.S.M., Frayer, D.W., Nadel, D., WishBaratz, S., & Arensburg, B. (1995). Ohalo II H2: a 19,000-year-old skeleton from a water-logged site at the Sea at Galilee, Israel. American Journal of Physical Anthropology, 96, 215–234. Nadel, D., Grinberg, U., Boaretto, E., & Werker, E. (2006). Wooden objects from Ohalo II (23,000 cal BP), Jordan Valley, Israel. Journal of Human Evolution, 50, 644–662. Pavlov Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Franciscus, R.G., & Vlček, E. (2006). The cranial remains. In: E. Trinkaus, & J. Svoboda (Eds), Early modern human evolution in Central Europe: the people of Dolní Vĕstonice and Pavlov (pp. 63–152). Oxford: Oxford University Press. Svoboda, J. (editor) (1997). Pavlov I – Northwest. Upper Paleolithic burial and its settlement context.

112

F. E. Grine

Dolnověstonické studie 4, Archeologický ústav. Brno: AV ČR (Academy of Sciences of the Czech Republic). Svoboda, J., van der Plicht, J. & Kuželka, V. (2002). Upper Palaeolithic and Mesolithic human fossils from Moravia and Bohemia (Czech Republic): some new 14C dates. Antiquity, 76, 957–962. Trinkaus, E., & Svoboda, J. (2006). Early modern human evolution in Central Europe: the people of Dolní Vĕstonice and Pavlov. Oxford: Oxford University Press. Peştera Cioclovina Uscată Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Olariu, A., Skog, G., Hellborg, R., Stenström, K., Faarinen, M., Persson, P., et al. (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. 222–226). Bucharest: Editura Academiei Romaine. Păunescu, A. (2001). Paleoliticul şi mezoliticul din spaţiul transilvan. Editura AGIR, Bucureşti. Rainer, F., & Simionescu, I. (1942). Sur le premier crâne d’homme Paléolithique trouvé en Roumanie. Analele Academiei Romane, Memoriile Seçtiunii Ştiinţifice, Seria III, 17, 489–503. Soficaru, A., Petrea, C., Dobos, A., & Trinkaus, E. (2007). The human cranium from the Peştera Cioclovina Ucastă Romania: context, age, taphonomy, morphology, and paleopathology. Current Anthropology, 48, 611– 619. Peştera Muierii Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Doboş, A., Soficaru, A., & Trinkaus, E. (2010). The prehistory and paleontology of the Pestera Muierii (Romania). Liege: ERAUL 24. Soficaru, A, Doboş, A., & Trinkaus, E. (2006). Early modern humans from the Peştera Muierii, Baia de Fier, Romania. Proceedings of the National Academy of Sciences of the United States of America, 103, 17196– 17201.

Peştera cu Oase Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Rougier, H., Milota, S., Rodrigo, R., Gherase, M., Sarcina, L., Moldovan, O., et al. (2007). Peştera cu Oase 2 and the cranial morphology of early modern Europeans. Proceedings of the National Academy of Sciences of the USA, 104, 1165–1170. Trinkaus, E., Milota, S., Rodrigo, R., Mircea, G., & Moldovan, O. (2003). Early modern human cranial remains from the Peştera cu Oase, Romania. Journal of Human Evolution, 45, 245–253. Trinkaus, E., Moldovan, O., Milota, S., Bilgar, A., Sarcina, L., Athreya, S., et al. (2003). An early modern human from the Peştera cu Oase, Romania. Proceedings of the National Academy of Sciences of the United States of America, 100, 11231–11236. Zilhăo, J., Trinkaus, E., Constantin, S., Milota, S., Gherase, M., Sarcina, L., et al. (2007). The Peştera cu Oase people, Europe’s earliest modern humans. In P. Mellars, O. Bar-Yosef, C. Stringer, & K. Boyle (Eds), Rethinking the human revolution: new behavioural and biological perspectives on the origins and dispersal of modern humans (pp. 249–262). Cambridge: Cambridge University Press. Prĕdmostí Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Matiegka, J. (1934). Homo prĕdmostensís. Fosilní človĕk z Predmosti na Moravĕ, I. Lebky (l'homme fossil de Prĕdmostí en Moravie, Tchecoslovaquie, I. Les cranes.). Praha: Česká Akademie vĕd a Umĕní. Smith, F.H. (1982). Upper Pleistocene hominid evolution in South-Central Europe: a review of the evidence and analysis of trends. Current Anthropology, 23, 667–703. Svoboda, J.A. (2008). The Upper Paleolithic burial area at Prĕdmostí: ritual and taphonomy. Journal of Human Evolution, 54, 15–33. Sunghir Alexeeva, T.I., & Bader, N.O. (2000). Homo sunghirensis. Upper Palaeolithic man: ecological and

6

Description of Skull Morphology

evolutionary aspects of the investigation. Moscow: Scientific World. Crevecoeur, I., Rougier, H., Grine, F.E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: evidence from Nazlet Khater, Peştera cu Oase and Hofmeyr. American Journal of Physical Anthropology, 140, 347–358. Dobrovolskaya, M., Richards, M.P., & Trinkaus, E. (2011). Direct radiocarbon dates for the Mid Upper Paleolithic (eastern Gravettian) burials from Sunghir, Russia. Bulletins et mémoires de la Société d'anthropologie de Paris, 24, 96–102. Jakimov, V.P. (1983). New data on skeletal remains of fossil men on the territory of the Soviet Union. Voprosy antropologii, 71, 14–24. Nalawade-Chavan, S., McCullagh, J., & Hedges, R. (2014). New hydroxyproline radiocarbon dates from Sunghir, Russia, confirm Early Mid Upper Palaeolithic burials in Eurasia. PLoS ONE, 9(1), e76896. https://doi. org/10.1371/journal.pone.0076896. Pettitt, P.B., & Bader, N.O. (2000). Direct AMS radiocarbon dates for the Sunghir mid Upper Palaeolithic burials. Antiquity, 74, 269–270. Trinkaus, E., Buzhilova, A.P., Mednikova, M.B., & Dobrovolskaya, M.V. (2014). The people of Sunghir. Burials, bodies, and behavior in the Early Upper Paleolithic. Oxford: Oxford University Press. Wadi Kubbaniya Pazdur, A., Pazdur, M.F. & Zastawny, A. (1994). Gliwce radiocarbon dates XII. Radiocarbon, 36, 281–302. Stewart, T.D., Angel, J.L., Kelly, J.O., & Tiffany, M. (1986). The Wadi Kubaniya skeleton: a late Paleolithic burial from southern Egypt. In: F. Wendorf, & R. Schild (Eds), The Prehistory of Wadi Kubbaniya. Volume 1. (pp. 49–70). Dallas: Southern Methodist University Press. Wendorf, F., Schild, R., Close, A.E., & Hillman, G.C. (1988). New radiocarbon dates and Late Paleolithic diet at Wadi Kubbaniya, Egypt. Antiquity, 62, 279–283.

References Aboshi, H., Takahashi, T., & Komuro, T. (2010). Age estimation using microfocus X-ray computed tomography of lower premolars. Forensic Science International, 200, 35–40. Acer, N., Sahin, B., Bas, O., Ertekin, T., & Usanmaz, M. (2007). Comparison of three methods for the estimation of total intracranial volume. Stereologic, planimetric, and anthropometric approaches. Annals of Plastic Surgery, 58, 48–53. Acsádi, G. Y., & Nemeskéri, J. (1970). History of human life span and mortality. Akademiai Kiado.

113 Acuña, L. E. B., Aristeguieta, L. M. R., & Mantilla, G. M. (2011). Mandibular fossa depth variations: Relation to age and dental state. International Journal of Morphology, 29, 1189–1194. Adeloye, A., Kattan, K. R., & Silverman, F. N. (1975). Thickness of the normal skull in American blacks and whites. American Journal of Physical Anthropology, 43, 23–30. Adisen, M. Z., Okkesim, A., & Misirlioglu, M. (2018). A possible association between medial depression of the mandibular ramus and maximum bite force. Folia Morphologica, 77, 711–716. Agarwal, S. K., Malhotra, V. K., & Tewari, S. P. (1979). Incidence of the metopic suture in adult Indian crania. Acta Anatomica, 105, 469–474. Ajmani, M. L., Mittal, R. K., & Jain, S. P. (1983). Incidence of the metopic suture in adult Nigerian skulls. Journal of Anatomy, 137, 177–183. Al-Hindi, D. R., Reynolds, A. W., & Henn, B. M. (2022). Genetic divergence within southern Africa during the Later Stone Age. In: F. E. Grine (Ed.), Hofmeyr: A Late Pleistocene human skull from South Africa (pp. 19–28). Springer. AlQahtani, S. J., Hector, M. P., & Liversidge, H. M. (2010). The London Atlas of human tooth development and eruption. American Journal of Physical Anthropology, 142, 481–490. Altman, D. G., & Bland, J. M. (2005). Standard deviations and standard errors. BMJ, 331, 903. Ambrose, S. H. (1998). Chronology of the Later Stone Age and food production in East Africa. Journal of Archaeological Science, 25, 377–392. Asami, R., Aboshi, H., Iwawaki, A., Ohtaka, Y., Odaka, K., Abe, S., et al. (2019). Age estimation based on the volume change in the maxillary premolar crown using micro-CT. Legal Medicine, 37, 18– 24. Asdullah, M., Aggarwal, A., Khawja, K. J., Khan, M. H., Gupta, J., & Ratnakar, K. (2019). An anatomic and radiographic study of medial sigmoid depression in human mandible. Journal of the Indian Academy of Oral Medicine and Radiology, 31, 123–1277. Bailey, R. (1995). The Bantu languages of South Africa: Towards a sociohistorical perspective. In: R. Mesthrie (Ed.), Language and social history: Studies in South African sociolinguistics (pp. 19–38). David Philip. Balzeau, A. (2006). Are thickened cranial bones and equal participation of the three structural bone layers autapomorphic traits of Homo erectus? Bulletins Et Mémoires De La Société D’anthropologie De Paris, 18, 145–163. Balzeau, A. (2007). Variation and characteristics of the cranial vault thickness in Krapina and Western European Neandertals. Periodicum Biologorum, 109, 369–377. Barash, A., Bastir, M., & Been, E. (2015). 3D morphometric study of the mandibular fossa and its implications for species recognition in Homo erectus. Advances in Anthropology, 5, 58691, 11 p. Belfer-Cohen, A., & Goring-Morris, A. N. (2003). Current issues in Levantine Upper Palaeolithic research. In: A. N. Goring-Morris & A. Belfer-Cohen (Eds.), More than meets the eye: Studies on Upper Palaeolithic diversity in the Near East (pp. 1–12). Oxbow Books. Benington, C. R., & Pearson, K. (1912). A study of the Negro skull with special reference to the Congo and Gaboon crania. Biometrika, 81, 292–339. Bertsatos, A., Chovalopoulou, M., Brůžek, J., & Bejdová, Š. (2020). Advanced procedures for skull sex estimation using sexually dimorphic morphometric features. International Journal of Legal Medicine, 13 p. https://doi.org/10.1007/s00414-020-02334-9 Black, W., & Grine, F. E. (2022). The dentition of the Hofmeyr skull. In: F. E. Grine (Ed.), Hofmeyr: A Late Pleistocene human skull from South Africa (pp. 213–233). Springer. Botha, D., & Steyn, M. (2015). Dental health of the late 19th and early 20th century Khoesan. Homo, 66, 187–202.

114 Bousman, C. B., & Brink, J. S. (2018). The emergence, spread, and termination of the Early Later Stone Age event in South Africa and southern Namibia. Quaternary International, 495, 116–135. Boyd, K. L., Villa, C., & Lynnerup, N. (2015). The use of CT scans in estimating age at death by examining the extent of ectocranial suture closure. Journal of Forensic Sciences, 60, 363–369. Bräuer, G. (1988). Osteometrie. In: R. Knußmann (Ed.), Anthropologie. Handbuch der vergleichenden Biologie des Menschen. Band I. Wesen und Methoden der Anthropologie. 1 Teil. Wissenschaftstheorie, Geschichte, morphologischen Methoden (pp. 160–285). Gustav Fischer. Bräuer, G., & Rösing, F. W. (1989). Human biological history in southern Africa. Rassengeschichte Der Menschheit, 13, 6–137. Brown, P. (1994). Cranial vault thickness in Asian Homo erectus and Homo sapiens. Courier Forschungsinstitut Senckenberg, 171, 33– 46. Burrell, C. L., Emery, M. M., Gonzalez, S., & Irish, J. D. (2016). Incidence and variance of the metopic suture. In Proceedings of the 18th Annual Conference British Association of Biological Anthropology and Osteoarchaeology (pp. 66–67). Butaric, L., Buck, L., Balzeau, A., du Plessis, A., & Grine, F. E. (2022). The paranasal sinuses of the Hofmeyr Cranium. In F. E. Grine (Ed.), Hofmeyr: A Late Pleistocene human skull from South Africa (pp. 179–211). Springer. Buyle-Bodin, Y. (1981–1982). Etude biométrique de la surface articulaire supérieure temporo-mandibulaire chez l’homme. L’Anthropologie, 85–86, 660–682. Cameriere, R., Ferrante, L., De Angelis, D., Scarpino, F., & Galli, F. (2008). The comparison between measurement of open apices of third molars and Demirjian stages to test chronological age of over 18-year-olds in living subjects. International Journal of Legal Medicine, 122, 493–497. Carvalho, I. M. M., Damante, J. H., Tallents, R. H., & Ribeiro-Rotta, R. F. (2001). An anatomical and radiographic study of medial depression of the human mandibular ramus. Dentomaxillofacial Radiology, 30, 209–213. Churchill, S. E., & Smith, F. H. (2000). Makers of the Early Aurignacian of Europe. Yearbook of Physical Anthropology, 43, 61–115. Clark, M. J., & McAnear, J. T. (1983). Pseudocyst in the coronoid of the mandible. Oral Surgery Oral Medicine Oral Pathology, 57, 231. Constant, D. A., & Grine, F. E. (2001). A review of taurodontism with new data on indigenous southern African populations. Archives of Oral Biology, 46, 1021–1029. Copeland, S. R., Grimes, V., Neveling, J., Lee-Thorp, J. A., Grine, F. E., Yang, Z., et al. (2022). Isotopic evidence for the geographic origin, movement and diet of the Hofmeyr individual. In: F. E. Grine (Ed.), Hofmeyr: A Late Pleistocene human skull from South Africa (pp. 47–68). Springer. Copes, L. E., & Kimbel, W. H. (2016). Cranial vault thickness in primates: Homo erectus does not have uniquely thick vault bones. Journal of Human Evolution, 90, 120–134. Corruccini, R. S., & Handler, J. S. (1980). Temporomandibular joint size decrease in American blacks: evidence from Barbados. Journal of Dental Research, 59, 1528. Cray, J., Mooney, M. P., & Siegel, M. I. (2011). Cranial suture biology of the Aleutian Island inhabitants. Anatomical Record, 294, 676– 682. Crevecoeur, I. (2008). Étude anthropologique du squelette du Paléolithique supérieur de Nazlet Khater 2 (Égypte). Apport à la compréhension de la variabilité des hommes modernes. Leuven University Press, Egyptian Prehistory Monographs 8. Crevecoeur, I., Rougier, H., Grine, F. E., & Froment, A. (2009). Modern human cranial diversity in the Late Pleistocene of Africa and Eurasia: Evidence from Nazlet Khater, Peştera cu Oase and

F. E. Grine Hofmeyr. American Journal of Physical Anthropology, 140, 347– 358. Crevecoeur, I., Thibeault, A., Boucheneb, L., Matu, M., Maureille, M. B., & Ribot, I. (2022). The Hofmeyr bony labyrinth: Morphological description and affinity. In: F. E. Grine (Ed.), Hofmeyr: A Late Pleistocene human skull from South Africa (pp. 165–178). Springer. Daegling, D. J. (1989). Biomechanics of cross-sectional size and shape in the hominoid mandibular corpus. American Journal of Physical Anthropology, 80, 91–106. Dart, R. A. (1924). The Rooiberg cranium. South African Journal of Science, 21, 556–568. Deacon, J. (1984). The Later Stone Age of southernmost Africa. Oxford: British Archaeological Reports International Series (Vol. 213, pp. 1–441). Oxford. De Boer, H. H., Van der Merwe, A. E., & Soerdjbalie-Maikoe, V. (2016). Human cranial vault thickness in a contemporary sample of 1097 autopsy cases: Relation to body weight, stature, age, sex and ancestry. International Journal of Legal Medicine, 130, 1371–1377. Dekaban, A., & Lieberman, J. E. (1964). Calculation of cranial capacity by linear dimensions. Anatomical Record, 115, 215–219. de Paiva, L. A. S., & Segre, M. (2003). Sexing the human skull through the mastoid process. Revista Do Hospital Das Clínicas Da Faculdade De Medicina Universidade De São Paulo, 58, 15–20. de Villiers, H. (1968). The skull of the South African Negro. Johannesburg: University of the Witwatersrand Press. Drennan, M. R. (1938). Archaeology of the Oakhurst Shelter, George. Part III. The cave dwellers. Transactions of the Royal Society of South Africa, 25, 259–293. Drontschilow, K. (1913). Metrische Studien an 93 Schädeln aus Kamerun. Archiv Für Anthropopologie, 40, 161–183. Durband, A. (2008). Mandibular fossa morphology in the Ngandong and Sambungmacan fossil hominids. Anatomical Record, 291, 1212–1220. Eboh, D. E., Okoro, E. C., & Iteire, K. A. (2016). A cross-sectional anthropometric study of cranial capacity among Ukwuani people of South Nigeria. Malaysian Journal of Medical Science, 23, 72–85. Elamin, F., Hector, M. P., & Liversidge, H. M. (2017). The timing of mandibular tooth formation in two African groups. Annals of Human Biology, 44, 261–272. Elhassan, N., Gebremeskel, E. I., Elnour, M. A., Isabirye, D., Okello, J., Hussien, A., et al. (2014). The episode of genetic drift defining the migration of humans out of Africa is derived from a large East African population. PLoS One, 9, e97674. https://doi.org/10.1371/ journal.pone.0097674 Fagundes, N. J. R., Ray, N., Beaumont, M., Neuenschwander, S., Salzano, F. M., Bonatto, S. L., et al. (2007). Statistical evaluation of alternative models of human evolution. Proceedings of the National Academy of Science of the USA, 104, 17614–17619. Fawcett, C. D., & Lee, A. (1902). A second study of the variation and correlation of the human skull, with special reference to the Naqada crania. Biometrika, 1, 408–467. Fewlass, H., Talamo, S., Wacker, L., Kromer, B., Tuna, T., Fagault, Y., et al. (2020). A 14C chronology for the Middle to Upper Palaeolithic transition at Bacho Kiro Cave, Bulgaria. Nature Ecology & Evolution, 4, 794–801. Franklin, D., Freedman, L., Milne, N., & Oxnard, C. E. (2007). Geometric morphometric study of population variation in indigenous southern African crania. American Journal of Human Biology, 19, 20–33. Funda, A., Cirpan, S., Mas, N. G., Karabekir S., & Magden, A. O. (2014). Anatomic features of metopic suture in adult dry skulls. Journal of Craniofacial Surgery, 25, 1044–1046. Galloway, A. (1937a). The characteristics of the Boskop physical type. American Journal of Physical Anthopology, 23, 31–47.

6

Description of Skull Morphology

Galloway, A. (1937b). The skeletal remains of Mapungubwe. In A. Fouche (Ed.), Mapungubwe: An Ancient Bantu Civilization on the Limpopo (pp. 127–174). Cambridge: Cambridge University Press. Gauld, S. C. (1996). Allometric patterns of cranial bone thickness in fossil hominids. American Journal of Physical Anthropology, 100, 411–426. Gavin, H. M., Sholts, S. B., & Mosca, L. A. (2014). Sexual dimorphism in human cranial trait scores: Effects of population, age, and body size. American Journal of Physical Anthropology, 154, 259–269. Górny, S. (1957). Crania Africana, Uganda. Materiaty i Prace Antropologiczne, 14, 1–400. Green, H., & Curnoe, D. (2005). Mandibular fossa of fossil Australians. Homo, 56, 233–247. Grine, F. E. (2016). The Late Quaternary hominins of Africa: The skeletal evidence from MIS 6-2. In: S. C. Jones & B. A. Stewart (Eds.), Africa from MIS 6-2: Population dynamics and paleoenvironments (pp. 323–381). Springer. Grine, F. E., Bailey, R. M., Harvati, K., Nathan, R. P., Morris, A. G., Henderson, G. M., et al. (2007). Late Pleistocene human skull from Hofmeyr, South Africa, and modern human origins. Science, 315, 226–229. Grine, F. E., Gunz, P., Betti-Nash, L., Neubauer, S., & Morris, A. G. (2010). Reconstruction of the Late Pleistocene human skull from Hofmeyr, South Africa. Journal of Human Evolution, 59, 1–15. Grine, F. E., Wurz, S., & Marean, C. W. (2017). The Middle Stone Age human fossil record from Klasies River Main Site. Journal of Human Evolution, 103, 53–78. Groucutt, H. S., Petraglia, M. D., Bailey, G., Scerri, E. M., Parton, A., Clark-Balzan, L., et al. (2015). Rethinking the dispersal of Homo sapiens out of Africa. Evolutionary Anthropology, 24, 149–164. Gunz, P., & Freidline, S. E. (2022). Cranial form of the Hofmeyr skull: Comparative 3D geometric morphometrics. In: F. E. Grine (Ed.), Hofmeyr: A Late Pleistocene human skull from South Africa (pp. 143–150). Springer. Gustafson, G. (1950). Age determination on teeth. Journal of the American Dental Association, 41, 45–54. Hefner, J. T. (2009). Cranial nonmetric variation and estimating ancestry. Journal of Forensic Science, 54, 985–995. Henn, B. M., Cavalli-Sforza, L. L., & Feldman, M. W. (2012). The great human expansion. Proceedings of the National Academy of Sciences of the USA, 109, 17758–17764. Hinton, R. J. (1981). Changes in articular eminence morphology with dental function. American Journal of Physical Anthropology, 54, 439–455. Hinton, R. J. (1983). Relationships between mandibular joint size and craniofacial size in human groups. Archives of Oral Biology, 28, 37–43. Hinton, R. J., & Carlson, D. S. (1979). Temporal changes in human temporomandibular joint size and shape. American Journal of Physical Anthropology, 50, 325–334. Howells, W. W. (1973). Cranial variation in man. A study by multivariate analysis of patterns of difference among recent human populations. Papers of the Peabody Museum of Archaeology and Ethnology, Harvard University, Vol. 67. Harvard University Press. Howells, W. W. (1989). Skull shapes and the map. Craniometric analyses in the dispersion of modern Homo. Papers of the Peabody Museum of Archaeology and Ethnology, Harvard University, Vol. 79. Harvard University Press. Hublin, J. J., Sirakov, N., Aldeias, V., Bailey, S., Bard, E., Delvigne, V., et al. (2020). Initial Upper Palaeolithic Homo sapiens from Bacho Kiro Cave, Bulgaria. Nature, 581, 299–302. Hwang, K., Hollinger, J. O., Chung, R. S., & Lee, S. I. (2000). Histomorphology of parietal bones versus age and race. Journal of Craniofacial Surgery, 11, 17–23.

115 Irish, J. D., Black, W., Sealy, J., & Ackermann, R. (2014). Questions of Khoesan continuity: Dental affinities among the indigenous Holocene peoples of South Africa. American Journal of Physical Anthropology, 155, 33–44. Ishida, H., & Dodo, Y. (1990). Cranial thickness of modern and Neolithic populations in Japan. Human Biology, 62, 389–401. Ishwarkumar, S., Pillay, P., DeGama, B. Z., & Satyapal, K. S. (2016). An osteometric evaluation of the mandibular condyle in a black KwaZulu-Natal population. International Journal of Morphology, 34, 848–853. Ishwarkumar, S., Pillay, P., DeGama, B. Z., & Satyapal, K. S. (2019). Osteometric and radiological study of the mandibular notch. International Journal of Morphology, 37, 491–497. Ivanhoe, F. (1979). Direct correlation of human skull vault thickness with geomagnetic intensity in some northern hemisphere populations. Journal of Human Evolution, 8, 433–444. Jacob, T. (1967). The morphologic variations of the supramastoid crest and tubercles. Anthropologica, 9, 59–72. Jain, A., & Thenmozhi, M. S. (2018). A study on morphology and morphometry of mandibular fossa. Drug Invention Today, 10, 2289–2292. Janković, I., Ahern, J. C. M., & Smith, F. H. (2016). On some aspects of Neandertal zygomatic morphology. Homo, 67, 89–99. Jerardino, A., Yates, R., Morris, A. G., & Sealy, J. C. (1992). A dated human burial from the Namaqualand coast: Observations on culture, biology and diet. South African Archaeological Bulletin, 47, 75–81. Jöris, O., & Street, M. (2008). At the end of the 14C time scale—The Middle to Upper Paleolithic record of western Eurasia. Journal of Human Evolution, 55, 782–802. Jung, H., & Woo, E. J. (2016). Evaluation of mastoid process as sex indicator in modern white Americans using geometric morphometrics. Journal of Forensic Sciences, 61, 1029–1033. Jung, Y. S., Kim, H. J., Choi, S. W., Kang, J. W., & Cha, I. H. (2003). Regional thickness of parietal bone in Korean adults. International Journal of Oral and Maxillofacial Surgery, 32, 638–641. Kaye, K. W. (1965). Report on a fossil human skull from Hofmeyr, Eastern Cape. B.Sc. honours thesis, University of the Witwatersrand, Johannesburg. Keen, J. A. (1951). Standardization of the technique of cranial capacity determination. South African Journal of Clinical Science, 2, 170–182. Key, C. A., Aiello, L. C., & Molleson, T. (1994). Cranial suture closure and its implications for age estimation. International Journal of Osteoarcheology, 4, 193–207. Khamanarong, K., Tuamsuk, P., Woraputtaporn, W., Namking, M., Sawatpanich, T., Toomsan, Y., et al. (2015). Incidence of metopism in adult Thai skulls. International Journal of Morphology, 33, 51–54. Kitson, E. (1931). A study of the Negro skull with special reference to the crania from Kenya Colony. Biometrika, 23, 271–314. Kuhlwilm, M., Gronau, I., Hubisz, M. J., de Filippo, C., PradoMartinez, J., Kircher, M., et al. (2016). Ancient gene flow from early modern humans into eastern Neanderthals. Nature, 530, 429–433. Le Cabec, A., Tang, N. K., Rbio, V. R., & Hillson, S. (2018). Nondestructive adult age at death estimation: Visualizing cementum annulations in a known age historical human assemblage using synchrotron X-ray microtomography. American Journal of Physical Anthropology, 168, 25–44. Lieberman, D. E. (1996). How and why humans grow thin skulls: Experimental evidence for systemic cortical robusticity. American Journal of Physical Anthropology, 101, 217–236. Liu, W., Martinón-Torres, M., Cai, Y. J., Xing, S., Tong, H. W., Pei, S. W., et al. (2015). The earliest unequivocally modern humans in southern China. Nature, 526, 696–699.

116 Liversidge, H. M. (2008). Timing of human mandibular third molar formation. Annals of Human Biology, 35, 294–321. Lynnerup, N. (2001). Cranial thickness in relation to age, sex and general body build in a Danish forensic sample. Forensic Science International, 117, 45–51. Lynnerup, N., Astrup, J. G., & Sejrsen, B. (2005). Thickness of the human cranial diploe in relation to age, sex and general body build. Head & Face Medicine, 1, 13. https://doi.org/10.1186/1746-160X1-13 Macchiarelli, R., & Sperduti, A. (1998). Mandibular fossa size variation in past and extant human populations. Homo, 49, 172–192. Macdonell, W. R. (1904). A study of the variation and correlation of the human skull, with special reference to English crania. Biometrika, 3, 191–202. Manica, A., Amos, W., Balloux, F., & Hanihara, T. (2007). The effect of ancient population bottlenecks on human phenotypic variation. Nature, 448, 346–348. Manjunath, K. Y. (2002). Estimation of cranial volume: An overview of methodologies. Journal of the Anatomical Society of India, 51, 85–91. Manyaapelo, T. (2007). An odontological analysis of 18th and 19th century burial sites from in and around Cape Town. M.Sc. thesis, University of Cape Town, Cape Town. Marean, C. W., Nilssen, P. J., Brown, K., Jerardino, A., & Stynder, D. (2004). Paleoanthropological investigations of Middle Stone Age sites at Pinnacle Point, Mossel Bay (South Africa): Archaeology and hominid remains from the 2000 field season. PaleoAnthropology, 2004, 14–83. Marquer, P. (1972). Nouvelle contribution a l’étude du squelette des Pygmées occidentaux du Centre Africain comparé a celui des Pygmées Orientaux. Mémoires Du Muséum Nationale D’histoire Naturelle, Série A, 72, 1–122. Marsh, H. E. (2013). Beyond thick versus thin: Mapping cranial vault thickness patterns in recent Homo sapiens. Ph.D. dissertation, University of Iowa, Ames. Martin, R. (1957). Lehrbuch der Anthropologie. In systematischer Darstellung mit besonderer Berüksichtigung der anthropologishen Methoden, 3 Aufl. von Karl Saller, 2 volumes. Gustav Fischer. Masset, C. (1989). Age estimation on the basis of cranial sutures. In M. Y. İşcan (Ed.), Age markers in the human skeleton (pp. 71–103). Springfield (IL): Charles C. Thomas. Meindl, R. S., & Lovejoy, C. O. (1985). Ectocranial suture closure: A revised method for the determination of skeletal age at death based on the lateral-anterior sutures. American Journal of Physical Anthropology, 68, 47–56. Mellars, P. (2006a). Why did modern human populations disperse from Africa ca. 60,000 years ago? A new model. Proceedings of the National Academy of Science of the USA, 103, 9381–9386. Mellars, P. (2006b). A new radiocarbon revolution and the dispersal of modern humans in Eurasia. Nature, 439, 931–935. Mitchell, P. J. (2002). The archaeology of southern Africa. Cambridge: Cambridge University Press. Monnazzi, M. S., Passeri, L. A., Gabrielli, M. F. R., Bolini, P. D. A., de Carvalho, W. R. S., & da Costa, M. H. (2019). Anatomic study of the mandibular foramen, lingula and antilingula in dry mandibles, and its statistical relationship between the true lingula and the antilingula. International Journal of Oral and Maxillofacial Surgery, 41, 74–78. Moreira-Gonzalez, A., Papay, F. E., & Zins, J. E. (2006). Calvarial thickness and its relation to cranial bone harvest. Plastic and Reconstructive Surgery, 117, 1964–1971. Morris, A. G. (1984). An osteological analysis of the proto-historic populations of the northern Cape and western Orange Free State, South Africa. Ph.D. dissertation, University of the Witwatersrand, Johannesburg.

F. E. Grine Morris, A. G. (2022). Lost and found: The discovery and rediscovery of the Hofmeyr skull. In: F. E. Grine (Ed.), Hofmeyr: A Late Pleistocene human skull from South Africa (pp. 9–18). Cham, Switzerland: Springer. Morris, D., Steyn, M., & Ribot, I. (2006). A burial from Driekopseiland, Northern Cape. South African Archaeological Bulletin, 61, 133–141. Nagaoka, T., Shizushima, A., Sawada, J., Tomo, S., Hoshina, K., Sato, H., et al. (2008). Sex determination using mastoid process measurements: Standards for Japanese human skeletons of the medieval and early modern periods. Anthropological Science, 116, 105–113. Nalawade-Chavan, S., McCullagh, J., & Hedges, R. (2014). New hydroxyproline radiocarbon dates from Sungir, Russia, confirm Early-Mid Upper Palaeolithic burials in Eurasia. PLoS One, 9(1), e76896. https://doi.org/10.1371/journal.pone.0076896 Nawrocki, S. P. (1991). A biomechanical model of cranial vault thickness in archaic Homo. Ph.D. dissertation, State University of New York, Binghamton. Neubauer, S. (2022). The endocast of the Late Pleistocene human skull from Hofmeyr. In: F. E. Grine (Ed.), Hofmeyr: A Late Pleistocene human skull from South Africa (pp. 153–163). Cham, Switzerland: Springer. Neubauer, S., Hublin, J. J., & Gunz, P. (2018). The evolution of modern human brain shape. Science Advances, 4, eaao5961. https:// doi.org/10.1126/sciadv.aao5961 Nielsen, R., Akey, J. M., Jakobsson, M., Pritchard, J. K., Tishkoff, S., & Willerslev, E. (2017). Tracing the peopling of the world through genomics. Nature, 541, 302–310. Nikita, E., & Michopoulou, E. (2018). A quantitative approach for sex estimation based on cranial morphology. American Journal of Physical Anthropology, 165, 507–517. Pacher, H. M. (1961). Anthropologische Untersuchungen an den Skeletten der Rudolf Pöchschen Buschmannsammlung. 1 Heft. Herkunft des Sammlungsgutes, Massbefunde und Lichtbilder der Schädel. Rudolf Pöchs Nachlass. In Serie A: Physische Anthropologie, XII Band. Harmann Böhlaus, Austrian Academy of Science. Pagani, L., Lawson, D. J., Jagoda, E., Mörseburg, A., Erikson, A., Mitt, M., et al. (2016). Genomic analyses inform on migration events during the peopling of Eurasia. Nature, 538, 238–242. Pearson, K., & Davin, A. G. (1924). On the biometric constants of the human skull. Biometrika, 16, 328–363. Pensler, J., & McCarthy, J. G. (1985). The calvarial donor site: an anatomic study in cadavers. Plastic and Reconstructive Surgery, 75, 648–651. Perizonius, W. R. K. (1984). Closing and non-closing sutures in 256 crania of known age and sex from Amsterdam (A.D. 1883–1909). Journal of Human Evolution, 13, 201–216. Pfeiffer, S. (2012). Conditions for evolution of small adult body size in southern Africa. Current Anthropology, 53(S6), S383–S394. Pfeiffer, S., & Sealy, J. (2006). Body size among Holocene foragers of the Cape ecozone, southern Africa. American Journal of Physical Anthropology, 129, 1–11. Picq, P. (1984). The temporomandibular joint of the fossil hominids of the genus Homo of the Plio-Pleistocene of East-Africa. Comptes Rendus De L’académie Des Sciences Serie II, 298, 501–506. Powers, R. (1962). The disparity between known age and sex as estimated by cranial suture closure. Man, 84, 52–54. Protsch, R., & Semmel, A. (1978). Zur Chronologie des KelsterbachHomininen. Eiszeitalter Und Gegenwart, 28, 200–210. Rightmire, G. P. (1978). Human skeletal remains from the southern Cape Province and their bearing on the Stone Age prehistory of South Africa. Quaternary Research, 9, 219–230.

6

Description of Skull Morphology

Ribot, I. (2003). Craniometrical analysis of Central and East Africans in relation to history: a case study based on unique collections of known ethnic affiliation. Anthropologie Et Préhistoire, 114, 25–50. Ribot, I. (2004). Differentiation of modern sub-Saharan African populations: Morphology, geography and history. Bulletins Et Mémoires De La Société D’anthropologie De Paris, 16, 143–165. Ribot, I. (2011). A study through skull morphology on the diversity of Holocene African populations in a historical perspective. British Archaeological Reports International Series, S2215. Oxford. Ribot, I., Ghalem, Y., & Crevecoeur, I. (2022). The position of the Hofmeyr skull within Late Pleistocene and Holocene African regional diversity: 2D and 3D morphometric analyses. In: F. E. Grine (Ed.), Hofmeyr: A Late Pleistocene human skull from South Africa (pp. 119–141). Springer. Ricklan, D. E., & Tobias, P. V. (1986). Unusually low sexual dimorphism of endocranial capacity in a Zulu cranial series. American Journal of Physical Anthropology, 71, 285–293. Ried, H. A. (1915). Zur Anthropologie des abflusslosen Rumpfschollenlandes im nordöstlichen Deutsch-Ostafrika. Abhandlungen Des Hamburgischen Kolonialinstitute, 31, 1–129. Rikhotso, R. E., & Munsamy, C. (2017). A morphological study of the lingula in South Africans in relation to sagittal split osteotomy. South African Dental Journal, 72, 408–412. Roksandic, M., Vlak, D., Schillaci, M. A., & Voicu, D. (2009). Applicability of tooth cementum annulation to an archaeological population. American Journal of Physical Anthropology, 40, 583– 588. Ross, A. H., Jantz, R. L., & McCormick, W. F. (1998). Cranial thickness in American females and males. Journal of Forensic Science, 43, 267–272. Sahin, B., Acer, N., Sonmez, O. F., Emirzeoglu, M., Basaloglu, H., Uzun, A., et al. (2007). Comparison of four methods for the estimation of intracranial volume: A gold standard study. Clinical Anatomy, 20, 766–773. Sahni, D., Jit, I., & Sanjeev, N. (2005). Time of closure of cranial sutures in northwest Indian adults. Forensic Science International, 148, 199–205. Saini, V., Srivastava, R., Rai, R. K., Shamal, S. N., Singh, T. B., & Tripathi, S. K. (2012). Sex estimation from the mastoid process among north Indians. Journal of Forensic Sciences, 57, 434–439. Schulter, F. P. (1976). A comparative study of the temporal bone in three populations of Man. American Journal of Physical Anthropology, 44, 453–468. Schwidetzky, I. (1969). Zur Rassengleiderung des subsaharischen Afrika. Statistische Untersuchungen Zu Hiernaux’s Afrikabuch. Homo, 29, 11–34. Sealy, J., & Pfeiffer, S. (2000). Diet, body size, and landscape use among Holocene people in the southern Cape, South Africa. Current Anthropology, 41, 642–655. Sealy, J., Ludwig, B., & Henderson, Z. (2006). New radiocarbon dates for Matjes River Rock Shelter. South African Archaeological Bulletin, 61, 98–101. Šefčáková, A., Katina, S., Mizera, I., Halouzka, R., Barta, P., & Thurzo, M. (2011). Late Upper Palaeolithic skull from Moča (the Slovak Republic) in the context of Central Europe. Acta Musei Nationalis Pragae Ser. B, Historia Naturalis, 67, 3–24. Shea, J. J. (2008). Transitions or turnovers? Climatically forced extinctions of Homo sapiens and Neanderthals in the East Mediterranean Levant. Quaternary Science Reviews, 27, 2253– 2270. Singer, R. (1953). Estimation of age from cranial suture closure. A report on its unreliability. Journal of Forensic Medicine of South Africa, 1, 52–59. Singer, R., & Wymer, J. J. (1982). The Middle Stone Age at Klasies River Mouth in South Africa. Chicago: University of Chicago Press.

117 Smith, F. H. (1984). Fossil hominids from the Upper Pleistocene of Central Europe and the origin of modern Europeans. In: F. Spencer (Ed.), The origins of modern humans: A world survey of the fossil evidence (pp. 137–210). New York: Alan R. Liss. Smith, F. H., & Raynard, G. C. (1980). Evolution of the supraorbital region in Upper Pleistocene fossil hominids from South-Central Europe. American Journal of Physical Anthropology, 53, 589–610. Smith, F. H., Simek, J. F., & Harrill, M. S. (1989). Geographic variation in supraorbital torus reduction during the later Pleistocene (c. 80 000–15 000 BP). In: P. Mellars & C. Stringer (Eds.), The human revolution (pp. 62–108). Princeton University Press. Smith, F. H., Falsetti, A. B., & Simmons, T. (1995). CircumMediterranean biological connections and the pattern of late Pleistocene human evolution. In: H. Ullrich (Ed.), Man and environment in the Palaeolithic (pp. 197–207). Etudes et Recherches Archeologiques de l’Universite de Liège. Soares, P., Alshamali, F., Pereira, J. B., Fernandes, V., Silva, N. M., Afonso, C., et al. (2012). The expansion of mtDNA haplogroup L3 within and out of Africa. Molecular Biology and Evolution, 29, 915–927. Solheim, T. (1990). Dental cementum apposition as an indicator of age. Scandinavian Journal of Dental Research, 98, 510–519. Stewart, T. D. (1933). The tympanic plate and external auditory meatus in the Eskimo. American Journal of Physical Anthropology, 17, 481–496. Steyn, M., Whitelaw, G., Botha, D., Vicente, M., Schlebusch, C. M., & Lombard, M. (2019). Four Iron Age women from KwaZulu-Natal: Biological anthropology, genetics and archaeological context. Southern African Humanities, 32, 23–56. Street, M., Terberger, T., & Orschiedt, J. (2006). A critical review of the German Paleolithic hominin record. Journal of Human Evolution, 51, 551–579. Stynder, D. D. (2006). A quantitative assessment of variation in Holocene Khoesan crania from South Africa’s western, southwestern, southern and southeastern coasts and coastal forelands. Ph.D. thesis, University of Cape Town. Stynder, D. D. (2009). Craniometric evidence for South African Later Stone Age herders and hunter–gatherers being a single biological population. Journal of Archaeological Science, 36, 798–806. Stynder, D. D., Ackermann, R. R., & Sealy, J. C. (2007a). Craniofacial variation and population continuity during the South African Holocene. American Journal of Physical Anthropology, 134, 489– 500. Stynder, D. D., Ackermann, R. R., & Sealy, J. C. (2007b). Early to mid-Holocene South African Later Stone Age human crania exhibit a distinctly Khoesan morphological pattern. South African Journal of Science, 103, 349–352. Stynder, D., Brock, F., Sealy, J., Wurz, S., Morris, A., & Volman, T. (2009). A mid-Holocene AMS 14C date for the presumed Upper Pleistocene human skeleton from Peers Cave, South Africa. Journal of Human Evolution, 56, 431–434. Suazo Galdames, I. C., Zavando, M. D. A., & Smith, R. L. (2009). Performance evaluation as a diagnostic test for traditional methods for forensic identification of sex. International Journal of Morphology, 27, 381–386. Sullivan, W. G., & Smith, A. A. (1989). The split calvarial graft donor site in the elderly: A study in cadavers. Plastic and Reconstructive Surgery, 84, 29–31. Takasaka, T., Kitamura, T., Sugimoto, C., Guo, J., Zheng, H. Y., & Yogo, Y. (2006). Phylogenetic analysis of major African genotype (Af2) of JC virus: Implications for origin and dispersals of modern Africans. American Journal of Physical Anthropology, 129, 465–472. Tellioğlu, A. T., Yilmaz, Ş, Baydar, Ş, Tekdemir, İ, & Elhan, A. H. (2001). Computed tomographic evaluation before cranial bone

118 harvesting to avoid unexpected hazards during aesthetic procedures. Aesthetic Plastic Surgery, 25, 198–201. Terhune, C. E., Kimbel, W. H., & Lockwood, C. A. (2007). Variation and diversity in Homo erectus: A 3D geometric morphometric analysis of the temporal bone. Journal of Human Evolution, 53, 41–60. Todd, T. W. (1924). Thickness of the male white cranium. Anatomical Record, 27, 245–256. Trevor, J. C. (1949). The metrical features of Ashanti and other African Negro crania. Homenaje a Don Luis de Hoyos Sainz (Vol. 1, pp. 317–332). J.R. Sanchez. Tryon, C. A. (2019). The Middle/Later Stone Age transition and cultural dynamics of Late Pleistocene East Africa. Evolutionary Anthropology, 28, 267–282. Truswell, A. S., & Hanson, J. D. L. (1976). Medical research among the !Kung. In R. B. Lee & I. DeVore (Eds.), Kalahari huntergatherers (pp. 166–194). Harvard University Press. Turgut, H. B., Anil, A., Peker, T., Pelin, C., & Gülekon, I. N. (2003). Supraarticular, supramastoid and suprameatal crests on the outer surface of the temporal bone and the relation between them. Surgical & Radiologic Anatomy, 25, 400–407. Twiesselmann, F. (1941). Méthodes pour l’évaluation de I’épaisseur des parois crâniennes. Bulletin du Musée d’Histoire naturelle de Belgique, 17, l–33. Vandevoort, F. M., Bergmans, L., Cleynenbreugel, J. V., Bielen, D. J., Lambrechts, P., Wevers, M., et al. (2004). Age calculation using Xray microfocus computed tomographical scanning of teeth: A pilot study. Journal of Forensic Science, 49, 787–797. Van Reenen, J. F. (1964). Dentition, jaws and palate of the Kalahari Bushmen. Journal of the Dental Association of South Africa, 19, 1–17. Van Reenen, J. F. (1982). The effects of attrition on tooth dimensions of San (Bushmen). In: B. Kurten (Ed.), Teeth: Form, function and evolution (pp. 182–203). New York: Columbia University Press.

F. E. Grine van Vark, G. N. (1984). On the determination of hominid affinities. In: G. N. van Vark & W. W. Howells (Eds.), Multivariate statistical methods in physical anthropology (pp. 323–349). Doordrecht: Reidel. Villa, P., Soriano, S., Tsanova, T., Degano, I., Higham, T. F. G., d’Errico, F., et al. (2012). Border Cave and the beginning of the Later Stone Age in South Africa. Proceedings of the National Academy of Sciences of the USA, 109, 13208–13213. Vogel, J. C., & Marais, M. (1971). Pretoria radiocarbon dates I. Radiocarbon, 13, 378–394. Walrath, D. E., Turner, P., & Bruzek, J. (2004). Reliability test of the visual assessment of cranial traits for sex determination. American Journal of Physical Anthropology, 125, 132–137. Wangai, L., Mandela, P., Butt, F., & Ongeti, K. (2013). Morphology of the mandibular condyle in a Kenyan Population. Anatomy Journal of Africa, 2, 70–79. Welch, B. L. (1939). On confidence limits and sufficiency, with particular reference to parameters of location. Annals of Mathematical Statistics, 10, 58–69. Wilson, M. L., & Lundy, J. K. (1994). Estimated living statures of dated Khoisan skeletons from the south-western coastal region of South Africa. South African Archaeological Bulletin, 49, 2–8. Wish-Baratz, S., Hershkovitz, I., Arensburg, B., Latimer, B., & Jellema, L. M. (1996). Size and location of the human temporomandibular joint. American Journal of Physical Anthropology, 101, 387–400. Wittwer-Backofen, U., Gampe, J., & Vaupel, J. (2004). Tooth cementum annulation for age estimation: Results from a large known-age validation study. American Journal of Physical Anthropology, 123, 119–129. Wu, X. J., & Trinkaus, E. (2014). The Xujiayao 14 mandibular ramus and Pleistocene Homo mandibular variation. Comptes Rendus Palevol, 13, 333–341.

Chapter 7

The Position of the Hofmeyr Skull within Late Pleistocene and Holocene African Regional Diversity: 2D and 3D Morphometric Analyses Isabelle Ribot, Yassmine Ghalem, and Isabelle Crevecoeur

Abstract In light of debates over population continuity or discontinuity for Late Pleistocene and Holocene humans in Africa, we re-assess the morphometric position of the Hofmeyr skull within an expanded sample of past and present-day Africans. Two case studies using 2D and 3D morphometrics are presented. Case Study 1 analyses linear dimensions on 801 skulls and provides an overview of temporal variation. Case Study 2 is an analysis of 3D landmarks for 65 individual temporal bones, which is considered to be a structure with high heritability. The results of Case Study 1 show that, depending upon the region of the cranium that is chosen, Hofmeyr is closest not only to Late Pleistocene Europeans but also to various Late Pleistocene African specimens. Features such as low orbits, a tall mastoid and/or a wide mandibular ramus are similar to Late Pleistocene individuals from North Africa (Nazlet Khater and Wadi Kubbaniya) and West Africa (Iwo Eleru). Case Study 1 also underscores that Hofmeyr is not always on the margin of modern variation, as some of its facial features (e.g., narrow orbits, a curved frontal and large maxillae) are similar to various recent groups from North, South and West Africa. Case Study 2 complements Case Study 1 by showing that Late Pleistocene temporal bones (including Hofmeyr) are not only often different from those of recent groups but are also very diverse themselves. Both case studies serve to highlight unique morphological aspects of the Hofmeyr skull, which presents a mosaic of features that probably reflect different signatures and regional population diversification during the Late Pleistocene.

I. Ribot (&)  Y. Ghalem Département d’Anthropologie, Université de Montréal, Montréal, QC 3T 1N8, Canada e-mail: [email protected] I. Crevecoeur Chargée de Recherche CNRS, Université de Bordeaux, UMR 5199-PACEA, 33615 Pessac Cedex, France



 

 



Keywords Cranium Diversification Human Mandible Morphometrics Population Temporal bone

Introduction The Hofmeyr skull is a key fossil that helps us to fathom the paleoanthropological evidence in sub-Saharan Africa between 50,000 and 30,000 BP, and to analyze population history in relation to genetic hypotheses (e.g., modern human diversification and the Late Pleistocene exodus from Africa between 65,000 and 25,000 BP). Previous craniometric studies have shown that the Hofmeyr skull reflects a unique phenotypic diversity, like other Late Pleistocene African and penecontemporaneous European crania. The present study aims to reassess this finding by employing a considerably larger dataset that represents present day populations from various regions of Africa. By so doing, it may be able to establish a genealogical link for Hofmeyr among modern African groups. It is essential to comprehend continental-wide variation of African crania, especially in light of recent work showing past regional genetic diversity (Schlebusch & Jacobson, 2018), as well as the possible persistence of distinct Late Pleistocene material cultures and “archaic” morphologies into the Holocene in West Africa (Harvati et al., 2011; Stojanowski, 2014). Craniometric studies have extensively researched Middle and Late Pleistocene human fossils from Africa (e.g., Crevecoeur, 2008; Harvati & Weaver, 2006; Harvati et al., 2011; Howells, 1995; Rightmire, 2009; Stringer, 1974a, 1974b). In particular, the Hofmeyr skull, which derives from a key period in African prehistory (ca. 45,000–35,000 BP), has been analyzed morphometrically in several studies (Crevecoeur et al., 2009; Grine et al., 2007, 2010; Gunz & Freidline, 2022). These studies employed large comparative samples to assess morphological variation in a diachronic manner and place the individuals within patterns of past as

© Springer Nature Switzerland AG 2022, corrected publication 2023 F. E. Grine (ed.), Hofmeyr: A Late Pleistocene Human Skull from South Africa, Vertebrate Paleobiology and Paleoanthropology, https://doi.org/10.1007/978-3-031-07426-4_7

119

120

well as modern variation. Recently, both genetic studies (Hammer et al., 2011; Mendez et al., 2013; Stewart & Stringer, 2012) and morphometric analyses (Harvati et al., 2011; Stojanowski, 2014) have suggested some possible introgression of archaic genes into the modern gene pool. Researchers have therefore emphasized the importance of further analyzing both inter- and intra-regional diversity throughout the Pleistocene and Holocene periods (Scerri et al., 2018). The persistence of distinct Late Pleistocene material cultures and archaic morphologies in various regions may suggest that our species evolved in a variety of local habitats within the African continent. This is consistent with genetic hypotheses regarding deep sub-structures of the Late Pleistocene populations and their complex evolutionary histories (Gurdasani et al., 2015; Schlebusch et al., 2012). Concerning the Hofmeyr skull, its mosaic of both plesiomorphic and derived features, as well as its similarities to penecontemporaneous Upper Paleolithic Eurasians, appear to support an hypothesis of Late Pleistocene population diffusion from sub-Saharan Africa towards higher latitudes (Grine et al., 2007). Its morphological and morphometrical features (e.g., long cranium, low orbits, large facial height, prominent glabella, molar megadontia) appeared to be very distinct from modern African diversity and perhaps reflect an ancient genetic stock that did not persist (Crevecoeur et al., 2009). However, an extensive, exhaustive comparison between the Hofmeyr skull and more diverse regional and temporal groups within and outside southern Africa has yet to be undertaken. The Early to Mid-Holocene Later Stone Age crania from South Africa appear to share a common morphological pattern similar to Khoesan groups, suggesting a level of population continuity throughout the Holocene (Stynder, 2006, 2009; Stynder et al., 2007a, 2007b). However, the question of population continuity and/or discontinuity between the Late Pleistocene and Holocene still requires further exploration (Howells, 1995; Rightmire, 2009; Rightmire & Deacon, 1991). In other regions, such as in West (Central) Africa, the terminal Pleistocene and Early Holocene periods also reflect unique population diversification and/or isolation. For example, the Early Late Stone Age human remains from Ishango, which are dated to the Last Glacial Maximum (25,000–20,000 BP) provide a glimpse of Late Pleistocene diversity in Central Africa that seems to have been lost in modern populations (Crevecoeur et al., 2016). The cranial and mandibular morphology (e.g., large bimental breadth and a long corpus) together with that of the bony labyrinth morphology of the Ishango sample show plesiomorphic features that are shared with other Late Pleistocene African fossils. In addition, the Iwo Eleru neurocranium from the terminal Pleistocene of Nigeria appears to present some persistent archaic morphology (e.g., low frontal subtense, large cranial

I. Ribot et al.

and bifrontal breadths) (Harvati et al., 2011; Stojanowski, 2014; Stringer, 1974a, 1974b). It has been suggested that the paleoenvironmental setting (a humid, forested habitat) might have favored paleoecological isolation and thus persistence of both archaic human groups and technologies (Middle Stone Age industries) (Allsworth-Jones et al., 2010; Brothwell & Shaw, 1971). The Early to Mid-Holocene human remains from Shum Laka, Cameroon, also show some plesiomorphic features (e.g., a wide mandibular ramus and large teeth) (Lipson et al., 2020; Ribot, 2011; Ribot et al., 2001). This might suggest that terminal Pleistocene and Early Holocene West African foragers retained a phenotype that expressed some archaic population diversity. Selective pressures related to a coarse diet in relation to a mixed forest-savanna habitat might have affected both mandibular and dental morphologies (Hanihara & Ishida, 2005; von Cramon-Taubadel, 2011a; Romero et al., 2018). Previous morphometrical analyses of the Hofmeyr specimen (Crevecoeur et al., 2009; Grine et al., 2007) examined its possible phylogenetic links based on the whole cranium, using landmark coordinates together with linear measurements. These studies, which focused on the placement of Hofmeyr within the context of worldwide craniometric variation, found similarities with Upper Paleolithic Eurasians, as well as differences from most recent humans. However, these studies did not analyze the different anatomical regions of the skull separately, which can track different evolutionary signatures related to population history, climate and/or diet (Harvati & Weaver, 2006; Smith et al., 2007; von Cramon-Taubadel, 2011b). The objective of the present chapter is to re-evaluate in detail the position of the Hofmeyr skull within Late Pleistocene and Holocene African population diversity. Unlike previous studies, which did not focus on regional diversity (Crevecoeur et al., 2009; Grine et al., 2007), the present research employs a substantially larger data set for African populations that represent different regions and chronological groups. This will enable exploration of regional diversification and the possible phylogenetic links of Hofmeyr skull in relation not only to southern African, but also to various other areas, especially West and Central Africa (Crevecoeur et al., 2016; Harvati et al., 2011; Ribot, 2011; Ribot et al., 2001; Stojanowski, 2014). To address the issue of African regional cranial diversification, two complementary approaches will be utilized. These are presented separately as two case studies. Case Study 1 uses 2D linear morphometric analyses different regions of the skull (neurocranium, face, mandible), whereas Case Study 2 entails a 3D geometric morphometric analysis of the temporal bone, whose shape is considered as a reliable tool to track neutral genetic distances (Harvati & Weaver, 2006). Although the two studies are based on slightly different data sets in that Case Study 2 focuses on a smaller

7

Cranial 2D and 3D Morphometric Analyses

sample than Case Study 1, both add a significant number of West and Central African individuals (Fig. 7.1) into the analyses compared with previous studies. The results of these case studies will be compared with one another and with the results of previous investigations (Crevecoeur et al., 2009; Grine et al., 2007).

121

Case Study One Although linear craniometric analyses do not necessarily present some of the detail and power that can be achieved through 3D morphometric studies in the exploration of biological affinities (Hennessy & Stringer, 2002; Pan et al.,

Fig. 7.1 Map showing the skeletal sample used for the two case studies presented in this chapter. The definition of the codes of the comparative groups are detailed in Table 7.1. Symbols without a black outline correspond to Case Study 1; symbols with a black outline correspond to Case Study 2

122

2014), 2D studies have the advantage of including very large comparative samples and therefore a broader picture (Howells, 1989, 1995; Relethford, 1998, 2001; Pilloud & Hefner, 2016). Different anatomical regions of the skull may provide different signatures in relation to population history, climate and/or diet, such as the nasal region in relation to thermoregulation (Beals, 1972; Carey & Steegmann, 1981; Franciscus & Long, 1991; Harvati & Weaver, 2006; Hiernaux & Froment, 1976; Lieberman et al., 2004; Smith, 2009; Smith et al., 2007; Terhune et al., 2007; von Cramon-Taubadel, 2009a, 2011b, 2014). Other aspects of the cranium, such as the shapes of the basicranium, temporal, sphenoid, frontal and parietal bones, are less influenced by external factors and therefore may be more reliable when studying evolutionary relationships (Harvati & Weaver, 2006; Smith, 2009; Smith et al., 2007; Terhune et al., 2007; von Cramon-Taubadel, 2009a, 2009b, 2014). In view of the fact that different aspects of the skull may track different evolutionary signatures, the objective of Case Study 1 is to re-evaluate the position of the Hofmeyr skull in relation to the question of whether its different components provide different results in terms of affinities with various regions in response to different environments/selective pressures within Africa. The aim is to assess variable phenotypic plasticity by examining the data from separate regions of Africa and of the skull. Individual cranial “modules” are affected by both neutral and non-neutral forces (von Cramon-Taubadel, 2014), and those that are associated with a single function (e.g., orbital, nasal and auditory modules) tend to be less genetically congruent than multi-functional regions such as the vault, the face and the basicranium. Although some morphological features of the Hofmeyr skull appear to be distinct from patterns that are evident within present-day African variation (Grine et al., 2007), they may have similarities with past populations dated to the terminal Pleistocene and Mid-Holocene from elsewhere in Africa (Crevecoeur et al., 2009). The present study expands upon previous investigations by including Late Pleistocene and Holocene samples from West-Central Africa (e.g., Iwo Eleru, Shum Laka). This region might have comprised a refugium for groups with plesiomorphic features that disappeared in southern African populations.

Materials and Methods The linear morphometric analysis of the Hofmeyr skull entails examination of 17 variables recorded for a sample of 801 adult skulls from across the continent of Africa (Fig. 7.1). This data set is three times larger than that employed by Grine et al. (2007) and is entirely different from the one used by Crevecoeur et al. (2009). Significantly,

I. Ribot et al.

in comparison to these previous studies, the Later Stone Age sample from South Africa has also been greatly expanded. Comparative samples. The comparative human sample includes both recent individuals (N = 605) and those derived from archaeological contexts (N = 196) (Table 7.1). Seventy-two percent of the crania (N = 577) were directly measured by the first author (I. Ribot) and F. E. Grine (pers. comm.) in various museums as detailed in previous papers (Grine et al., 2007; Ribot, 2003, 2004, 2011; Ribot et al., 2006). The remaining 28% of the dataset (N = 224) was compiled from previous studies by Morris (1992a, 1992b). The recent African sample (N = 581) was subdivided into five geographical regions. Namely, 61 South Africans (SA) (Sotho-Tswana: Morris, 1992a), 84 East Africans (EA) (Bahutu from Rwanda: Ribot, 2003; various groups from Tanzania: Grine et al., 2007), 155 Central Africans (CA) (Basuku and various groups from Democratic Republic of Congo: Ribot, 2003, 2004, 2011), 226 West Africans (WA) (sampling eight countries: Grine et al., 2007; Ribot, 2003, 2004, 2011) and 55 North Africans (NA) (sampling three countries: Ribot et al., 2006). A small sample of present-day Europeans (EUR) (N = 24) (Grine et al., 2007) were employed as an “outgroup” for comparison. The comparative archaeological materials represent different periods and regions, with 136 Early to Later Holocene South Africans (ELHSA), 42 Late Pleistocene North Africans (LPNA), one Late Pleistocene West African (LPWA), one Late Pleistocene South African (LPSA), 15 Late Pleistocene Eurasians (LPE), and one Middle Pleistocene South African (MPSA) (Table 7.1). More precisely, the ELHSA group was represented by 45 San (RR) dated to the 17th– 19th c. A.D the from Riet River, South Africa (Morris, 1992a); 91 Early to Late Holocene Khoesan associated with radiocarbon sites (6 EH: Early Holocene; 45 MH: Middle Holocene; and 41 LH: Late Holocene) (Stynder, 2006, 2009; Stynder et al., 2007a, 2007b). The Fish Hoek 4 cranium, which was previously considered to be Late Pleistocene in age has been recently redated to 7,346–7,145 cal BP (Stynder, 2009). The Late Pleistocene North African (LPNA) group was represented by two Iberomaurusian sites (AFAL: Afalou-bou-Rhummel and TAF: Taforalt (Chamla, 1978; Hachi, 1996; Mariotti et al., 2009), as well as by Wadi Kubbaniya (WK) (Stewart et al., 1986, Wendorf et al., 1986) and Nazlet Khater (NK2) (Crevecoeur, 2008). Unfortunately, the LPWA group is represented by only one individual from Iwo Eleru (IE) (Brothwell & Shaw, 1971; Ribot, 2011). Variables employed. Seventeen linear dimensions (eight relate to both vault and base, eight are facial variables and one is mandibular) were used in the Case Study 1 analyses. These are listed in Table 7.2 together with their descriptions and abbreviations. The reconstructed Hofmeyr cranium

7

Cranial 2D and 3D Morphometric Analyses

123

Table 7.1 List of the comparative materials used for linear craniometrical analysis of the skull Group/Region (code used)

Country/Site (group or fossil code)

N

Additional information

Sources for measurements/dates

Modern groups

West Africa (WA)

226

Ribot (2003, 2004, 2011)

Central Africa (CA) East Africa (EA) South Africa (SA) North Africa (NA) Europe (EUR) Riet River (RR) Last 2,000 BP Late Holocene (LH) 5,000–2,000 BP Mid-Holocene (MH) 10,000–5,000 BP Early Holocene (EH) Afalou-bou-Rhummel (AFAL) Taforalt (TAF)

155 84 61 55 24 45 41

Mali (24), Senegal (38), Benin (20), Ghana (60), Liberia (3), Togo (10), Nigeria (38), Cameroon (33) D.R.C. (120 Basuku, 30 various) Rwanda (59 Bahutu), Tanzania (25 various) R.S.A (Sotho-Tswana) Morocco, Algeria, Tunisia Germany, Czech Republic Northern Cape Province (17th–19th c. A.D.) Various sites**

Wadi Kubbaniya (WK)

1

Egypt (19,000–17,000 BP)

Nazlet Khater (NK2) Iwo Eleru* (IE)

1 1

Egypt (*38,600 BP) Nigeria (*11,700–16,300 BP)

Chancelade* Cro-Magnon (1*, 2*) Grotte des Enfants* Dolni Vestonice Mladec 1 Predmosti (3*, 4*, 9, 10) Oberkassel Ohalo (II h2) Florisbad 1

1 2 1 3 1 2 4 1 1

France

Early to Late Holocene South Africa (ELHSA)

Late Pleistocene North Africa (LPNA)

Late Pleistocene West Africa (LPWA) Late Pleistocene Europeans (Middle East) (LPE)

45 6

Various sites**

27

Knysna Heads, Oakhurst, Plettenberg Bay, Matjes River Algeria (13,000–11,000 BP)

13

Morocco (12,000–11,000 BP)

Ribot et al. (2006) Grine et al. (2007) Morris (1992a, 1992b) Grine pers. comm. Stynder (2006), Stynder et al. (2007a, 2007b)

Chamla (1978), Hachi (1996) Chamla (1978), Mariotti et al. (2009) Stewart et al. (1986), Wendorf et al. (1986) Crevecoeur (2008) Brothwell and Shaw (1971), Ribot (2011), Harvati et al. (2011) Grine et al. (2007), Grine pers. comm

Czech Republic

Germany Israel South Africa (*259,000 BP)

Middle Pleistocene Grün et al. (1996), Grine South African pers. comm (MPSA) * Cast measurements by the authors ** List of the sites in alphabetical order: Blaauberg Strand, Bokbaai (Darling), Bredasdorp (Cape Aghulas), Buffels River, Clanwilliam, Die Dam, Elands Bay Cave, Fish Hoek 4, Gordon’s Groenriver, Groot Brak River Cave, Henkries, Hermanus, Hout Bay, Humansdorp District, Kabeljaauw’s Cave, Klein Melkbos, Kommetjie, Kruidfontein (Prince Albert), Ladysmith (Cape Province), Langebaar Lagoon, Matjes River, Melkbosch (Cape), Melkbosstrand, Mossel Bay (Gouritz River, Little Brak River), Mud River (Darling), Paardefontein (Jansenville), Paradysstrand (Jeffrey’s Bay), Plettenber Bay, Oudefontein (Koffiefontein). Robberg, Simonstown, Saldanha, Sand River and Thysbaai (Cape St. Francis), Seal Point, S. Lynch Point, Spitzkop Farm (Springvale), Tikoslip (Saldanha), Touws River (George), “Travalai” Kareespruit Farm, Vrendal, Witklip Farm (Vredenberg), Ysterfontein

(Grine et al., 2010), was used in this analysis as it offers a more complete and accurate morphometric picture. This entailed corrections to five variables from those values that were employed by the previous studies (Crevecoeur et al., 2009; Grine et al., 2007). These variables include WFB (99 mm instead of 98 mm), DKB (26 mm instead of 28 mm), OBH (31 mm instead of 32 mm), NLB (29 mm instead of 31 mm), and GLS (6 mm instead of 5 mm).

Nine variables related to various parts of the skull were employed (FRS, FRF, MDH, GLS, DKB, OBB, NLB, MAB; war) in comparison to previous linear morphometric study by Grine et al. (2007) that used only eight variables (GOL, XCB, FRC, WFB, ZYB, OBH, NLH, NPH). Six variables related to the frontal (FRS, FRF, MDH, GLS, DKB, MAB) were added in comparison to the study of Crevecoeur et al. (2009). The latter has 11 (out of 22)

124

I. Ribot et al.

Table 7.2 The seventeen measurements of the skull employed in Case Study 1 Anatomical region

Code used

Anatomical description

Maximal vault length (glabella – opisthocranion) XCB Maximal vault breadth (euryon – euryon) FRC Frontal chord FRS Frontal subtense (nasion – bregma) FRF Nasion subtense fraction WFB Minimal frontal breadth (between frontotemporalia) MDH Mastoid height perpendicular to Frankfurt plane GLS Glabellar projection Face ZYB Bizygomatic breadth (zygion – zygion) DKB Interorbital breadth OBB Orbital breadth (dacryon – ectoconchion) OBH Orbital height (perpendicular to orbital breadth) NLH Nasal height (nasion – nariale) NLB Nasal breadth (alare-alare) NPH Upper facial height (nasion – prosthion) MAB Maxillo-alveolar breadth (at level of M2) Mandible war Mandibular minimal rameal width * Variable used in the multivariate analysis by Grine et al. (2007) Vault and base

GOL

Martin and Saller (1959)

Howells (1989)

Crevecoeur et al. (2009)

Grine et al. (2007, 2010)

1

GOL

GOL

mcl*

8

XCB

XCB

mcb*

29 29b – 9

FRC FRS FRF WFB

FRC – – BFT

frc* nbs nsf mfb*

19a

MDH



mdh

45

GLS ZYB

– ZYB

gp bzb*

49a 51a

DKB OBB

– MEB

iob orb

52

OBH

OBH

orh*

55 54 48

NLH NLB NPH

NLH NLB NPH

nh* nb ufh*

61

MAB



mab

71a



RMB

Mrw

variables in common with the present study (GOL, XCB, FRC, WFB, ZYB, OBB, OBH, NLH, NLB, NPH and war). The measurements that were chosen for use here not only maximize sample sizes, but also serve to provide greater coverage of the frontal region, which often appears as a key anatomical area when studying population variation through time (Crevecoeur et al., 2009; Rightmire, 2009; Stynder et al., 2007a, 2007b). Statistical analyses. Univariate and multivariate statistics were performed with SPSS (Version 25), and the graphs were created with SYSTAT (Version 10.2) and improved in CorelDraw X3 (Version 13). The statistical approach employed here is broadly similar to that used in the previous linear morphometrical study (Grine et al., 2007). All variables were checked for normality within each group or region. As males and females were pooled, in contrast to the previous study (Grine et al., 2007), all variables were size-adjusted and transformed into C-scores according to the method of Howells (1989, 1995). Principal Component Analyses (PCAs) were performed with varimax rotation, as this procedure minimizes the number of variables with high loadings (or correlation coefficients that are used to express the weight assigned to each factor). Three sets of variables were selected to represent different parts of the cranium

(from general to more detailed analysis focusing on the frontal) and maximize the African data set (archaeological individuals often did not possess a full set of measurements). Three PCAs were conducted. In the first PCA (PCA I), nine variables were entered focusing on both the vault and the face (GOL, XCB, WFB, ZYB, OBB, OBH, NLH, NLB, NPH), which allowed for a comparison of Hofmeyr with three archaeological groups (LPNA, N = 42; LPE, N = 13; ELHSA, N = 123) and six modern ones (WA, N = 225; CA, N = 155; EA, N = 80; SA, N = 61; NA, N = 55; EUR, N = 23). In the second PCA (PCA II), eight variables were entered focusing predominantly on the vault despite a mixture of anatomical regions as it included one mandibular variable (GOL, XCB, FRS, FRF, WFB, MDH, GLS and war). This allowed for a comparison of Hofmeyr with three ancient groups (LPE N = 4, LPWA N = 1, ELHSA N = 41) and four modern groups (WA, N = 50; CA, N = 24; EA, N = 17; SA, N = 61). In the third PCA (PCA III), seven variables focusing predominantly on the frontal were entered (FRC, FRS, FRF, WFB, DKB, OBB, MAB). This enabled a comparison of Hofmeyr with four ancient (MPSA, N = 1; LPNA, N = 1; LPE, N = 8, ELHSA, N = 39) and four modern African

7

Cranial 2D and 3D Morphometric Analyses

groups (WA, N = 218; CA, N = 154; EA, N = 84; SA, N = 61). To assess the degree of differentiation of the comparative groups (N  15), one-way analyses of variance (ANOVA testing the equality of means) were performed on the regression scores of the principal components. Scatter plots of the principal components that showed significant interregional differences were produced to visualize the distribution of the entire sample in multivariate space. Comparative groups (N  15) were represented by an ellipse of variation (level of confidence: p = 0.95). In addition, to evaluate the closeness of Hofmeyr to the selected comparative groups or individuals, the regression scores obtained from the components were also used to compute proximity matrices of squared Euclidean distances and produce dendrograms (hierarchical clustering).

Results of Case Study One Descriptive statistics (sample size, mean and standard deviation) for the seventeen variables (transformed into Z-scores) for Hofmeyr and the comparative groups and individuals are presented in the Appendix (Table 7.A1). The three multivariate Principal Component Analyses are presented separately below. The components that presented eigenvalues (or total variance explained by each factor) greater than one were retained for the varimax rotation. For the three PCAs, the total cumulative percentage of variance varied between 73 and 85% and four components were obtained for each analysis. Principal Component Analysis I (PCA I). For PCA I, the variables with the highest component loadings (  0.7) on component 1 were CNLH, CNPH and CWFB, on component 2, COBH, on component 3, CNLB, and on component 4, COBB (Fig. 7.2a). One-way analyses of variance were all significant at the highest level of significance (P < 0.001) and inter-group variation was highest for components 2 and 3 (as they both had the highest F-values). The post-hoc multiple comparisons showed that the differences were spread out between all ten groups, and especially between five groups (CA, LH, MH, AFAL or LPE) and most of the others. However, for the scatter plots, some groups were pooled together into larger groups (e.g. EH, MH and LH grouped into ELHSA; TAF and AFAL grouped into LPNA), as they did not differ significantly from one another. Along component 1, Hofmeyr is positioned in the right half of the graph within the highest values of the modern variation and like some LPE fossils who tend to have relatively high face and nose. Along component 2, it is located in the upper right quadrant of the graph within the variation of both LPNA and LPE groups, who all exhibit lower orbits than modern groups (Fig. 7.2a). A clinal variation related to

125

nasal aperture is broadly visible along component 3: Hofmeyr falls even better within the modern and Late Pleistocene variation especially from Africa (Fig. 7.2b). Along component 4, Hofmeyr is located on the edge of most of the variation and is more distant to other Pleistocene Africans, who tend to have narrower orbits (Fig. 7.2c). According to the proximity matrix (Table 7.3), Hofmeyr is closest (in decreasing order) to LPE, RR, EUR, AFAL, SA and most distant to Wadi Kubbaniya (WK). The latter appeared very distant to all groups, except for TAF. The dendrogram showed clearly that Hofmeyr, like a few other fossils (WK, FH 4), did not cluster with all other groups (Fig. 7.3). However, the modern Africans clustered well together, as did the various Holocene South Africans (RR, LH, MH, EH, except for Fish Hoek 4) and as did the various Late Pleistocene Africans (AFAL, TAF) and Europeans (LPE). Principal Component Analysis II (PCA II). For PCA II, the variable with the highest component loadings (  0.7) on component 1 was CWFB, on component 2, CFRS, on component 3, CMDH, and on component 4, Cwar (Fig. 7.4). One-way analyses of variance were often significant (component 2: P < 0.01: component 3: P < 0.001; component 4: P < 0.01). Inter-group variation was highest for component 3 (as the latter has the highest F-value). The post-hoc multiple comparisons showed that differences for component 2 were between WA and SA, for component 3 among mainly WA, EA and RR and between the latter and CA, and for component 4 between WA and SA and between the latter and RR. In all scatter plots, Hofmeyr falls within the ellipses of the modern variation, although it is slightly distant from the centroid. It is located along component 1 in the right half of the graph, as it has a relatively large minimal frontal breadth like the LPE and LPWA fossils. However, it is positioned in the lower right quadrant for the two other components (IE tends to be at the opposite end), as it presents a very small frontal subtense (component 2) (Fig. 7.4a) and high mastoid (component 3) (Fig. 7.4b). Hofmeyr also clusters well along component 4 with some of the LPE and LPWA individuals, who tend to have rather wide mandibular ramus (Fig. 7.4c). According to the proximity matrix (Table 7.4), Hofmeyr is closest (in decreasing order) to WA, CA, SA, RR and LPE and most distant to one LPWA individual (IE). The latter is very distant to all groups, although slightly closer to the LPE group and Hofmeyr, as illustrated by the dendrogram that portrays group clustering (Fig. 7.5). Principal Component Analysis III (PCA III). For PCA III, the variables with the highest component loadings (  0.7) on component 1 were CDKB and CFRS, on component 2, COBB, on component 3 CFRF, and on component 4 CMAB (Fig. 7.6). All one-way analyses of variance were significant (at least P < 0.01) and, according to the F-values, the inter-group variation was highest in decreasing order

126

I. Ribot et al.

Fig. 7.2 Scatter plots of the four principal components of PCA I (total cumulative variance: 73%). Comparative groups (N  15) are represented with 95% confidence ellipses and their centroid. a Components 1 and 2; b Components 1 and 3; c Components 1 and 4

Table 7.3 Proximity matrix of squared Euclidean distances (with highlighted lowest value for each group or individual) derived from the regression scores of the four components of PCA I Groups or Individuals (code used) HOFMEYR West Africa (WA) Central Africa (CA) East Africa (EA) South Africa (SA) North Africa (NA) Europe (EUR) Riet River (RR) Last 2,000 BP (LH) 5,000-2,000 BP (MH) 10,000-5,000 BP (EH) Fish Hoek (FH4) Afalou-bou-R. (AFAL) Taforalt (TAF) Wadi Kubbaniya (WK) Nazlet Khater (NK2) Late Pleistocene Europe (LPE)

HOFMEYR 0.000 13.480 14.466 13.020 9.941 12.042 9.230 7.796 14.776 14.754 21.462 16.027 8.371 10.842 34.481 11.110 6.512

WA

CA

EA

SA

NA

EUR

RR

LH

MH

EH

FH4

AFAL

TAF

WK

NK2

LPE

0.000 0.225 0.433 0.440 2.300 6.530 1.435 2.407 3.621 7.772 21.543 4.019 3.617 14.236 3.649 7.711

0.000 0.524 0.428 2.880 6.686 1.584 3.144 3.768 8.535 19.742 5.280 5.195 17.134 4.977 9.588

0.000 0.599 0.957 3.935 1.083 2.058 2.378 7.369 18.425 3.949 3.858 15.410 3.373 6.920

0.000 2.376 5.072 0.526 2.903 3.398 8.345 16.788 3.579 3.950 17.761 4.101 6.849

0.000 1.811 2.001 2.426 2.464 7.908 18.667 3.731 3.616 14.928 2.427 4.748

0.000 3.731 6.635 4.875 12.975 13.101 6.725 7.929 25.300 5.076 6.133

0.000 2.008 1.981 6.280 12.892 1.925 2.778 16.976 4.891 4.322

0.000 0.753 1.846 18.261 1.417 1.247 8.288 6.803 4.194

0.000 2.257 12.660 2.576 3.228 12.973 8.775 5.281

0.000 18.809 3.629 3.781 7.687 15.466 7.216

0.000 17.022 22.019 47.961 29.646 18.911

0.000 0.377 10.158 6.501 1.212

0.000 7.158 5.486 1.715

0.000 17.234 12.699

0.000 6.521

0.000

7

Cranial 2D and 3D Morphometric Analyses

Fig. 7.3 Dendrogram using the mean distances from the proximity matrix (PCA I) for Hofmeyr and sixteen comparative groups or individuals

127

along components 1, 3, 4 and 2. The post-hoc multiple comparisons indicated that differences for component 1 were between SA and all groups except for RR, and between CA and all groups except for EA and EUR. For component 2 the differences were between WA and CA, for component 3 between SA and EA and CA, and for component 4 between WA and SA and between the latter and RR. In all scatter plots, along component 1, Hofmeyr is located in the right half of the graph (as it has a relatively wide interorbital area and long frontal subtense) and it is at the opposite end of one MPSA individual (Florisbad 1) with one LPNA fossil (NK2) in between. Along component 2, Hofmeyr falls within the modern variation, but remains relatively isolated from the other fossils that have large orbits (Fig. 7.6a). This trend is accentuated along the component 3 and 4, as Hofmeyr has a rather high frontal fraction (Fig. 7.6b) and broad maxillae (Fig. 7.6c). According to the proximity matrix (Table 7.5), Hofmeyr, as well as Florisbad 1 are rather marginal, as they appear very distant to all

Fig. 7.4 Scatter plots of the four principal components of PCA II (total cumulative variance: 73%). Comparative groups (N  15) are represented with 95% confidence ellipses and their centroid. a Components 1 and 2; b Components 1 and 3; c Components 1 and 4

128

I. Ribot et al.

Table 7.4 Proximity matrix of squared Euclidean distances (with highlighted lowest value for each group or individual) derived from the regression scores of the four components of PCA II Groups or Individuals (code used) HOFMEYR West Africa (WA) Central Africa (CA) East Africa (EA) South Africa (SA) Riet River (RR) Iwo Eleru (IE) Late Pleistocene Europe (LPE)

HOFMEYR 0.000 1.817 3.449 1.404 3.650 5.205 25.379 5.398

Fig. 7.5 Dendrogram using the mean distances from the proximity matrix (PCA II) for Hofmeyr and seven comparative groups or individuals

groups, although slightly closer to one modern African group (SA) and LPE group respectively. The dendrogram also shows that the modern groups tend to cluster together in contrast to the Late (and Middle) Pleistocene Africans (Florisbad 1, Hofmeyr, Nazlet Khater 2) and one South African LSA individual (Fish Hoek 4) (Fig. 7.7).

Discussion of Case Study One In sum, Hofmeyr appears closest (in decreasing order), not only to Late Pleistocene Europeans but also to various archaeological groups (Riet River and Afalou-bou-Rhummel (PCA I: vault and face), but also to various modern Africans (PCA II and III: predominantly vault and frontal respectively). These results are discussed below, focusing on the affinities or differences of Hofmeyr with the various comparative groups or individuals. Hofmeyr compared to Florisbad – the Middle Pleistocene South African (MPSA). The PCA III that included Florisbad I allowed us to compare Hofmeyr with one MPSA

WA

CA

EA

SA

RR

IE

LPE

0.000 0.391 0.052 1.275 3.144 20.916 3.294

0.000 0.635 0.712 2.484 22.964 4.048

0.000 1.284 3.120 20.374 2.987

0.000 1.291 23.816 3.763

0.000 22.279 3.147

0.000 9.348

0.000

individual. Although it did not provide supplementary information, it underlined the overall morphological modern features of the Hofmeyr cranium, with robust features (e.g., strong supraorbital torus, projecting glabella, alveolar prognathism) (Grine et al., 2007, 2010). In addition, PCA III showed that in comparison to Florisbad, Hofmeyr presents a much higher frontal, narrower interorbital region, smaller orbital breadth and wider maxillo-alveolar breadth. These features suggest a morphological evolution leading to anatomically modern humans and were different from this MPSA fossil as is also supported by the highest squared Euclidean distance value (Table 7.5; Fig. 7.6). Hofmeyr compared to various Late Pleistocene groups (LPE, LPNA, LPWA). The multivariate analyses (especially PCA I) allowed Hofmeyr to be compared with the Late Pleistocene Europeans (LPE). Hofmeyr appears to be located consistently close or within the LPE variation, which tended to be characterized by a high face, a high nose and high orbits in PCA I; a large frontal, a high mastoid and a wide mandibular ramus in PCA II. Most of these observations were already noted in Grine et al. (2007), except for the last two variables. However, the PCA III shows that Hofmeyr is more different from the LPE group, as it has narrower orbits, narrower maxillae and a lower frontal fraction, which is also supported by the high squared Euclidean distance value (Table 7.5; Fig. 7.6). Similar observations related to its robusticity were made earlier, showing not only morphological similarities but also differences between Hofmeyr and LPE (Crevecoeur et al., 2009), as well as within the latter (Balzeau & Badawi-Fayad, 2005). As expected, Hofmeyr is closer to the earliest Late Pleistocene North Africans (LPNA) individuals (Nazlet Khater and Wadi Kubbaniya) than the more recent LPNA individuals (the Iberomaurusians, Taforalt and Afalou-bou-Rhummel) (Figs. 7.2 and 7.3). Orbital height in Hofmeyr is similar to WK, in contrast to NK2, which is in the lower range. When Hofmeyr is compared with one Late Pleistocene West African (LPWA: Iwo Eleru), it appears rather close to the latter for various morphological features reflecting robusticity (e.g., mastoid height, width of mandibular ramus) (Figs. 7.4 and 7.5). Despite these similarities shared with

7

Cranial 2D and 3D Morphometric Analyses

129

Fig. 7.6 Scatter plots of the four principal components of PCA III (total cumulative variance: 85%). Comparative groups (N  15) are represented with 95% confidence ellipses and their centroid. a Components 1 and 2; b Components 1 and 3; c Components 1 and 4

Table 7.5 Proximity matrix of squared Euclidean distances (with highlighted lowest value for each group or individual) derived from the regression scores of the four components of PCA III Groups or Individuals (code used) HOFMEYR West Africa (WA) Central Africa (CA) East Africa (EA) South Africa (SA) Europe (EUR) Riet River (RR) Fish Hoek (FH4) Nazlet Khater (NK2) Late Pleistocene Europe (LPE) Florisbad 1 (FL1)

HOFMEYR 0.000 20.966 24.544 21.024 18.988 24.487 25.752 25.512 26.028 28.140 84.185

WA

CA

EA

SA

EUR

RR

FH4

NK2

LPE

FL1

0.000 0.666 0.515 1.087 0.694 1.858 7.967 1.621 1.502 29.892

0.000 1.308 1.288 0.325 1.110 6.028 1.634 1.597 33.734

0.000 1.635 0.567 2.608 5.815 3.789 1.412 27.882

0.000 1.271 0.587 6.322 3.437 1.259 35.595

0.000 1.267 4.571 3.035 0.949 30.481

0.000 6.344 3.524 0.876 34.191

0.000 13.667 6.706 46.958

0.000 4.441 35.099

0.000 24.914

0.000

Late Upper Pleistocene Europeans, it is still very different from Iwo Eleru, who presents a much lower and less curved frontal as often observed in the LPE group (Harvati et al., 2011; Stringer, 1974a, 1974b). However, as the Iwo Eleru

skull has been affected by both distortion and considerable reconstruction (Brothwell & Shaw, 1971) as well as by a questionable U-series dating (Pike & Hedges, 2002; Allsworth-Jones et al., 2010), this does not allow us to

130

I. Ribot et al.

In conclusion, Case Study 1 enabled Hofmeyr to be examined within the context of both past and modern African cranial variation. Although we failed to completely separate the different anatomical parts (face, base, vault, mandible) because of the desire to maximize sample size, the three PCAs provided a means to analyze its morphology through different combinations of variables. As previously observed (Crevecoeur et al., 2009; Grine et al., 2007), these craniometrical observations suggested that Hofmeyr is often relatively closer to: (i) most of the penecontemporaneous Late Pleistocene individuals (e.g., in its low orbits, tall mastoid, and wide ramus), and (ii), some recent samples from North, West and South Africa (e.g., in its narrow orbits, curved frontal, and large maxillae). Hofmeyr does not consistently reside at the margin of modern African variation. Fig. 7.7 Dendrogram using the mean distances from the proximity matrix (PCA III) for Hofmeyr and ten comparative groups or individuals

Case Study Two

propose very reliable interpretations regarding its persistent archaic nature in the terminal Pleistocene. Currently, signatures of archaic admixture in Africa have been detected through genetics rather than morphology (Hammer et al., 2011). This high range of variability in Late Pleistocene human crania reflects a variable retention of robusticity, as noted by Crevecoeur et al. (2009), who compared Hofmeyr with penecontemporaneous remains from Europe (Peştera cu Oase) and Egypt (Nazlet Khater). Interestingly, Hofmeyr is dated to a chronological period (Marine Isotope Stage 3) that corresponds to significant climatic changes. The latter might have encouraged some geographical isolation and therefore regional diversification (but also extinctions) of archaic populations in both Africa and Europe (Bräuer, 2008; Grine et al., 2010; Howell, 1999; Lahr, 2016). Hofmeyr compared to Early to Late Holocene South Africans (ELHSA). In general, Hofmeyr falls comfortably in ELHSA group variation for different features (e.g., nose, orbits, frontal convexity, mandibular ramus), although its morphology (e.g., maxillary breadth, mastoid height) remains distinct (Figs. 7.2 and 7.6). Nevertheless, it is difficult to support a hypothesis of discontinuity between Hofmeyr and the ELHSA group, as was previously thought (Grine et al., 2007; Stynder, 2009). The Hofmeyr cranium reflects a mosaic of features that are only shared by the ELHSA group (Morris, 2002, 2003; Morris & Ribot, 2006). In this instance, morphology does not help to assess if the Khoesan lineage arose earlier than the Holocene, in contrast to the very clear genetic evidence that it did (Henn et al., 2011; Schlebusch & Jacobson, 2018; Schlebusch et al., 2013).

Several studies have argued that some components of the skull, and especially the temporal bone, are more reliable than the facial skeleton and mandible for inferring phylogenetic relationships (Harvati & Weaver, 2006; von Cramon-Taubadel, 2009a, 2014). The hypothesis supporting the differential preservation of information among different cranial regions was first tested on the temporal bone to distinguish taxa (Lockwood et al., 2002, 2004, 2005). In this instance, the phylogenetic tree determined from the shape of the temporal bone of modern humans, orangutans, gorillas, chimpanzees and bonobos was identical to the molecular phylogeny for these species (Lockwood et al., 2004; Smith et al., 2007). In their comparative study of morphological and genetic distances among modern populations, Harvati and Weaver (2006) found that the morphology of the cranial vault and temporal bone correlated with molecular distances, a finding supported by Smith et al. (2007). Although the temporal bone has been argued to be reliable for studying past population history (Wood & Lieberman, 2001), the influence of environmental factors cannot be excluded. Thus, while Harvati and Weaver (2006) found a significant correlation between temporal bone shape and molecular distance, it remained low. Furthermore, temporal bone size correlates with temperature and latitude, and Smith et al. (2007) have suggested that factors affecting the shape of the temporomandibular joint could explain part of the variation. Given that the temporal bone is linked to masticatory function, its form might be affected by diet, and given its role in other functions (e.g., housing the bony labyrinth) there may be a number of extraneous influences and selective pressures affecting its form. Nevertheless, there are two

7

Cranial 2D and 3D Morphometric Analyses

principal reasons that have been argued in support of its use in phylogenetic studies, especially when molecular data are missing. The first is that is a key element because of its functional and anatomical complexity, and the second is that, because it comprises part of the basicranum, it is considered reliable for reflecting genetic relationship between taxa (Lockwood et al., 2004; von Cramon-Taubadel, 2009a, 2009b; Wood & Lieberman, 2001). Indeed, most studies support the notion that temporal bone shape is a reliable anatomical region by which to infer past population history and phylogenetic relationships (Harvati & Weaver, 2006; Smith, 2009; Smith et al., 2007; Terhune et al., 2007; von Cramon-Taubadel, 2014). Knowing these advantages, Case Study 2 focuses on the temporal bone with a more detailed morphometric approach (Mitteroecker & Gunz, 2009). For Hofmeyr and other sub-Saharan African fossils, this approach is very suitable, as specimens are often fragmentary. Although the Hofmeyr cranium has been badly affected by mishandling, the temporal bone is almost entirely intact. Nevertheless, this region of the fossil has never been studied morphometrically as a separate entity. The objective of Case Study 2 was to analyze and compare the temporal to a small but adequate sample of homologous elements to address the question whether the Hofmeyr element is as distinct as its general cranial morphology from modern African populations.

Material and Methods For the present geometric morphometric analysis of the Homeyr temporal bone, the 3D model of the cranium was obtained courtesy of Philipp Gunz, who was responsible for the virtual reconstruction presented by Grine et al. (2010). The comparative data set (N = 65) and the list of the

131

landmarks that were used are detailed in Tables 7.6 and 7.7 respectively (see also Fig. 7.1). Only left temporal bones were selected for this study to avoid variation related to asymmetry. Comparative samples. Temporal bones from recent African individuals (N = 57) and from a smaller number derived from archaeological contexts (N = 8) were employed in comparative contexts with Hofmeyr (Table 7.6). Only left temporal bones were selected for study to avoid potential variation related to bilateral asymmetry. With the exception of the Hofmeyr temporal bone, all of the comparative elements were three-dimensionally modelled by one of us (Y. Ghalem) by photogrammetry on skeletal collections at the Royal Institute of Natural Sciences of Belgium (IRScNB) (Ribot, 2011). The modern African sample was divided into two main geographical regions: East Africa (N = 20) and Central Africa (N = 37). The East Africans represent a homogeneous ethnic group, the Bahutu from Rwanda. The Central African sample was composed of four ethnic groups: the Basuku (N = 20), the Azande (N = 7), the Bassoko (N = 3) and the Mongo (N = 4). Pygmies s from the Irumu region in D.R.C. (N = 3) were also included in the recent sample. The archeological sample originated mainly from equatorial Africa and is dated to the Early Holocene (N = 8). It included specimens from four different sites from the Upemba Depression, D.R.C. – six individuals from Katongo, Kikulu, Malemba-Nkulu and Sanga, dated to some 1,300 BP (de Maret, 1977, 2016), and two individuals from Shum Laka, Cameroon, dated to about 3,000 BP (Lipson et al., 2020; Ribot et al., 2001). Analytical and statistical methods. The data consisted of 16 3D temporal bone landmarks (Table 7.7) as defined in the studies by Harvati and Weaver (2006) and Smith et al. (2007). The 3D models were produced with photogrammetry

Table 7.6 Comparative materials used for the geometric morphometric analysis of the temporal bone Population/Site (code of skeleton)

Country (region)

Bahutu Rwanda Basuku R.D.C Azande Bassoko Mongo Pygmies R.D.C. (Irumu) Katongo (T3) Katongo (KTG_T3) R.D.C. (Upemba Depression) Kikulu (KUL_T1, T7, T14) Malemba-Nkulu (MAK_T9) Sanga (SGA_T18) Shum Laka (6/SEIII) Cameroon (Grassfields) Shum Laka (6/SEIV) * Ribot (2011) ** de Maret (1977) *** Ribot et al. (2001), Lipson et al. (2020)

Period and date

Group name

Total

Modern 19th–20th century AD*

East Africa Central Africa

20 20 7 3 4 3 1 3 1 1 2

Iron Age circa 1,300 BP**

Stone to Metal Age 3,820–3250 BP***

Pygmies Upemba Depression

Shum Laka

132

I. Ribot et al.

Table 7.7 The sixteen anatomical landmarks used in the geometric morphometric analysis of the temporal bone Landmark number

Landmark name

Definitions from Harvati and Weaver (2006) and Smith et al. (2007)

1 2 3 4 5

Most anterior point on articular surface of auricular eminence Point at center of articular eminence Most inferior point on medial margin of articular surface of articular eminence Most inferior point on entoglenoid process (inferior one) Midpoint of lateral margin of articular surface of auricular eminence

6 7 8 9

Anterior auricular eminence Center of the articular eminence Inferior auricular eminence Inferior entoglenoid Midpoint of the lateral margin of the articular surface of the auricular eminence Lateral margin in the articular eminence Mandibular fossa Postglenoid process Stylomastoid foramen

10 11 12 13

Mastoidale Origin of tympanic crest Porion Auricular

14

Radicular

15

Parietal notch

16

Asterion

Deepest point of lateral margin of auricular eminence Deepest point within mandibular fossa Most inferior point on postglenoid process Point on stylomastoid foramen; between styloid process and mastoid process of temporal bone Most inferior point on tip of mastoid process Lateral origin of petro-tympanic crest Most superior point along upper margin of external acoustic meatus Most posterolateral point on tympanic element of temporal bone/on top of porion Point of lateral margin of zygomatic process of temporal bone in coronal place of postglenoid process; point on lateral aspect of root of zygomatic process at deepest incurvature Postero-superior border of temporal bone where squamous and parietomastoid sutures meet Point where temporal, parietal and occipital bones meet

and the landmark coordinates were recorded using Landmark Editor (Wiley, 2006). Because the Hofmeyr right temporal bone is not preserved wholly intact, it was necessary to estimate three missing 3D coordinates utilizing the reconstruction presented by Grine et al. (2010): the inferior auricular eminence (landmark n°3), the stylomastoid foramen (landmark n°9) and mastoidale (landmark n°10). For this additional step, a simple reflection operation of the three coordinates was done using R software (R Core Team, 2018). Shape analysis of landmark configurations were conducted using the MorphoJ program (Klingenberg, 2011). The raw landmark coordinates were first imported into MorphoJ and then processed with a Generalized Procrustes Superimposition (GPA). This step allows to separate shape from overall size, position, and orientation of landmark configurations and, as a result, Procrustes shape coordinates were obtained and used for statistical analysis (Mitteroecker & Gunz, 2009). Principal Component Analysis (PCA) was conducted to examine the relationships between Hofmeyr and the comparative samples. Shape differences were visualized using wireframe diagrams and scatter plots were made for the first two components (for the comparative group with a 95% confidence ellipse when N  6). Procrustes residuals were used to calculate a matrix of Procrustes distances to evaluate group or individual differences into a morphospace, and then, permutation tests were conducted to assess the

significance of shape differences (Franklin et al., 2010). Following Smith et al. (2007), the significance of distances between groups or individuals was assessed using a permutation test with 1000 replicates (P-values were calculated as the number of times the original Procrustes distance between populations was exceeded, divided by the total number of iterations).

Results of Case Study Two The shape variation of the temporal bone among modern and archaeological groups (with 95% confidence ellipses) and individuals can be visualized through the PCA scatter plot (Fig. 7.8). Axis 1 explains 15% of the sample variance (Table 7.8) and represents variation related the latero-posterior length of the temporal bone as well as to its articular eminence, mandibular fossa and mastoid region. Although the comparative groups slightly overlapped along Axis 1, they tended to cluster well together: while the East African sample tended to load positively (they have a long temporal bone in the latero-posterior region), the Central Africans loaded both positively and negatively (they are very variable having both long and short temporal bone in the latero-posterior region). Iron Age groups from the Upemba Depression clustered well together, slightly closer to Central Africans than East Africans. The two individuals from Shum Laka diverged highly along Axis 1: 6/SEIV is

7

Cranial 2D and 3D Morphometric Analyses

133

Fig. 7.8 Scatter plot of first two principal components obtained from the geometric morphometric analysis of the temporal bone (total cumulative variance: 28%). Confidence ellipses encompass 95% of shape variation Table 7.8 Eigenvalues and distribution of variance for the first 13 principal components obtained from the geometric morphometric analysis on the temporal bone Principal component

Eigenvalues

% Variance

Cumulative %

1 2 3 4 5 6 7 8 9 10 11 12 13

0.002043 0.001921 0.001426 0.001303 0.001094 0.000746 0.000703 0.000567 0.000507 0.000460 0.000379 0.000332 0.000297

14.745 13.865 10.289 9.407 7.896 5.384 5.073 4.091 3.660 3.317 2.737 2.398 2.142

14.745 28.610 38.899 48.306 56.202 61.585 66.658 70.749 74.409 77.726 80.463 82.861 85.003

located in the middle of the modern Central African variation; and 6/SEIII is located in the top right quadrant with the longest temporal bone. Like 6/SEIII, Hofmeyr did not cluster within the modern or the archeological samples but is located within the 95% confidence ellipse of modern Central Africans for the shortest latero-posterior length. Axis 2 explains 13% of the sample variance (Table 7.8) and represents shape variation related to the squama region

of the temporal bone (Fig. 7.8). Comparative groups overlapped substantially along Axis 2 and did not cluster well within the Central African sample. The main variation is related to the mediolateral shape of the bone. Individuals that loaded positively have a wide temporal bone, while individuals that loaded negatively have a narrower one. Additionally, shape variation on Axis 2 reflects changes related to both the squama and the mastoid regions. One individual

134

I. Ribot et al.

Table 7.9 Procrustes distances among groups and/or individuals obtained from the geometric morphometric analysis on the temporal bone. Highlighted values correspond to significant permutation tests (see next table) Central Africa Central Africa East Africa Pygmies Upemba Depression Shum Laka 6/SEIII Shum Laka 6/SEIV HOFMEYR

0.0000 0.0593 0.0939 0.0678 0.2392 0.1222 0.1958

East Africa 0.0000 0.0764 0.0807 0.2349 0.1346 0.2083

Pygmies

0.0000 0.1049 0.2448 0.1516 0.2257

Upemba Depression

0.0000 0.2309 0.1348 0.1956

Shum Laka 6/SEIII

Shum Laka 6/SEIV

0.0000 0.2315 0.3597

0.0000 0.2040

Table 7.10 P-values obtained from permutation tests for the Procrustes distances obtained from the geometric morphometric analysis of the temporal bone (highlighted P-values: