Genes, Language, and Culture History in the Southwest Pacific
HUMAN EVOLUTION SERIES SERIES EDITORS Russell L. Ciochon, The University of Iowa Bernard A. Wood, George Washington University EDITORIAL ADVISORY BOARD Leslie Aiello, University College, London Alison Brooks, George Washington University Fred Grine, State University of New York, Stony Brook Andrew Hill, Yale University David Pilbeam, Harvard University Yoel Rak, Tel-Aviv University Mary Ellen Ruvolo, Harvard University Henry Schwarcz, McMaster University African Biogeography, Climate Change, and Human Evolution Edited by Timothy G. Bromage and Friedemann Schrenk Meat-Eating and Human Evolution Edited by Craig B. Stanford and Henry T. Bunn The Skull of Australopithecus afarensis William H. Kimbel, Yoel Rak, and Donald C. Johanson Early Modern Human Evolution in Central Europe: The People of Dolni Vˇestonice and Pavlov Edited by Erik Trinkaus and Jiˇrí Svoboda Evolution of the Human Diet: The Known, the Unknown, and the Unknowable Edited by Peter S. Ungar Genes, Language, and Culture History in the Southwest Pacific Edited by Jonathan Scott Friedlaender
Genes, Language, and Culture History in the Southwest Pacific
EDITED BY
Jonathan Scott Friedlaender
1 2007
1
Oxford University Press, Inc., publishes works that further Oxford University’s objective of excellence in research, scholarship, and education. Oxford New York Auckland Cape Town Dar es Salaam Hong Kong Karachi Kuala Lumpur Madrid Melbourne Mexico City Nairobi New Delhi Shanghai Taipei Toronto With offices in Argentina Austria Brazil Chile Czech Republic France Greece Guatemala Hungary Italy Japan Poland Portugal Singapore South Korea Switzerland Thailand Turkey Ukraine Vietnam
Copyright © 2007 by Oxford University Press, Inc. Published by Oxford University Press, Inc. 198 Madison Avenue, New York, New York 10016 www.oup.com Oxford is a registered trademark of Oxford University Press All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise, without the prior permission of Oxford University Press. Library of Congress Cataloging-in-Publication Data Genes, language, and culture history in the Southwest Pacific / edited by Jonathan Scott Friedlaender. p. cm. — (Human evolution series) ISBN 978-0-19-530030-7 1. Oceania—Civilization—Congresses. 2. Population genetics— Oceania—Congresses. 3. Linguistics—Oceania—Congresses. 4. Human population genetics—Oceania—Congresses. I. Friedlaender, Jonathan Scott. II. Series. DU28.P66 2007 995—dc22 2006014236
9 8 7 6 5 4 3 2 1 Printed in the United States of America on acid-free paper
Preface
The immediate stimulus for this book was a symposium entitled “A Second Garden of Eden: Island Melanesian Genetic Diversity” that was held at the American Association of Physical Anthropology in Tampa, Florida, April, 15, 2004. Its focus was the genetic diversity in the area just to the east of New Guinea referred to as Northern Island Melanesia. Our intensive field studies there in 1998, 2000, and 2002 in more than 25 populations were yielding remarkable patterns of genetic variability that were the focus of the related studies reported at the symposium. The presentations at the symposium consisted of 12 papers authored by 21 anthropological geneticists, linguists, and archeologists. After the symposium, I asked the authors to revise their papers for incorporation into an edited volume, and additional authors were asked to add complementary chapters in particular areas that would supplement and extend the coverage of this research. These included both anthropological geneticists and linguists. All chapters were not only reviewed and edited by me, but in most instances, and particularly where my expertise was less than desirable, anonymous outside reviewers were solicited. The result was accepted for publication by Oxford University Press (New York), which has overseen a most prompt publication process. While a number of chapters have special Acknowledgments sections, the entire work could not have been possible without the remarkable cooperation and interest of hundreds, or actually thousands, of people from Papua New Guinea, Solomon Islands, and neighboring regions of the Pacific. Over the years, we have been extremely fortunate the have their good will and trust, and we hope that the publication of this book will, to some degree, fulfill our commitment to them to make some new sense of their own history.
Our collaborative arrangement with the Papua New Guinea Institute for Medical Research was essential in carrying out this work. Although he is acknowledged as a co-author in a number of the papers involving the analysis of collected blood samples, as a group we are indebted one way or another to George Koki, Principal Technical Officer in Genetics and Molecular Microbiology at the IMR. George was not only the phlebotomist on the expeditions of the last decade, but was effectively co-leader of those field expeditions. He was particularly important in gaining the cooperation of participating groups and was essential in educating them as to the purpose of our endeavor. While some people decided they did not wish to participate, most did. The essential point is that George made sure they were properly informed as to what we were attempting to accomplish. At the Institute, the Directors and Assistant Directors, Michael Alpers, Charles Mgone, and more recently John Reeder, as well as Martina Yambun, have devoted critical support to the project. In Solomon Islands, Lawrence Foanaota, Director and Curator of the National Museum, played a similar role. Daniel Hrdy was a most effective volunteer participant in the 1998 field expedition, and deserves special thanks. Roger Green has been a constant source of encouragement and advice throughout this process, for which I am particularly grateful. Benjamin Friedlaender made a significant contribution to database construction. The photo of the Tanenuiarangi meeting house carving (figure 10.2) was taken by Tim Mackrell, Auckland University. The color portraits were taken by Eva Lindström (for New Ireland), Ger Reesink (for New
Preface
Britain), Lot Page (for Malaita), and me (for Bougainville and of George Koki and Heather Norton). As a matter of intellectual descent, this book has three grandfathers: the late William W. Howells and Albert Damon, who led the Harvard Solomon Islands Expeditions in the 1960s and 1970s; and particularly Douglas Oliver, pioneer in Pacific anthropology, who persuaded a number of graduate students in cultural anthropology, archeology, and biological anthropology to develop thesis projects in Island Melanesia and elsewhere in the Pacific. My participation in the first of those expeditions began a career-long fascination with understanding Island Melanesian diversity.
vi
Finally, we acknowledge the grants and material support provided by the National Science Foundation (USA), the Wenner-Gren Foundation for Anthropological Research, The National Geographic Society Exploration Fund, Temple University, The University of Michigan, Binghamton University, The National Institutes of Health, The Coriell Institute for Medical Research, the University of Pennsylvania, and The Pennsylvania State University.
Jonathan S. Friedlaender, March 10, 2006
Contents
Contributors
ix
Part I The Framework 1 Introduction
3
Jonathan S. Friedlaender
2 Island Melanesian Pasts: A View from Archeology
10
Glenn R. Summerhayes
3 Recent Research on the Historical Relationships of the Papuan Languages, or, What Does Linguistics Say about the Prehistory of Melanesia? 36 Andrew Pawley
Part II Core Studies in Northern Island Melanesia 4 Mitochondrial DNA Variation in Northern Island Melanesia 61 Jonathan S. Friedlaender, Françoise R. Friedlaender, Jason A. Hodgson, Stacy McGrath, Matthew Stoltz, George Koki, Theodore G. Schurr, D. Andrew Merriwether
5 Y Chromosome Variation in Northern Island Melanesia
81
Laura B. Scheinfeldt, Françoise R. Friedlaender, Jonathan S. Friedlaender, Krista Latham, George Koki, Tatiana Karafet, Michael Hammer, Joseph Lorenz
6 Pigmentation and Candidate Gene Variation in Northern Island Melanesia 96 Heather L. Norton, George Koki, Jonathan S. Friedlaender
Contents
7 The Distribution of an Insertion/Deletion Polymorphism on Chromosome 22 113 Renato Robledo
8 The Languages of Island Melanesia
118
Eva Lindström, Angela Terrill, Ger Reesink, Michael Dunn
9 Inferring Prehistory from Genetic, Linguistic, and Geographic Variation 141 Keith Hunley, Michael Dunn, Eva Lindström, Ger Reesink, Angela Terrill, Heather Norton, Laura Scheinfeldt, Françoise R. Friedlaender, D. Andrew Merriwether, George Koki, Jonathan S. Friedlaender
Part III Regional Studies and Conclusion 10 Animal Translocations, Genetic Variation, and the Human 157 Settlement of the Pacific Elizabeth Matisoo-Smith
11 Viral Phylogeny and Human Migration in the Southwest Pacific 171 Jill Czarnecki, Jonathan S. Friedlaender, Gerald Stoner
12 Origins of Plant Exploitation in Near Oceania: A Review
181
Robin Allaby
13 Extraordinary Population Structure among the Baining of New Britain 199 Jason A. Wilder and Michael F. Hammer
14 Immunoglobulin Allotypes as a Marker of Population History in the Southwest Pacific 208 Moses S. Schanfield, Frank B. Austin, Peter B. Booth, D. Carlton Gajdusek, Richard W. Hornabrook, Keith P. W. McAdams, Jan J. Saave, Susan W. Serjeantson, Graeme W. Woodfield
15 Contributions of Population Origins and Gene Flow to the Diversity of Neutral and Malaria Selected Autosomal Genetic Loci of Pacific Island Populations 219 J. Koji Lum
16 Conclusion
231
Jonathan S. Friedlaender
Index
viii
239
Contributors
Robin Allaby Assistant Professor, Warwick Horticulture Research International, University of Warwick, Wellesbourne, CV35 9EF, UK
Frank B. Austin Formerly of the University of Otago, Dunedin, New Zealand
Peter B. Booth (deceased) Formerly of the Christchurch Hospital, Christchurch, New Zealand
Jill Czarnecki National Institute of Neurological Disorders and Stroke, National Institutes of Health, Bethesda, MD 20892. E-mail:
[email protected] Head of Immunodiagnostics, Biological Defense Research Directorate, Naval Medical Research Center, Silver Spring, MD 20910. E-mail:
[email protected]
Michael Dunn Research Fellow, formerly Department of Linguistics, Research School of Pacific and Asian Studies, Australian National University, Canberra ACT 0200, Australia; currently Centre for Language Studies, Radboud University Nijmegen, Erasmusplein 1, 6525 HT Nijmegen, The Netherlands. E-mail:
[email protected]
Françoise R. Friedlaender Independent Scientist. 7 North Columbus Boulevard, Philadelphia, PA 19106. E-mail:
[email protected]
Jonathan S. Friedlaender Emeritus Professor of Biological Anthropology, Temple University, Philadelphia, PA 19122. E-mail:
[email protected]
D. Carlton Gajdusek Formerly of the NINDS, National Institutes of Health, Bethesda, MD
List of Contributors
Michael Hammer Director, Genomic Analysis and Technology Core Facility, and Professor in the Departments of Anthropology and Ecology and Evolutionary Biology, University of Arizona, Tucson, AZ 85721. E-mail:
[email protected]
Jason A. Hodgson Department of Anthropology, University of Binghamton, Binghamton, NY 13902-6000. E-mail:
[email protected]
Richard W. Hornabrook Formerly of the Papua New Guinea Institute of Medical Research, Goroka, New Guinea
Keith Hunley Assistant Professor of Anthropology , University of New Mexico Albuquerque, NM 87131. E-mail:
[email protected]
Joseph Lorenz Research Associate, Molecular Biology Laboratory, Coriell Institute for Medical Research, Camden, NJ 08103. E-mail:
[email protected]
Tatiana Karafet Genomic Analysis and Technology Core Facility, University of Arizona, Tucson, AZ 85721. E-mail:
[email protected]
George Koki Department of Human Genetics, Papua New Guinea Institute for Medical Research, Goroka, Papua New Guinea. E-mail:
[email protected]
Krista Latham Department of Anthropology, Temple University, Philadelphia PA 19122. E-mail:
[email protected]
Eva Lindström Research Fellow, Department of Linguistics, Stockholm University, 106 91 Stockholm, Sweden. E-mail:
[email protected]
J. Koji Lum Associate Professor of Anthropology and Biological Sciences, Laboratory of Evolutionary Anthropology and Health, Binghamton University, Binghamton, NY 13902-6000. E-mail:
[email protected]
Elizabeth Matisoo-Smith Senior Lecturer in Biological Anthropology and Principal Investigator, Allan Wilson Centre for Molecular Ecology and Evolution, Department of Anthropology, University of Auckland. E-mail:
[email protected]
Keith P. W. McAdams Formerly of the Papua New Guinea Institute of Medical Research, Goroka, New Guinea
Stacy McGrath Department of Anthropology, University of Binghamton, Binghamton, NY 13902-6000. E-mail:
[email protected]
D. Andrew Merriwether Associate Professor of Anthropology, University of Binghamton, Binghamton, NY 13902-6000. E-mail:
[email protected] x
List of Contributors
Heather Norton Post-doctoral Fellow, University of Arizona, Tucson, AZ 85721. E-mail:
[email protected]
Andrew Pawley Professor of Linguistics, Research School of Pacific & Asian Studies, Australian National University, Canberra, ACT 0200. E-mail:
[email protected]
Ger Reesink Research fellow, Centre for Language Studies, Radboud University Nijmegen, Erasmusplein 1, 6525 HT Nijmegen, The Netherlands. E-mail:
[email protected]
Renato Robledo Associate Professor, University of Cagliari, Cagliari 09100, Italy. E-mail:
[email protected]
Jan J. Saave Formerly of the Territory Papua New Guinea Health Service, Sydney, Australia
Moses S. Schanfield Chair and Professor, Department of Forensic Sciences, George Washington University, Washington, DC 20052. E-mail:
[email protected]
Laura Scheinfeldt Post-doctoral Fellow, Department of Human Genetics, Children’s Hospital of Philadelphia, PA 19104. E-mail:
[email protected]
Theodore G. Schurr Assistant Professor, Department of Anthropology, University of Pennsylvania, Philadelphia, PA. E-mail:
[email protected]
Susan W. Serjeantson Australian Academy of Sciences
Matthew Stoltz Department of Anthropology, University of Binghamton, Binghamton, NY 13902-6000. E-mail:
[email protected]
Gerald Stoner (deceased) Formerly of the NINDS, National Institutes of Health, Bethesda, MD
Glenn Summerhayes Head, Department of Anthropology, University of Otago, Dunedin. E-mail:
[email protected]
Angela Terrill Research Fellow, Centre for Language Studies, Radboud University Nijmegen, Erasmusplein 1, 6525 HT Nijmegen, The Netherlands. E-mail:
[email protected]
Jason Wilder Assistant Professor of Biology, Williams College, Williamstown, MA 01267. E-mail:
[email protected]
Graeme W. Woodfield Former Director, Papua New Guinea Red Cross Blood Transfusion Service, Boroko, Papua New Guinea xi
This page intentionally left blank
part i The Framework
This page intentionally left blank
1 Introduction Jonathan S. Friedlaender
While its significance for understanding human prehistory is much broader, the focus of this book is actually rather narrow: reconstructing the population history of Northern Island Melanesia, primarily through analyses of contemporary genetic and linguistic variation, but with reference to the archeological record. Although often overlooked and often assumed to be genetically impoverished, this region just to the east of New Guinea is truly critical for understanding the genetics and prehistory of the Pacific. At first, this group of large islands was at the easternmost edge of the human species range, and as a result retained a number of distinctive relic populations in semi-isolation. Then tens of thousands of years later, it served as the homeland for the remarkable series of explorations leading to the settlement of the uninhabited islands of the rest of the Pacific by the ancestors of the Polynesians. Identifying the genetic composition of these various populations has been a very contentious issue, but no comprehensive sampling in this core area has been done before this time.
Some Themes Before introducing the individual chapters, I will describe some central, but generally unstated, themes of the book.
The Geographic Setting: The End of the Line, and Also the Launching Pad The Bismarck Archipelago and Bougainville Island are near the eastern end of the great island chain that extends from Southeast Asia out towards the Central Pacific, thrown up from the collision of major tectonic plates. Almost all the islands in the chain are “intervisible,” or
within sight of neighboring islands. The implied easy movement across these stepping stones ends with the Solomon Islands, where larger water crossings sometimes occur, particularly at the southeastern end. Moving further east into the central Pacific, the islands themselves rarely attain the considerable size of the major continental islands of Melanesia and Indonesia, and the water crossings are much longer. As a result, people were able to move as far east as sections of Northern Island Melanesia very early in modern human prehistory, at more than 40,000 years ago. This area, along with New Guinea, has been referred to as Near Oceania, contrasting it with the then uninhabited islands further to the east, in Remote Oceania (Green, 1991). Chapter 3 has an extended description of the biogeographical regions here (see particularly figure 3.1). Sea levels were considerably lower then, so that there were even fewer major water gaps to cross between the larger land masses in Near Oceania and Island Southeast Asia. Major sections of Island Southeast Asia and Taiwan were then a single land mass (referred to as Sundaland) that remained separated from the ancient continent of Sahul, which merged today’s New Guinea and Australia. In Northern Island Melanesia, the small water gaps persisted throughout the Pleistocene, except, importantly, for an enlarged Bougainville Island, which included nearby islands of the Solomons almost as far as Guadalcanal. From the beginning, then, people in this region were relatively isolated from major population events that swept through Eurasia. Their peripheral position continued for over 30,000 years thereafter—an unparalleled situation that has played the decisive force in shaping the remarkable variability of the peoples inhabiting these islands today. However, this region was not just an isolated backwater to human cultural and population developments.
3
the framework
New Guinea and Northern Island Melanesia clearly became regional centers of plant, animal, and human diversity during this very long period. Also, the populations in this region were not without their own internal developments and dynamics.
Dynamics of Small but Mobile Populations Contemporary Northern Island Melanesian populations are extraordinarily diverse. The underlying cause has been assumed to be the isolation of the small clusters of extended family units that was typical over most of the region. This should be an overriding dynamic causing diversity to accumulate if it could be generalized over time. For example, in the 1960s, I interviewed people in inland Bougainville villages and found that most of the villagers (men and women alike) had married and settled within a few kilometers of their birthplaces (see figure 1.1, and Friedlaender, 1975). Jeff Long reported much the same thing among the Gainj in the New Guinea Highlands (Long et al., 1986), so this was probably a general regional pattern, at least for inland areas. In 2003, George Koki and I visited beach villages in north Bougainville and collected comparable marital migration data there as well, and the results were very different. In that recent series, many more individuals settled further away from their birthplaces (figure 1.2). While the greater movement may have partly been caused by the availability of more roads, trucks, and outboard motors, it may reflect a longer-term contrast between “beach” and “bush” populations. It should have been far easier for people to move along the coastlines by boat or foot than to slog over the rugged inland terrain. If this were a longlasting dichotomy, beach-living groups on a large island would be expected to be genetically similar, while bush groups would be expected to develop, or at least to retain, greater genetic distinctiveness. Figure 1.1 Marital migration rates for inland Bougainville villages in 1967.
4
Figure 1.2 Marital migration rates for shoreland Bougainville villages in 2003.
On the other hand, there is no question that populations here have moved around a great deal over many centuries. Of course, the last 50 years have witnessed major population shifts, with the Second World War, mining, logging, road construction, and civil strife all causing relocations in many instances. Also, recently established centralized high schools have been a vector for contacts that lead to relocation and marriage away from birth villages. Our simple family questionnaires could take account of many of these recent residential changes (so that, for example, our New Guinea sample consisted of men born in New Guinea who had married into this region), but the ethnographic record suggests there were always considerable population displacements and movements in the region. In the first place, New Britain and Bougainville have both had many episodes of population extinction and subsequent replacement because of extensive volcanic activity. Concerning our New Britain series, the Tolai migrated centuries ago into the Gazelle Peninsula from southern New Ireland subsequent to a major eruption, displacing and sometimes enslaving the Baining who had occupied that region; the Sulka recently moved their settlements from one part of Wide Bay to another; and the Anêm and Kove both moved to their current locations in far West New Britain from regions much further east, also an area of major volcanic activity. In New Ireland, the Kuot have apparently been displaced southward by the northern Nalik, and the Notsi established settlements on the east coast, coming from Tabar Island (Lindström, 2002). In Bougainville, there was a history of a number of population displacements and movements in different regions, especially near the active volcanoes in the north, and along the coasts (most recently reviewed in Nash, 2005; Ogan, 2005; Spriggs, 2005).
Introduction
Also, interior population aggregates must have been smaller and more dispersed than coastal ones, to judge by recent village census information, the relative abundance of shoreline resources, and archeological settlement evidence.
Intensifying Population Mobility Over Time Archeological evidence suggests that, while such a beach–bush dichotomy in movement was likely, it intensified over time. Once the first small populations colonized this region, probably drawn there by lush marine and avian resources, some groups did begin to exploit the large island interiors, at first only intermittently. Our genetic data are consistent with not just one, but a number of small population movements into the region from the time of first settlement, most immediately from New Guinea, but in a number of instances, showing no genetic relationship to anything found there or even further to the west. Over the last 20,000 years, there are a number of indicators, archeological and genetic, that suggest this was a period of developing exploitative abilities of these ancient peoples, with increasing trade and communication among islands (although there appears to have been a period of increased sedentism subsequent to first settlement—see figure 2.5 and accompanying text). However, the rate of penetration of the large island interiors could not have been markedly transformed from the earliest days. The introduction of Oceanic languages and the subsequent development of the Lapita Cultural Complex in some coastal environments of the region at 3,500–3,200 years ago certainly represented a major transformation, but it was hardly the only outside influence to make itself felt in this region during preceding times.
The Relevance of Language Distributions and Distinctions One cannot assume that people speaking one language will have the same general genetic constitution (just think of the ethnic mix of native English speakers in the United States, Australia, or New Zealand). Likewise, language boundaries or distinctions may not noticeably impede intermarriage rates. In the Bougainville marital migration study, I found that, in villages near language boundaries, marriages across boundaries happened only slightly less frequently than those at equivalent distances within the language area. This was probably because people at the boundaries were often bi- or trilingual. Nevertheless, language distributions can still be very informative with regard to prehistory and old population relationships. For instance, agricultural revolutions and the associated population expansions have been linked to major language expansions (Bellwood, 2002), and in Northern Island Melanesia, a tie has been established
between the development of Oceanic languages and the Lapita Cultural Complex some 3,300 years ago (Bellwood et al., 1995). Some simple associations between particular genes and languages have also been proposed in the Pacific, most notably between Oceanic language dispersals and a particular mitochondrial DNA variant (the “Polynesian Motif”), which will be reconsidered in chapters 4 and 5. The Oceanic languages predominate in most of Island Melanesia and certain coastal regions of New Guinea, but other languages, called non-Austronesian or Papuan, are spread over most of New Guinea and in certain scattered sections (often inland) of Island Melanesia. This distribution, and the remoteness of any ties of the various Papuan languages, reflects their status as descendants of earlier settlers and linguistic strata in the region. There have been indications that the Papuan groups were, at least in some cases, genetically distinguishable from Oceanic-speaking populations (Giles et al., 1965; Melton et al., 1995), but not always (Serjeantson, 1985). My early work in Bougainville suggested the major genetic distinction on that island was actually between different Papuan-speaking groups, so that this entire matter of language and gene associations in the region is clearly not a simple one. A number of chapters of this book address this issue, and we now have a much better understanding of the relationship between language and genetic differentiation here. In general, we find Oceanic-speaking groups, especially the coastal-dwelling ones, are more intermixed, while some Papuan-speaking groups, particularly the more inland ones, retain the greatest genetic distinctiveness.
Chapter Descriptions Chapters in this volume represent both review articles of recent innovative approaches and reports of new findings pertaining to Island Melanesian uniqueness and importance as a center of dispersal. They also show the strength of the traditional anthropological and linguistic comparative approaches to prehistory. This volume consist of three parts: The Framework (chapters 1–3), Core Studies in Northern Island Melanesia (chapters 4–9), and Regional Studies and Conclusion (chapters 10–16), which all address the general issue of the demographic processes that affected the peopling of the Southwest Pacific Islands.
Part I: The Framework In chapter 2, Glenn Summerhayes presents a comprehensive review of the archeology of this region. Its most pervasive
5
the framework
theme is the developing network of interactions between the islands of Near Oceania over its 40,000+ years of settlement. Summerhayes is able to identify incipient trading networks and spheres of influence across the region, dating well back into the Upper Pleistocene. Over the more recent period, including the entire Holocene, there is considerable evidence for intensification of these networks and influences. From this perspective, the Lapita phenomenon at 3,300 years ago, while still something quite new and transforming, clearly builds on many developing local themes and represented only the latest (and most important) influence from Sunda, beyond the Wallace Line. The chapter is an argument against earlier simplistic notions of a two-stage settlement history that envisioned an initial settlement by a homogeneous population perhaps 40,000 years ago that persisted in isolation until an intrusion by a distinctive Southeast Asian/ Taiwanese population with a completely distinctive cultural complex (the Lapita phenomenon). Instead, there are clear archeological indications of introductions and innovations to Northern Island Melanesia subsequent to initial settlement at ~20,000 years ago, and again in the Holocene at ~8,000 years ago, both well before Lapita (but apparently not always extending as far as Bougainville). This theme of developing complexity and interaction is carried through to the following chapters on population genetics and linguistics. In chapter 3, Andrew Pawley discusses historical linguistic connections among the “Papuan” languages of Near Oceania, “Papuan” being a residual category consisting of the more than 700 non-Austronesian languages of New Guinea, Island Melanesia and East Nusa Tenggara. Pawley refers to recent work by Ross (2005) which indicates that there are, in fact, over 20 separate language families within this residual category, none of which can be convincingly shown to be related, together with a number of isolates (single member families). This degree of linguistic diverisity is unsurpassed in any other region of comparable size in the world. Indeed, it exceeds that of the whole of Africa. However, recent research by Pawley and Ross, using the classical Comparative Method, has confirmed the existence of a large Trans New Guinea language family, which had previously been proposed on more speculative grounds. Members of the Trans New Guinea family occupy the central mountain cordillera of New Guinea and some regions to the north and south of this. Pawley suggests that the creation of this large family is associated with a farming-based population expansion in that region during the last 10,000 years. The major concentrations of unrelated Papuan language families occur in the Sepik/Ramu river regions, along the north coast to the west of the Sepik as far as the Bird’s Head, and in Northern Island Melanesia. These concentrations of diverse families appear to be relics of ancient
6
linguistic strata, the product of in situ diversification that began in each region in the late Pleistocene. The chapter also looks in some detail at how far linguistic evidence complements, corroborates or contradicts the evidence of archeology and certain other historical disciplines such as geomorphology and paleo-botany.
Part II: Core Studies in Northern Island Melanesia The major theme of this part is that there is an enormous amount of population genetic structure in Northern Island Melanesia that is consistent to some degree across genetic systems regardless of whether the loci are maternally, paternally, or biparentally inherited. The authors argue that this general pattern is determined by: (a) the very ancient historical settlement pattern of this region, which consisted of a series of population overlays; (b) subsequent relative population isolation in the region, illustrated by the extremely limited marital migration rates in island interiors; and (c) developing contacts along the coastlines in the Holocene that have, to some extent, blurred earlier distinctions among populations. Included in this last phase are the effects of recent population movements including (and also subsequent to) the arrival of the Lapita People. Chapter 4 reports the (maternally inherited) mitochondrial DNA diversity in the Southwest Pacific and then concentrates on the extensive structure of this variation in the core region, Northern Island Melanesia. This chapter shows that a constellation of mitochondrial variants in Northern Island Melanesia is particularly old and not found anywhere to the west, beyond the Wallace Line. These variants must have developed subsequent to initial settlement some 40–50,000 years ago. Also, the mitochondrial DNA (mtDNA) evidence suggests a subsequent series of expansions into the region from the west, through the Upper Pleistocene and into the Holocene. The most recent expansions concern haplogroup E and the so-called Polynesian Motif (haplogroup B4a1a1 in technical terminology). This “Motif” clearly developed in Near Oceania from a haplogroup that was introduced from Island Southeast Asia, and is closely associated with the Lapita phenomenon. There are some oddities about the distribution of the “Motif” in Island Melanesia that still require explanation, however. Overall, the mitochondrial DNA population diversity is organized on a clear island-by-island basis, with the Papuan-speaking groups of the island interiors showing the greatest diversity, and the Oceanic-speaking groups on the coastlines the least. Chapter 5 reports the paternally inherited Y chromosome variation (specifically its non-recombining fraction, or NRY), following the same format as for the mitochondrial DNA. Using an expanded battery of regionally
Introduction
informative markers, this chapter greatly expands the known NRY chromosome variation in the Southwest Pacific, and shows that, in fact, the paternally inherited variation in the region is generally comparable to the maternally inherited mtDNA pattern of variation—contrary to hypotheses that had been put forward in the last decade. Again, there is a constellation of variants that seem to have arisen tens of thousands of years ago within Near Oceania and specifically in Northern Island Melanesia. There is an identifiable footprint of recent Island Southeast Asian influence in the NRY data, but it is considerably fainter than the mtDNA “Polynesian Motif.” Chapter 6 reviews the complementary skin and hair pigmentation variation across the core study area of Northern Island Melanesia. Technical advances in reflectance instruments now allow for detection of considerable regional variation in pigmentation both in the skin and hair that had not been detectable using the crude reflectance tools of earlier generations. This chapter reveals an island-by-island cline in skin pigmentation, with increasing M Index (heavily pigmented) values towards Bougainville Island. This skin color gradient is apparent in the portraits in the color insert. In fact, the skin pigmentation M values for Bougainville populations are as high as any surveyed population elsewhere, including West Africans. Also, Papuan speakers in different islands tend to have somewhat lighter hair pigmentation than their Austronesian-speaking neighbors, according to this account. The distribution of six candidate genes for possible association/causation with pigmentation suggests that at least two (OCA2 and ASIP) vary in a fashion that suggests an association with melanin phenotype variation in this region. While natural selection clearly must have an effect on pigmentation in this intensely irradiated region so close to the equator, it clearly does not dictate the pattern of melanin variation among these groups, which must be the result of ancient population associations. In chapter 7 Renato Robledo presents another comparative genetic dataset, this time a particularly informative autosomal variant (an insertion/deletion on chromosome 22) that is presumed to be selectively neutral. It has a highly variable distribution across this set of populations in Northern Island Melanesia, and is known to be polymorphic world-wide. It is monomorphic in the more inaccessible regions of Bougainville, New Britain, and New Guinea. Chapter 8 gives a state-of-the-art overview of what is known about the linguistic characteristics and internal relations of each of the two overarching groups of languages in Island Melanesia: Papuan languages and the Oceanic branch of the Austronesian family, as well as phenomena arising through contact between these groups. It also shows by what methods linguistics can
contribute to our understanding of the history of languages and speakers, and what the findings of those methods have been to date. Several scholars have successfully applied the Comparative Method to the Oceanic languages, and have been able to form strong hypotheses as to the location of the homeland of speakers of ProtoOceanic (in northeast New Britain); many facets of the lives of those speakers; as well as the patterns of their subsequent spread across Island Melanesia and beyond into Remote Oceania, followed by a second wave overlaying the first into New Guinea and as far southeast as halfway through the Solomon Islands. Regarding the Papuan languages of this region, at least some of them clearly go back much further in time than the 6,000–10,000 ceiling of the Comparative Method, and the authors explore the linguistic relations of these languages with the aid of a database of 125 non-lexical structural features. Their results reflect archipelago-based clusterings, with the Central Solomons Papuan languages forming a clade either with the Bismarcks or with Bougainville languages. They also investigate contact issues and, among other things, find Papuan languages in Bougainville less influenced by Oceanic languages than those in the Bismarcks and the Solomons. The authors consider a variety of scenarios to account for their findings, concluding also that their results are compatible with multiple pre-Oceanic waves of arrivals into the area after initial settlement. Chapter 9 investigates the fit of genetic, phenotypic, and linguistic data to two well-known models of population history. The first of these models, termed the population fissions model, emphasizes population splitting, isolation, and independent evolution. It predicts that genetic and linguistic data will be perfectly tree-like. The second model, termed isolation by distance, emphasizes genetic exchange among geographically proximate populations. It predicts a monotonic decline in genetic similarity with increasing geographic distance. While these models are overly simplistic, deviations from them were expected to provide important insights into the population history of northern Island Melanesia. The authors examine the fit of the population genetic, phenotypic, and linguistic data to these models using established methods. They find scant support for either model. Additional analyses combined with archeological, ethnographic, and historical data reveal that neither model fits because the prehistory of the region has been so complex. This prehistory includes population fissions, long-range movements, temporal and spatial variation in migration rates and patterns, and differing histories of genetic and linguistic exchange. Nonetheless, the genetic and linguistic data are consistent with an early radiation of proto-Papuan speakers into the region followed by a much later migration of Austronesian-speaking peoples.
7
the framework
While these groups subsequently experienced substantial genetic and cultural exchange, this exchange has been insufficient to erase this history of separate migrations. This chapter also emphasizes the need for higher resolution genetic data to more accurately assess the prehistory of the region.
Part III: Regional Studies and Conclusion Chapters in this final part place the intensive study of human genetic variation in Island Melanesia in the context of genetic variation across the Pacific region. Chapter 10 presents a complementary genetic approach to population relationships across the Pacific, utilizing information from animals closely affiliated with humans. Matisoo-Smith, who is the primary innovator in this area, describes how her analyses of genetic variation in commensals (the Pacific rat, pig, dog, and chicken) are being used as a proxy for understanding prehistoric human mobility and contacts. In particular, mitochondrial DNA studies of the Pacific rat, Rattus exulans, are providing intriguing insight into the relationships and level of interactions among Near and Remote Oceanic human populations. These are also providing valuable data on the timing and degree of population interactions in the region. The basic conclusion of her work is that there has been considerably more continued interaction between populations in different areas of the Pacific than many suspected before, and this includes interactions between “Near” and “Remote” Oceania. Chapter 11 deals with another sort of human commensal, viruses, which are providing additional insights into human contacts throughout the region. This is particularly true for viruses that infect many people, have low rates of re-infection, and remain in the host for decades, if not lifetimes, without causing major illness. The focus of this chapter is the JC virus, which is primarily spread within families and apparently remains in the host for a lifetime—mimicking human genes in this regard. Nevertheless, viruses seem to be most informative concerning more recent contacts among human populations, although this certainly varies widely. In Chapter 12, Robin Allaby reviews the developing evidence from archeobotany (including the molecular evidence) on the history of plant exploitation in Near Oceania. The old notion that most domesticate crops were imported from Southeast Asia is not borne out by the botanical evidence. Rather, many of the principal crops of Near Oceania appear to have been domesticated locally, and over a time period that predates the arrival of the Proto-Oceanic/Lapita cultures. This evidence represents another corollary to the dynamic nature of population development in the region with a sophistication attributed to earlier peoples of the region previously
8
unacknowledged under the old paradigm of a two-wave colonization of Oceania. Chapter 13 sets out to examine in greater detail (with four unlinked loci) the extent of divergence between two linguistically related Baining groups in New Britain (speakers of the Mali and Kaket dialects). Although they are linguistically related and are less than 100 km apart, they are, by a number of measures, surprisingly different genetically. This dramatically emphasizes the overall population structure of the total Northern Island Melanesian dataset analyzed in Part II. Wilder and Hammer attempt to explain this Baining difference in terms of male and female demographic distinctions (their marital migration rates and effective population sizes). Early comparisons in global and regional mtDNA and NRY diversity indicated comparatively greater overall mtDNA variability, but greater amonggroup NRY variation. This led to considerably older estimates of mtDNA coalescence values, but greater population distinctions in the NRY. An early proposed explanation for the global distinction in population differentiation was that rates of migration among groups were generally less for males than females. However, more recently, this argument has been disputed by Wilder and others, who suggest that the key factor that can explain both discrepancies is the larger effective population size of women (since relatively few men contribute to following generations). This distinction could cause an acceleration in the effects of genetic drift, leading to less overall variation, but proportionately more among-group variation. In the Baining study, they find evidence for a much smaller male effective population size (only a third of the effective female population size, or 25% of the total). However, the proportion of males who migrate and successfully reproduce appears to be greater than for females. In considering the surprising degree of overall differentiation between these two Baining groups, the effects of drift are paramount, but there remains the question of whether the differences may be due to the residue of ancient lineage sorting. Chapter 14 reviews variation across the region for the single most informative polymorphism from the autosomes, the GM locus. There are over 11,000 samples analyzed for GM from Pacific populations. Their pattern of variation reinforces some themes established in the mitochondrial DNA. There is a GM connection between Australian Aboriginal and certain New Guinea populations; between Southeast Asian and Austronesian groups; and between certain Eastern Highlands New Guinea and Island Melanesian populations. The GM pattern also indicates a substantial level of intermixture between Austronesian- and Papuan-speaking populations in Northern Island Melanesia, which is also echoed in the mitochondrial DNA.
Introduction
In Chapter 15, Koji Lum deals with the effects of differential selection as well as population history simultaneously. Malaria is the only established important differential selective agent in the region, and Lum reviews three studies of either neutral or malaria-selected autosomal loci to illustrate the contribution of population origins, gene flow, and disease selection to genetic variation. Neutral genetic diversity within populations generally decreases with distance from Southeast Asia (with the exception of Austronesian-speaking populations of Vanuatu) resulting in a clustering of Melanesian populations regardless of linguistic affiliation. The malariaresistant B3∆27 allele is currently restricted to coastal areas of Papua New Guinea reflecting its origin with the Lapita colonists and a paucity of gene flow from the coast to inland areas. Within Vanuatu, an archipelago settled during the initial Lapita expansion, a North–South gradient of malaria endemicity has resulted in corresponding gradients of resistance and susceptibility conferring alleles. Nearly all of the α-thalassemia alleles of Vanuatu are of inferred PNG origin, consistent with an accumulation of alleles from Near Oceania over time. Chapter 16—the Conclusion—summarizes major findings of the book and then discusses the contradictions, the problems of interpretation, and the road forward. The pervasive diversity of populations in this relatively small region has been caused by their semiisolation over an extremely long time period, extending back tens of thousands of years. The diversity has an underlying pattern, with more distinctive populations in large island interiors, primarily Papuan-speaking groups. This presents a striking contrast to the comparative homogeneity of most human groups, and offers an alternative model for world-wide prehistoric human population variation.
References Bellwood P. 2002. Farmers, foragers, languages, genes: The genesis of agricultural societies. In: Bellwood P, Renfrew C, editors. Examining the farming/language dispersal hypothesis. Cambridge MA: McDonald Institute for Archaeological Research. pp 17–28. Bellwood P, Fox JJ, Tryon D. 1995. The Austronesians in history: Common origins and diverse transformations. In: Bellwood P, Fox JJ, Tryon D editors. The Austronesians: Historical and comparative perspectives. Canberra, Australia: Australian National University (Department of Anthropology). pp 1–16.
Friedlaender JS. 1975. Patterns of human variation: The demography, genetics, and phenetics of Bougainville Islanders. Cambridge, MA: Harvard University Press. Giles E, Ogan E, Steinberg AG. 1965. Gamma-globulin factors (Gm and Inv) in New Guinea: Anthropological significance. Science 150: 1158–60. Green RC. 1991. Near and remote Oceania—disestablishing “Melanesia” in culture history. In: Pawley A, editor. Man and a half: Essays in Pacific anthropology and ethnobiology in honour of Ralph Bulmer. Auckland, New Zealand: The Polynesian Society. pp 491–502. Lindström E. 2002. Topics in the grammar of Kuot, a non-Austronesian language of New Ireland, Papua New Guinea. Department of Linguistics (doctoral dissertation). Stockholm, Sweden: University of Stockholm. Long JC, Naidu JM, Mohrenweiser HW, Gershowitz H, Johnson PL, Wood JW, Smouse PE. 1986. Genetic characterization of Gainj- and Kalam-speaking peoples of Papua New Guinea. American Journal of Physical Anthropology 70: 75–96. Melton T, Peterson R, Redd AJ, Saha N, Sofro ASM, Martinson J, Stoneking M. 1995. Polynesian genetic affinities with Southeast Asian populations identified by mtDNA analysis. American Journal of Human Genetics 57: 403–14. Nash J. 2005. Nagovisi then and now, 1963–2000. In: Regan A, Griffin H, editors. Bougainville before the conflict. Canberra, Australia: Pandanus Press, RSPAS. pp 400–9. Ogan E. 2005. An introduction to Bougainville cultures. In: Regan A, Griffin H, editors. Bougainville before the conflict. Canberra, Australia: Pandanus Press, RSPAS. pp 47–56. Ross M. 2005. Pronouns as a preliminary diagnostic for grouping Papuan languages. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts: Investigations into the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra: Pacific Linguistics. pp 15–66. Serjeantson S. 1985. Migration and admixture in the Pacific: Insights provided by human leukocyte antigens. In: Kirk RL, editor. Out of Asia: peopling the Americas and the Pacific. Canberra, Australia: Australian National University Press. pp 133–54. Spriggs M. 2005. Bougainville’s early history: An archaeological perspective. In: Regan A, Griffin H, editors. Bougainville before the conflict. Canberra, Australia: Pandanus Books, RSPAS. pp 1–19.
9
2 Island Melanesian Pasts: A View from Archeology Glenn R. Summerhayes
This chapter outlines what we currently know from the archeological record about the colonization and contacts among different islands and regions in New Guinea and Island Melanesia. Modeling past connections between the groups of peoples that inhabited the western Pacific since first human occupation is not easy. When identifying societal interactions within a region, archeologists have relied on changes in the archeological records that may be part of a wider regional trend, or the identification of past movement of materials that may indicate parallel movements of peoples or ideas as well. In the chapter I will address the nature of past societal interactions within Papua New Guinea and the Bismarck Archipelago by identifying the nature of the transfer or movement of goods between geographically separated areas. Societal interactions here are an indicator of the closed or open nature of the societies inhabiting this region for over 40,000 years. Five discrete time periods will be used. The first concerns the initial colonization of New Guinea and the Bismarck Archipelago prior to 40,000 years ago. The second occurred some 20,000 years later when the movement of objects and people becomes highly visible. The third centers in the early to mid-Holocene, while the fourth looks at the Lapita phenomenon and the introduction of Austronesian-speaking peoples into the region. The last briefly looks at the regionalization that occurred within the region over the last 2,000 years and the development of the eclectic nature of societies we see today.
10
Colonization and Early Interactions Earliest Occupation in New Guinea and Western Melanesia The evidence available suggests that the initial colonization of New Guinea and the Bismarck Archipelago was rapid with little time separating the earliest dates for occupation on mainland New Guinea and New Ireland to the east (figure 2.1). From mainland New Guinea the earliest evidence for human occupation is found on the reef terraces of the Huon Peninsula where a number of stone tools, called “waisted blades,” were found in contexts underlying volcanic tephras dated by thermoluminescence to just under 40,000 BP (Groube et al., 1986; O’Connell and Allen, 2004). Similar age estimates based on radiocarbon dating have been reported for Lachitu Cave on the north coast of New Guinea between Vanimo and the present border with Indonesia (Gorecki et al., 1991; Gorecki, personal communication; Chappell, 2000: Table 1). Occupation in the mountainous interior occurs during the following ten millennia. The highland sites of Nombe located at 1669 m above sea level (Gillieson and Mountain, 1983: 55; Mountain, 1991) and Kosipe at 2,000 m above sea level (White et al., 1970) were both occupied by 30,000 years ago (see Chappell, 2000: Table 1, for calibration details). Occupation of the Bismarck Archipelago, on the other hand, occurs simultaneously with the earliest dates for lowland New Guinea. On New Britain the earliest evidence for occupation is found at
Island Melanesian Pasts: A View from Archeology
Figure 2.1 New Guinea and the Bismarck Archipelago showing the earliest human occupation sites.
35,000 BP at the site of Yombon that lies 35 km in from the south coast (Pavlides and Gosden, 1994). A possible second early colonization site, Kupona na Dari, was located and excavated at the base of the Willaumez Peninsula, and is argued to be between 35–45,000 years old using non-radiocarbon techniques (Torrence et al., 2004). From New Ireland to the east, the earliest occupation is earlier, at 40,000 BP at the site of Buang Merabak, a cave site from the east coast halfway down the island (Leavesley et al., 2002; Leavesley and Chappell, 2004). Further south also just in from the coast, the site of Matenkupkum (Summerhayes and Allen, 1993; Allen and Gosden, 1996: 186) has been dated to over 35,000 years (Chappell, 2000: Table 1). Occupation of the northern Solomon region soon followed with evidence for the occupation of Kilu cave on Buka Island at 32,000 BP (Wickler and Spriggs, 1988). Unfortunately little archeological work has been undertaken in the western half of New Guinea (West Papua, currently part of Indonesia). The earliest archeological evidence for people here is only at 26,000 BP (uncalibrated) at Toe Cave, Ayamaru region, on the central Bird’s Head of West Papua (see figure 2.1) (Pasveer, 2004). No doubt earlier occupation dates will be found there in the future. The evidence suggests that these early colonizers did not waste time in moving to the islands east of New Guinea soon after colonizing the mainland. Such movements
required some sea technology as the island of New Britain was never joined to mainland New Guinea, and New Britain was not joined to New Ireland, nor was the North Solomons to either New Ireland or New Britain.
Life in the Pleistocene This period of time has been described as a “world without any ethnographic parallel” that lasted for over 10,000 years (Gosden, 1993: 133). The evidence suggests that the earliest colonists were “small groups of mobile, broad-spectrum foragers” that exploited both maritime and terrestrial resources (Allen et al., 1989: 558–9; Allen and Gosden, 1996: 187). Jim Allen (2000) makes the point that these early migrants moving from Wallacea to mainland Papua New Guinea and then the Bismarck Archipelago would have focused on the coastal resources to which they had adapted so well. Evidence for this is found in all the early midden sites such as Lachitu, Matenbek, Matenkupkum, and Buang Merabak. It is important to note that terrestrial diet was also important, being made up of small animals (birds and bats) and a variety of reptiles (Leavesley and Allen, 1998). There is evidence of occupation away from the coasts in the interior of New Britain, at Yombon, but it seems likely this was a temporary occupation site because of the apparent lack of a subsistence base (see Bailey et al., 1989; Sillitoe, 2002).
11
the framework
Would coastal living fare any better? Yes; however, even here these resources would only have supported a small mobile population with low consumption levels. In many later Pacific sites, there is a marked change in shell size over time, with the largest of the shells becoming depleted, indicating a high level of human predation. Allen (2003: 35), noting the lack of change in the size of shells over the first 10,000 years of occupation in New Ireland, argued for low predation levels. Gosden (1993: 132), noting the small number of early sites in New Ireland, also pointed out the paucity of material in these earliest assemblages compared with occupation deposits later on. The evidence for such small mobile groups of people spread over wide geographical areas during this time period led Gosden to not only argue that they kept regular contact with each other in order to survive, but also that they were not “socially bounded groups” but socially mobile (Gosden, 1993: 133). These early colonists to New Guinea would not have encountered a different habitat to the ones they left, simply continuing to follow the coastlines of new islands. It is only when they moved inland to the upper altitudes that they showed new adaptation skills. Forty thousand years ago temperatures were 4–7°C cooler than today (Haberle et al., 1990: 36; Hope and Golson, 1995; Bird et al., 2004: 150; Haberle and David, 2004: 167). Sites such as Kosipe or Nombe would have been just below the tree line and on the boundary of the forest and grassland edges where people could hunt game from both ecological zones. Beech forests (Nothofagus) at these heights indicate cloudiness and mist (Hope and Golson, 1995: 820). Glaciers were located on the higher peaks, with large alpine grasslands found down to 2,000 m altitude. Occupation above 1,600 m shows adaptation to colder climates. What was the nature of occupation at these mainland New Guinea sites? The little evidence that is available suggests these early colonizers were “intermittent” hunters living a mobile way of life, perhaps making seasonal forays into the higher altitudes. Although larger prey would have been hunted, such as tree kangaroo, wallaby and bandicoot, hunters mostly caught the medium-sized to small prey such as possums, bats and frogs (see Mountain, 1991). There is also evidence that they may have hunted extinct marsupial megafauna as bones and teeth of Protemnodon (wallaby), Diprotodon and Thylacinus cynocephalus were found from Nombe and the later sites of Kafiavana and Kiowa. These large animals were montane forest and sub-alpine grassland browsers that may have attracted humans to these higher altitudes in the first place (see Mountain, 1991, for a description). Low population numbers are also suggested by Pasveer’s (2004) archeological research in the exploitation of wallaby (Dorcopsis Muelleri) in the Bird’s
12
Head archeological sites. On the basis of bone analyses she suggests that “caves were visited infrequently and that hunting in the vicinity of caves was sporadic and low intensity” (Pasveer, 2004: 338–9). These early inhabitants also altered the environment with clearing. Groube (1989) argued that the presence of waisted axes on the Huon indicated forest clearance, an argument extended to both Nombe and Kosipe. The presence of charcoal in pollen core data from across the highland area suggests forest clearance by fire at a number of sites: from the Baliem Valley in the west, to Telefomin near the current Indonesia–PNG border, and to Kosipe in the highlands of southeast Papua. Evidence from Kosipe suggests people could have been seasonally exploiting pandanus at 26,000 BP (White et al., 1970). In summary, the initial colonization of New Guinea and the Bismarck Archipelago occurred late in the Pleistocene primarily along the coastal routes then available. The spread was archeologically instantaneous and not entirely restricted to the coast, as evidenced by the occupation of Yombon, an interior rainforest site on New Britain. Within ten millennia or less, the interior of New Guinea was also being utilized. The expansion of peoples into different environmental niches and the speed of expansion is testimony to the “endurability” of these colonizing groups, their mobile way of life, and adaptability. It is reasonable to infer that continual moving across such expanses did not lend itself to the building of territories or group boundaries, and there was considerable interaction among groups as Gosden (1993) had suggested. However, there is no hard archeological evidence for the subsequent movement of people or goods between mainland New Guinea and Island Melanesia for another 20,000 years. Perhaps once the islands of New Britain and New Ireland were colonized they became isolated from mainland New Guinea to the west. A similar barrier could also have existed between the Bismarck Archipelago and North Solomons. Once Bougainville was colonized there is no evidence for any further interactions between it and the Bismarck Archipelago till only 3,300 years ago (see figure 2.2). Evidence to support this model of isolation can be seen in the recent work on Melanesian mitochondrial haplogroups P and Q by Friedlaender and his team (Friedlaender et al., 2005a; 2005b). They argue that the mitochondrial haplogroup Q2, which is not found on mainland New Guinea, developed within the Bismarck Archipelago with an estimated coalescence time of 36–37,000 years ago (a standard error of 11–12,000 years). In short, human populations (or at least the females as mtDNA evidence relates only to female isolation) of the Bismarck Archipelago remained in isolation from populations on mainland New Guinea.
Island Melanesian Pasts: A View from Archeology
Figure 2.2 Interaction barriers between mainland New Guinea and the Bismarck Archipelago following the initial colonization and up until 20,000 BP.
Origins There is no question that these early colonists originated from further west, but the particulars cannot be ascertained. We simply know much less about the archeological record of Southeast Asia than that of Papua New Guinea. There is no evidence of modern humans in Southeast Asia before about 45,000 years ago. Niah Cave provides the first evidence in contexts at least 43,000 to 42,000 BP, based on radiocarbon dating on charcoal fragments from just above the skull (Barker et al., 2001: 56). Dates of between 38,000 to 28,000 BP are also found in association with stone tools at Lang Rongrien cave from southern Thailand (Anderson, 1990). There are two proposed routes to enter Sahul (see figure 2.3). The first, a southern route, passes through Timor into either what is today the Sahul Shelf, or further north to what is today the Aru Islands. The second passes through a series of islands including Halmahera and Seram, ending up in West Papua (see Bellwood et al., 1998: 233). There is evidence for early human occupation following both these routes, although these settlements are younger than those found further east in New Guinea. Along the southern route, recent excavations at Lene Hara, East Timor, has occupation dated to between 30,000 to 35,000 BP (O’Connor et al., 2002: 45) and dates of 26,000 BP are from Liang Lemdubu in the Aru Islands
(Spriggs, 1998: 933, Veth et al., 1998). Liang Lemdubu would have been 45 km from the coast at the time of occupation. From the northern route, dates of 33,000 BP have also been obtained from east of Halmahera at Golo and Wetef Caves on Gebe Island (Bellwood, 1998: 958; Bellwood et al., 1998). This situation highlights the need for further work in Southeast Asia and also the problem that any earlier coastal occupation in this area would have been covered by the rising sea levels after the last glacial maximum.
From 20,000 Years Ago to the End of the Pleistocene While there is little archeological evidence to suggest major transformations in the nature of early societies in Near Oceania, this changes at 20,000 years ago when the barriers that had separated mainland New Guinea and the Bismarck Archipelago since initial colonization were broken. Novel animal species and goods were introduced from New Guinea across the Vitiaz Strait (still a difficult crossing today in a banana boat with an outboard motor). Seafaring abilities were clearly improved during this period as well because Manus was colonized from the north coast of New Guinea, which represents a significant ocean crossing.
13
the framework
Figure 2.3 Archeological sites from Southeast Asia with evidence of early humans.
The First Translocations
New Ireland
The long-distance movement of both animals and artifacts at 20,000 years ago has wider implications for the nature of society and groups occupying Near Oceania. The movement of animals from island to island was either the result of purposeful movements by humans or the accidental byproduct of their movement, i.e. as stowaways, a process recently defined as “ethnophoresy” (Heinsohn, 2003: 351; 2004a; 2004b). There are clear examples of purposeful introductions in both New Ireland and the Admiralties.
From New Ireland there is evidence for the introduction of new animal species and the beginning of the transportation of obsidian. In Matenbek (southern New Ireland) and Buang Merabak (central New Ireland) the cuscus, Phalanger orientalis, appeared at 23,500 to 20,000 BP and from Matenkupkum a little later at 16,000 BP (Flannery and White, 1991; Allen, 1996: 19; Leavesley and Allen, 1998: 72). These animals would have originated on mainland New Guinea and been physically transferred across to New Britain and then to New Ireland (see figure 2.4).
14
Island Melanesian Pasts: A View from Archeology
Figure 2.4 Late Pleistocene translocations and movement of materials.
Later translocations involved the introduction of Rattus praetor in the northern New Ireland site of Panakiwuk at 15,000 BP (Marshall and Allen, 1991). Why has evidence for animal translocations been found in New Ireland but not New Britain? The New Britain Pleistocene occupation sites, both coastal and interior, are dominated by volcanic activity, so that the soils retain little evidence of organics. However, the sites do contain stone artifacts, and some are from local obsidian sources in West New Britain. The distribution of the particular varieties of obsidian can yield important clues to the nature of exchange. At 20,000 BP, West New Britain obsidian already appears in archeological sites in southern and central New Ireland ~350 km away (Summerhayes and Allen, 1993; Leavesley and Allen, 1998: 72). The movement of obsidian follows a “down the line exchange” model where closeness to a particular obsidian source determines the amount of obsidian at a site, analogous to a genetic “isolation by distance” model. For instance, in the southern New Ireland site of Matenbek, obsidian from Mopir is much more common than obsidian from the more distant source area on the Willaumez Peninsula (see figure 2.4). The same pattern is repeated at Kupona na Dari in New Britain itself where the proportion of obsidian from source areas is directly proportional to distance to the source (see Table 4 in Torrence et al., 2004: 113). Unlike the southern New Ireland assemblages, however, obsidian from Kupona na Dari was probably obtained by direct procurement. Whether obsidian was exchanged as far west as New Guinea is not known because of the paucity of excavated sites there.
This type of exchange pattern has a bearing on the distribution of the Phalanger as well. Phalanger originated from mainland New Guinea, and it has been argued that they were brought to the Bismarck Archipelago as a breeding population (Allen and Gosden, 1996: 188) or as an accidental byproduct, being escaped pets or potential food (Allen et al., 1989: 557). Either way, they would have supplemented previous protein sources of reptiles, birds, bats, and seafood. Gosden (1993) argued that the introduction of Phalanger marks a change from people traveling to obtain resources, to people carrying resources with them. The importance of Phalanger is underlined in the southern and central New Ireland archeological sites, where it abruptly appears and then dominates the faunal assemblage. Why did the long-distance movement of goods began in this region, after the preceding 20,000 years of occupation? While Gosden suggested this represented a simple innovation in human management, it was more likely a byproduct of population increase. The nature of human society did not fundamentally change during the first 20,000 years since colonization. The evidence from the few archeological sequences available shows little change to subsistence patterns. What may have been a critical change is the gradual filling up of the landscape, which would have varied from the larger island of New Guinea that is rich in land animals, compared to the much smaller islands of the Bismarck Archipelago, which are depauperate in land animals. With population increase, group territories would be expected to slowly develop with defined boundaries,
15
the framework
the nature of interaction between groups of people was crossed resulting in the regular distribution of obsidian to settlements 350 km to the east, and the importation of animals from the west.
The Admiralties
Figure 2.5 Changing hunting and gathering territorial ranges in New Britain.
especially as different groups came into more frequent contact (figure 2.5). Evidence for long-distance downthe-line exchange of animals and obsidian appeared at 20,000 years ago because that was the first time a chain of regular and frequent contacts could have developed between mobile communities. That is, widespread resource distribution requires the development of dependable exchange links with other communities. Allen (1996: 21) saw evidence of this relationship between population increase, territoriality, and resource distribution in north and south New Ireland. He argued that as more Bismarck sites were settled, “relationships between territories around the archipelago presumably also evolved structurally.” Torrence argued that by the early–middle Holocene, “mobile groups moved shorter distances and exploited smaller areas than previously” (Torrence et al., 2004: 126). However, this reduction in the size of foraging zones should logically have begun earlier at c. 20,000 BP, along with the other changes associated with population increase at that time. It was at this time that a threshold in
16
While these changing patterns of interactions were occurring between New Britain and mainland New Guinea and New Ireland, Manus was settled for the first time. As mentioned, the colonization of Manus implies some form of sophisticated water transport, as it involves a substantial water crossing of either 230 km from the north coast of New Guinea, 200 km from Mussau, or 230 km from New Hanover/Lavongai (Irwin, 1992: 21). Excavations at Pamwak Cave produced a radiocarbon AMS (Accelerator Mass Spectroscopy) age of 21,000 BP for human occupation at a depth of 3 m. (Ambrose, 2002: 68). The initial occupation could have been much earlier, as another 80 cm of human deposit lies below the point where the date was collected (Ambrose, 2002: 68). This crossing was not an isolated event. There is a gap of several thousand years before occupation levels again appear at Pamwak, dated to 13,000 BP, with new deliberate introductions. They contain the remains of mainland New Guinea animals and nut trees; cuscus (Spilocuscus kraemeri), bandicoot (Echymipera kalubu), and Canarium indicum nuts (Kennedy, 2002: 20; Summerhayes, 2003a: 139; Specht, 2005: 252–9). Obsidian extracted from small island sources 25 km away also appear (Summerhayes, 2003a). Another phase is indicated soon afterwards at Pamwak between 12,000 to 10,000 BP, with the probably unintended introduction of Rattus praetor. As mentioned, evidence for this animal occurs earlier in New Ireland, at 15,000 BP at Panakiwuk (White et al., 2000: 106; Specht, 2005: 256). Once the Admiralties were colonized they did not remain in complete isolation, since many goods continued to be transported or introduced over long sea distances. It is unclear if these interactions were reciprocal, as, for example, there is no evidence of obsidian being transferred from the Admiralty sources where it was then in use, to the New Guinea north coast. However, this may again be because of the lack of archeological research there. The established contacts during this period with the Admiralties were only with the north coast of New Guinea and not with islands in the Bismarck Archipelago. Of the species mentioned above that were identified in both New Guinea and Manus, the bandicoot Echymipera kalubu was never identified in New Ireland assemblages, although it did appear later in New Britain. Jim Specht (2005: 258) and others (Allen, 2000: 155; Spriggs, 2000: 297) interpret its absence in New Ireland and presence in
Island Melanesian Pasts: A View from Archeology
Manus as its being a direct introduction from New Guinea. Also, the cuscus Spilocuscus kraemeri is absent in prehistoric assemblages of New Ireland and New Britain but is found today in the Admiralties and some of the western Admiralty Islands (Flannery, 1995: 104–5). So the evidence points to a lack of interaction, or some cultural/social/ economic boundary, between the Admiralties and New Ireland. The first evidence of any interaction between these two areas occurred in the mid-Holocene when obsidian was transferred from the Admiralties to the island chains in the western Pacific in association with Lapita (see below).
Bougainville and Buka Another social/economic/migration barrier also existed between New Ireland and Bougainville/Buka. Although Bougainville/Buka was colonized by 32,000 years ago, there is no evidence of subsequent interaction with New Ireland of any kind until 29,000 years later, when obsidian associated with Lapita settlements are found in both regions. Also, while Phalanger orientalis was found in New Ireland at 23,000 BP, it does not appear in the Bougainville region until Lapita levels along with the introduced wallaby, although Phalanger is found on Nissan at 5,000 BP (Wickler, 2003: 235). The Bougainville region was clearly isolated from the Bismarcks for most of the time since its initial colonization. The only question is the completeness of the isolation during that period.
Secondary Movement of Animals and Plants There are other early indications of human-mediated transfer from New Guinea to Island Melanesia, particularly of nut trees and edible plants. While most of the associated dates are younger than 20,000 BP, the possibility for their transfer at the times of first colonization is high, given the indications of isolation just discussed. The secondary introductions of uncertain initial date include nut trees native to New Guinea. The homeland of the almond Canarium is the north coast of New Guinea according to the archeobotanist Doug Yen (1990; 1995: 838) since wild forms are endemic there. The transfer of Canarium required considerable human intervention as it “must have been deliberately planted, tended, and harvested” (Spriggs, 2000: 297). The earliest evidence for Canarium indicum is found on mainland New Guinea at Kowekau Cave (known as Seraba) in contexts dated to about 14,000 BP (Yen, 1990: 262). It appears somewhat later at Pamwak in Manus in levels dated to 13,000 BP, by 8,000 B.P at Panakiwuk, northern New Ireland (Marshall and Allen, 1991), and further south at Kilu Cave in Buka at 10,000 BP (both Canarium indicum and C. solomonense and also some questionable identification of Canarium in
a lower Pleistocene layer). Another nut, Cocus nucifera, is found at a number of sites in Island Melanesia prior to 5,000 BP. It is found with Canarium throughout the Kilu archeological sequence in Buka (Wickler, 2003: Table 8:11), in New Ireland (Panakiwuk) from 10,000 BP (Marshall and Allen, 1991: 87) and at Lebang Takoroi from Nissan at 6,000 BP (Spriggs, 1991: 237). More problematic examples of early human-mediated transfers involve root crops, which could have been naturally distributed throughout the region (Spriggs, 1997). Colocasia and Alocasia taro starch residues and crystalline raphides were identified from stone tools in contexts from the first occupation of Kilu at 28,000 BP (Loy et al., 1992). Evidence for plant processing in the late Pleistocene comes from the New Ireland cave site of Balof where residue analysis on stone tools suggests the processing of both Cyrtosperma merkusii and Alocasia macrorrhiza, as well as yam (Dioscorea bulbifera), dated to 14,000 BP, and 9,400 BP (Barton and White, 1993: 175 and Table 1). Hay (1990) argues for their dispersal by humans.
Summary The dispersal of artifacts, animals and plants in the Bismarck and Bougainville archeological record beginning at 20,000 years ago indicated what Jim Specht (2005: 271) calls the “broadening of social horizons,” or what Peter White (2004: 157) suggested typically happened as “small-scale societies have become incorporated into larger state or world systems and there is some market demand for exotics.” This change in the nature of societal interactions and the movement of goods was driven by an increase in population levels that caused the formation of group boundaries and territories. The societies would however remain relatively small in scale. The colonization of Manus and the transfer of materials over a 200-km water distance is also testimony to the sea-faring skills of these Pleistocene travelers. It is a pity that the volcanic nature of the New Britain landscape does not allow the preservation of organics that would help complete the picture of regional island transfers.
Early to Mid-Holocene Transfers This is an important and controversial time interval in Near Oceania. Agriculture develops during this period, prior to the appearance of the Lapita Cultural Complex in some coastal regions, using plants from the coastal areas. There is an increase in the types of materials exchanged across large distances. Introductions from outside the region include pottery and pigs. All these developments have important implications for modeling the nature of
17
the framework
human interactions at this time. However, some archeologists (e.g. Spriggs, 1997) have argued that these items were all transported to Near Oceania later by Austronesianspeaking groups as part of the Lapita Cultural Complex (see below). My emphasis here is on the critical role of climate change in enabling these changes early on. The major warming trend underway by the beginning of the Holocene brought about major ecological changes that altered the possibilities for movements of people and goods. Vegetation patterns changed, with forest cover replacing alpine grasses. The tree line gradually increased from 2,000 m at the height of the last maximum glacial to 4,000 m above sea level today. This must have been salubrious for humans, and highlands sites in New Guinea indicate just this. At Nombe, dramatic increases in stone tools, burnt bone and other midden material indicate an increased human presence (Mountain, 1991: 517). The indicated intensified consumption of bats also suggests more permanent human camps. At the highlands site of Manim there is evidence for Pandanus processing (Christensen, 1975). These increases in site use intensification suggest the beginning of territories and settled landscapes and are also associated with first, the beginning of agriculture in the highlands and, second, evidence for the long-distance movement of materials. In contrast, in the lowlands during this period, major vegetation changes were mostly human induced (burning) with anthropogenic grasslands or open savannah found in many areas such as the Markham Valley, the Sepik-Wewak areas, Oriomo and Moresby environs (see Hope et al., 1983: 41). The warming after the last glacial also had a dramatic effect on the shape of coastlines in many locations, and caused the separation of New Guinea from Australia by 8,000 years ago. Along the north coast of New Guinea, the effect on the coastlines varied considerably. At the western end of northern Papua New Guinea near Lachitu rock shelter, the deep submarine trench off the coast meant that any decrease or increase in sea level would not greatly change the coastline. On the other hand, today’s Sepik and Ramu River drainage areas would have been a large inland sea up until 6,000 years ago, when the shoreline began to recede (Swadling et al., 1989; 1991; Swadling, 1997).
Changes in the Highlands Prior to 10,000 years ago, agriculture was not even a possibility in the highlands (Golson, 1991; 1997). The paleoecologist Simon Haberle outlines the reasoning: “The occurrence of infrequent but severe drought and associated frost particularly between 20,000 and 11,500 cal yr BP in highland New Guinea, would have put sustained production of most food plants out of the question”
18
(Haberle and David, 2004: 173). The warmer conditions of the early Holocene allowed the movement of these crops from the tropical lowlands to the highland region, whether carried by trade or with population movements. Golson (1991: 487) argued that whatever crops were used were first cultivated in lower altitudes. At these lower and mid-altitudes many indigenous vegetables, nut and fruit species were exploited in a hunter-gatherer economy leading to domestication (Yen, 1982: 292). Thirty years ago, it was assumed that most of the domesticates (yams, taro and bananas) were of Southeast Asian origin (Powell, 1977: 16). It is now clear that Australimusa banana, sugar cane (Saccharum officinarum), other canes, tubers (Pueraria lobata), and breadfruit (Artocarpus alitilis) were domesticated in New Guinea. Some aroids were independently domesticated in New Guinea and in Asia, notably taro (Colocasia esculenta, and Cyrtosperma chamissonis) (Yen, 1995: 835–40) and the Eumusa banana (Kennedy and Clarke, 2004). However, Yen (1995: 837) believes Alocasia macrorrhiza is an introduction as an agricultural crop, even though there are wild varieties in New Guinea. The situation with yams (Dioscorea) is unknown, with some having wild forms (Yen, 1995: 836). Denham et al. (2003) have recently confirmed that taro and banana were cultivated at Kuk, consistent with Yen and Golson’s early work. The major evidence for the development of agriculture in the highlands comes from the long record from the Kuk Tea Plantation excavations just outside Mt. Hagen in the upper Wahgi Valley, at 1,650 m above sea level. The major features of the Kuk excavation are prehistoric water control channels. These ditches were placed into a chronological sequence using the volcanic ashes found in the drainage fill. Golson defined six phases at Kuk. The top five phases lay above a gray clay layer while the oldest lies below it. Phase One dates from 9,000 BP and is made up of “gutters, hollows and stake holes.” It provides the earliest inference of agriculture in New Guinea. This is based on geomorphologic (erosion) and structural (water control measures) evidence. The second phase dates from 6,000 to 5,000 BP and overlies the gray clay, suggesting major forest clearance related to agricultural intensification. This phase has evidence of structures and channels in swampland gardens that suggest taro cultivation. Raised beds for other crops such as bananas are also present. Phase Three lasts from 4,000 to 2,500 BP and consists of a network of channels to drain water from agricultural areas. This is followed in Phase Four by a grid-like system of elaborate field ditches that drained a larger area. The last two phases (Five and Six) date from the last 400 years and had extensive drainage systems indicating sweet potato cultivation (see Golson and Gardner, 1990; Golson, 1991; Hope and Golson, 1995).The earliest evidence for major forest clearance elsewhere is in the
Island Melanesian Pasts: A View from Archeology
Baliem Valley, over 2,000 years later than at Kuk (Haberle and David, 2004: 175). By 6,000 BP there is evidence for agricultural intensification (Golson’s Phase Two at Kuk), with an increase in highland occupation sites and more evidence for vegetation change in swamp environments. This may indicate population increase and the beginning of territories within the highlands. The development of agriculture brought with it changes in settlement and subsistence strategies, and this would have had a major effect on the nature of societal interactions within this region.
Exchange within New Guinea and with the Bismarck Archipelago Indications for interactions solely within the highlands are scant during this period. Exchange is suggested by two “ground chips” of axes found in Kafiavana from 5,000-year-old levels that were sourced to the Kafetu quarry 50 km away (White, 1972: 95). Imported axe fragments were also reported from Yuku in levels dated to 4,500 BP (Bulmer, 1975: 31). Nevertheless, Golson and Gardner (1990: 404) argued that early axe production was mostly locally distributed until 2,500 to 1,500 years ago when there is the first evidence for the “development of complex exchange economy.” Interactions between the highlands and coastal areas are a little more numerous. Besides the movement of cultigens into the highlands, coastal–highland exchange systems apparently commenced involving the longdistance movement of shell and stone (figure 2.6). Marine shell (cowrie) dated to about 8,000 BP was found at the highland site of Kafiavana (White, 1972), and there is evidence of marine shell found at levels at Yuku dated to 4,500 BP (Bulmer, 1975: 30).
Early exchange between New Guinea and the Bismarck Archipelago is indicated by the wide distribution of similar pestles and mortars, particularly birdshaped pestles (see Swadling and Hide, 2005, for details). The ages for these stone pestles and mortars in the highland region range from between 8,000 to 3,750 years ago (significantly, none have been found in the later Lapita assemblages—see below). Swadling also makes a convincing argument that these pestles and mortars are associated with taro production. Their wide distribution suggests exchange networks operating across mainland New Guinea and the Bismarck Archipelago not seen in later periods. At this time, obsidian from New Britain was making its way across the Vitiaz Strait to the New Guinea mainland and then up into the highland region. Obsidian has no natural source on mainland New Guinea, occurring only in West New Britain, the Admiralties and Fergusson Island (see Summerhayes et al., 1998). Obsidian was found at Kafiavana in the eastern highlands in contexts dated to 4500 BP (White, 1972). Swadling reports the finding of four stemmed obsidian tools from the SepikRamu region along the north coast, three of which were analyzed and sourced to the Kutau/Bao area of West New Britain (Swadling and Hide, 2005: 307). These distinctive forms are only found in New Britain from about 6,000 to 3,500 years ago and were produced using obsidian from a number of local sub-sources (see Summerhayes et al., 1998). The distribution of this tool type across the north coast of New Britain appears to indicate a single or socially related set of groups of mobile hunter-gatherers who extracted and used obsidian from a particular source they happened to be near (Torrence and Summerhayes, 1997). Their territories ranged from the major source
Figure 2.6 Early to Mid Holocene movement of materials.
19
the framework
areas of Willaumez Peninsula to Mopir 60 km away. The stemmed tools found from the Mopir region were made using Mopir obsidian, while identically shaped tools found near the Willaumez Peninsula were made using those sources. The stemmed tools from the interior and south coast of New Britain were not made from obsidian, but from chert, the local suitable rocks. The time frame is the same. Stemmed chert tools appear at Yombon about 6,000 years ago, although chert had been exploited for stone tool use (expedient flakes), since the site was first occupied from 35,000 years ago (Pavlides, 1993; Pavlides and Gosden, 1994). What were these stemmed tools used for? Torrence et al. (2000: 235–7) and Fullagar (1993) argue that these tools were used for plant processing and had to be prepared in advance and carried around until groups visited the source again. A recent study has argued that stemmed tools like those transported to New Guinea may have had ceremonial functions as well (Araho et al., 2002). Whatever their function, the stemmed tools of New Britain disappeared under a blanket of ash produced by the massive eruption 3,600 years ago of Mount Witori which is located adjacent to the Mopir obsidian source. The eruption devastated populations in this entire region, with Yombon resettled 800 years afterwards, and the Willaumez Peninsula 250 years later (Torrence et al., 2000). Animal translocations between New Guinea and the Bismarck Archipelago also occurred during this period. The small terrestrial wallaby, Thylogale browni, has been found in mid-Holocene New Ireland archeological sites such as Balof 2 at 8,400 BP (White et al., 1991: 57), Buang Merabak unit two at 6,200 BP (Leavesley and Allen, 1998: 75, 78), and Panakiwuk at 2,000 BP (Marshall and Allen, 1991). It does not reach the North Solomons until 3,500 years ago. Of course, the movement of obsidian, pestles and mortars, and animals across the Vitiaz Straits does not have to equate with the movements of people. For example, the distribution of Trans New Guinea Phylum communities (TNGP) does not extend east of New Guinea. Although the initial spread of pestles and mortars may be linked to the initial dispersal of TNGP speakers from the central highlands around 10,000 years ago (see chapter 3), pestles and mortars clearly spread beyond the TNGP region (a point made by Pawley in chapter 3), and they could have been a special exchange item for taro processing (Swadling, 2005: 3).
Pigs and Pottery The pig was introduced into New Guinea from Asia and its presence in Near Oceanic archeological contexts is an important indicator of a direct introduction across the Wallace Line. Since it is normally a domesticate, it is also
20
an indicator of an agricultural and semi-sedentary society. Golson cited the early evidence for pig in early New Guinea contexts as indicative of agriculture (see Golson, 1991). Pig bone has been found in three highland sites in contexts dated to the early Holocene: Kafiavana, Yuku and Kiowa (see figure 2.6). At Kafiavana, pig bones were dated to 6,500 BP (White, 1972). A single incisor was found at Yuku in levels dated between 10,000 to 12,000 BP, and another in later levels at 4,500 BP. Another incisor comes from Kiowa at 10,350 BP (Bulmer, 1975; 1982: 187). Gorecki (et al., 1991: 121) recovered pig bone from all levels of Taora shelter that dates at its earliest levels to 5,600 BP. Swadling also found pig remains at Akari in the Lower Ramu, in levels that could be mid-Holocene (Swadling et al., 1991: 106). Pig bone has also been found in the Bird’s Head of West Papua and directly dated to 1,840 BP, with an upper range of 4,000 BP (Pasveer, 2004: 34, 337). Two pig teeth have also been found from Matenbek, New Ireland, in contexts dated to between 8,000 and 6,000 BP (Allen, 2000: 157). These associations of pig bone in Near Oceanic contexts as old as 10,000 years ago have been criticized because no pig bones were found in Australia, which would have been joined to New Guinea up until 8,000 years ago (White and O’Connell, 1982: 189). Also, others have dismissed associations of pig bone with any Near Oceanic context older than 3,300 years ago, arguing that pigs were introduced by Austronesian speakers associated with the Lapita cultural complex (Spriggs, 1997). Certainly there is a strong association between pigs and Austronesian language terminology. For instance, Wurm argued that the highland word for pig “is a reflex of the proto-Oceanic Austronesian term for pig” and suggested it was related to Austronesian appearances in coastal areas (cited in Hope et al., 1983). AMS dating of a number of the pig remains does support its later introduction. Pig teeth from New Guinea sites including Kafiavana were dated and argued to be less than 500 years in age (Hedges et al., 1995). Furthermore, a pig tooth found in levels dated between 8,000 and 6,000 BP from Matenbek was AMS dated and produced a modern result (Allen, 2000: 158). Despite these modern results, Allen (2000: 159) makes a strong case for the presence of pig bones before the appearance of Lapita on both archeological and stratigraphic grounds. Debate also rages about the introduction of pottery into New Guinea. It has been argued that the first pottery was brought along with pigs into the area by Lapita peoples (see below). Earlier pottery, however, has been excavated from five mainland New Guinea mid-Holocene sites (see figure 2.6) and argued by Gorecki (1992: 42) to have originated from Asia, predating Lapita pottery by some 2,000 years (Gorecki, 1992: 42). Specifically, coiled pottery dated to between 5,500 to 3,000 BP was excavated
Island Melanesian Pasts: A View from Archeology
from Wanelek (Bulmer, 1977: 68; 1982: 180). Wanelek is an open settlement site in the Kaironk Valley, Madang Province, at 1650 m above sea level. Bulmer also found red slipped pottery in pre-3,000 BP contexts that she associates with Austronesians and settled agriculture (Bulmer, 1977: 68). Bulmer sees support for Austronesian influences in the stone tools found at Wanelek in contexts dated from 4,450 to 2,800 BP such as drill points and blade production. Bulmer sees these as similar to those found in Lapita and Polynesian assemblages (Bulmer, 1991: 470), and says they would go unnoticed in a Lapita tool kit (Bulmer, 1991: 476). Bulmer explains the presence of the “Austronesian” type artifacts as the result of trade. She argues that there may have been a settlement of traders from the lowlands camping here to make stone tools and get salt (Bulmer, 1991: 476). Swadling also suggests that the pottery was probably a trade ware brought up from the Ramu area to the north (Swadling, 1990: 76). A problem for this scenario is the dating of “Austronesian” settlements (in the form of Lapita) much later in time. That is, Wanelek Horizon D began earlier than dates for Lapita sites (Bulmer, 1991: 476). Two additional pre-Lapita pottery-yielding midHolocene sites are Taora and Lachitu, located near Fichin on the north coast of New Guinea, west of Vanimo. These sites provided evidence for plain ware pottery dated to 5,400 B.P (Gorecki et al., 1991: 120; Gorecki, 1992: 35). Gorecki names the pottery the “Fichin Tradition,” and describes it as thin-walled ware made using the paddle and anvil. Only one sherd showed incision (Gorecki, 1992: 35). Finally, Swadling found early pottery with incision and lip notching at the Akari and Beri sites in the Lower Ramu, dated from 5,600 BP (Swadling et al., 1989: 108–9; 1991). There are simply too many archeological sites containing pottery and/or pig bone in contexts before 3,300 years to be ignored and passed off as accidental or problematic. They provide strong evidence that interactions with the west existed before the introduction of the full-blown Lapita Cultural Complex. One last point. Evidence for interactions with Asia at this time was also argued on the presence of betel nut (Areca catechu) at the waterlogged site of Dongan, a site in the Sepik catchment, excavated by Pam Swadling (1990) and dated to the mid-Holocene. Betel would have been a direct introduction from Asia. However AMS dating directly on the remains of Areca catechu from the site returned a modern age estimate (Fairbairn and Swadling, 2005).
Summary Allen and Gosden (1996: 193) argue that by 4,000 years ago there was “continuous two way interaction between Island Southeast Asia and areas to the east.” An early
introduction of pig and other cultigens such as yams would be prime indicators of this. There are a number of examples of animals being transported in the other direction as well, from New Guinea to Island Southeast Asia. Phalanger orientalis, a native of New Guinea, is found in the site of Uai Bobo Two in East Timor in levels dated to 6,000 years ago (Glover, 1986: Table 122; Flannery, 1995: 96). Also wallaby (Dorcopsis) and bandicoot (perhaps part of genus Echymipera), natives of New Guinea (Flannery, 1995), were found at Siti Nafisah cave, Halmahera, in contexts from 5,500 to 3,000 BP, and Dorcopsis only on Gebe Island. From Gebe, wallaby (Dorcopsis) appears in the early Holocene at 8,000 BP (Bellwood et al., 1998: 261). It is also found on Misool and Japen, islands off the western part of New Guinea (Flannery, 1995: 79–80). Taken together, the evidence suggests a number of species translocations and the movement and exchange of materials within Papua New Guinea and the Bismarck Archipelago, and between these regions and Island Southeast Asia.
Late Holocene: Lapita and Its Spread into the Pacific Three thousand years ago there clearly was a major movement of people and animals and a transformation of landscapes on a scale not seen before in this region. This was the colonization for the first time of the area known as Remote Oceania, being made up of the islands to the east of the Solomon Island chain including Vanuatu, New Caledonia, Fiji, Tonga and Samoa (figure 2.7). The archeological culture associated with this colonization movement is called Lapita, with its signature being a highly decorated dentate-stamped pottery (figure 2.8). Nearly 200 Lapita sites across the western Pacific have now been reported, with 80 sites in the Bismarck Archipelago alone (Anderson et al., 2001). From these archeological ceramic assemblages separated by over 3,000 km, more than 500 motifs, mostly dentate-stamped, have been identified (Anson, 1983). It was once thought that this colonization process was archeologically instantaneous (Kirch and Hunt, 1988). However, with more sites and better dating techniques, we now know that this was not the case, and that the Lapita colonization took half a millennium. The chronology of sites demonstrates a west to east gradient, with the Bismarck Archipelago occupied first by 3,300 years ago, the southeast Solomons and Vanuatu by 3,100 years ago, Fiji by 2,900 years ago, and Tonga and Samoa by at least 2,800 years ago (figure 2.7). The motifs and Lapita vessel types also changed over time, and since these changes occurred across the western Pacific, the sites can be placed into a chronological framework: Early Lapita, Middle Lapita and Late Lapita (see Summerhayes, 2000b; 2001b).
21
the framework
Figure 2.7 Lapita’s spread through the western Pacific.
Archeologists have primarily focused attention on the origins of Lapita societies as well as the nature of their interactions and colonization, based on the distribution of pottery and obsidian, and the nature of settlement subsistence.
Origins of Lapita Society There are three current models commonly used to account for Lapita’s origins (see Green, 2003, for details).
The Fast Train This origin model invokes a movement of Austronesianspeaking people out of Southeast Asia and into Remote Oceania, passing through the Bismarck Archipelago, carrying with them their complete material cultural repertoire (the Lapita Cultural Complex). Based on early
22
radiocarbon estimates, the initial occupation of the Bismarck Archipelago was seen as predating occupation in Fiji, 3,000 km to the east, by a century at the most (Kirch et al., 1987). This suggested the spread of Lapita colonization was archeologically instantaneous (see Kirch and Hunt, 1988), thereby accounting for the similarity in material culture over a vast region. This model considered the domestication of animals, the Austronesian languages of the Pacific, and many elements of the material cultural kit to be derived from Southeast Asia. In this model, any subsequent change in the Lapita pottery style was due to subsequent isolation of populations.
Indigenous Bismarck Archipelago Model The second model posits the development of the Lapita Cultural Complex as occurring within the Bismarck
Island Melanesian Pasts: A View from Archeology
Figure 2.8 Lapita dentate-stamped pottery (from the site of Kamgot). See color insert.
Archipelago from indigenous sources. That is, the colonization of Remote Oceania still has its origins in the Bismarck Archipelago, but the Lapita Cultural Complex arises as a result of local social and economic developments of the previous 35,000 or so years (White and Allen, 1980; Allen, 1984). Despite this indigenous emphasis, it is assumed people did not live in a vacuum and could have had contacts with the west from which they may have acquired the skills to make pottery, seen in this model as “culturally unaccompanied baggage” (Kennedy, 1983: 120). As Allen noted (1991: 7) “such contacts would have facilitated the flow of materials, technologies and people in both directions.”
The Slow Train The third model, although seeing the ultimate origin of the Lapita Cultural Complex in Southeast Asia (as in model 1), sees Lapita as evolving in the Bismarck Archipelago for 300 years before spreading out into Remote Oceania. Roger Green (1991) has developed a variant of this model that he calls the Triple I model: Intrusion/Innovation/ Integration. Intrusion equates with Austronesian speakers coming into the area from Southeast Asia, bringing with them items of material culture. Innovation equates with new developments within the Bismarck Archipelago, while integration equates with adopting elements of material culture from the area’s original inhabitants.
Thus people may have paused in the Bismarck Archipelago and indeed picked up local elements of material culture on the way (Green, 1991), perhaps learning to adapt “to an area with a complex continental island environment, which possessed a wide range of resources” (Green, 1979b: 45)—a kind of “homeland” (see also Spriggs, 1989: 608; 1996). Both the second and third models envision the Lapita Cultural Complex developing in the Bismarck Archipelago before colonizing groups left the area for Remote Oceania. The current archeological evidence lends itself to the third model. The inhabitants of the Lapita settlements would have spoken a Proto-Oceanic language and shared a similar culture. There is no evidence in the archeological record that suggests that Lapita social forms, settlement, or domesticated economy evolved directly from variants within the Bismarck Archipelago or New Guinea over the preceding 35,000 years of human settlement. While there are indications of continuity in some forms of material culture and subsistence, we must look elsewhere for structures for change, i.e. the kick-start that got the momentum for colonization possible. This is where we hit the major snag. Since the archeology of Island Southeast Asia is poorly known, many questions concerning the influences of that region on the Southwest Pacific will remain open to debate. We do know, however, that these Lapita peoples resided in the Bismarck Archipelago for 300 years before colonizing islands of
23
the framework
Remote Oceania, and as Green argues (1991), during that time they acquired elements of local material culture from interactions with the non-Austronesian, or Papuan, inhabitants, while also creating new ones as well. It does appear that these populations created the “Lapita character” in the Bismarck Archipelago during this short interval. The particular nature of the interactions with non-Austronesians is not known, primarily because of our poor archeological knowledge of occupation on contemporary non-Lapita sites. However, intermixing with the local populations who had inhabited this area for the previous 35,000 years would not be unexpected (Hayden, 1983), making Lapita a Northern Island Melanesian phenomenon.
Nature of Interaction between Lapita Communities: Pottery The assemblages from these early colonizing communities are very homogeneous and change in a regular fashion over time, indicating continuing interaction among the different Lapita settlements. Lapita is known mainly for its highly complex form of dentate-stamped ceramic decoration that was previously unknown anywhere else. The oldest Lapita assemblages that are found in the Bismarck Archipelago have the most complex decoration,
and motifs become less complex over time. Yet Lapita assemblages are made up of considerably more than a specialized dentate-stamped pottery, as shown in assemblages from the Arawe Islands. Within Lapita assemblages, there are two sets of vessels (dentate versus non-dentate), each having a different tempo or rate of change over time (Summerhayes, 2000c). The less ornate non-dentate ware changed little over Lapita’s duration. This included the plain globular pots with everted rims, plain and decorated carinated jars, and plain bowls, some with red slip, which on the basis of shape could have been used either for cooking or as containers for storing water (see also Kirch, 1997: 122). The slow rate of change in these vessels is likely related to their ongoing domestic/utilitarian role. On the other hand, the dentate ware changed dramatically over time, not only in its representation within the assemblage, but also in shape, motif type, and production (see Summerhayes, 2000a, for details). Its decline over time has to be related to its lessening importance. These changes are clear in the major Lapita assemblages of Mussau, the Arawe Islands, and Anir (figure 2.9), located at the furthest extremes of the Bismarck Archipelago, suggesting similar roles within these three assemblages. The production of pottery in these assemblages was mostly local. These three communities kept
Figure 2.9 Lapita sites in the Bismarck Archipelago.
24
Island Melanesian Pasts: A View from Archeology
up some level of communication over many generations, but this did not have to involve complex exchange networks or even frequent contacts. For example, the same new dentate motifs, which cannot be derived directly from the original corpus, appear in all three assemblages in the same late time interval. The same sorts of contemporaneous changes in Lapita styles occur in other parts of the southwestern Pacific, including the Santa Cruz and the Reef Islands, Fiji, and Tonga (Summerhayes, 2000a). Dentate-stamping itself is informative on the nature of Lapita society. The importance of these motifs can be seen in their systematic spatial distribution (Summerhayes, 2000c). The listing of over 500 dentate-stamped motifs by Anson (1983) and their arrangement in structured patterns from the Bismarcks to Samoa on these specialized pots suggest that they were not meaningless scribbles. Motif similarities and change across such a wide expanse were not the result of exchange from a small number of production centers. They are consistent with Rathje’s (1978) contention that these motifs were active social/ideological signifiers that conveyed information, foster group identity, and maintained social boundaries (see Hill, 1985: 367). Conversely, the decline in the dentate-stamped component and its eventual loss in the area from the Bismarck archipelago to Samoa not only indicates the lessening of the role of ceramic decoration within that society in maintaining social paths, but also reflects changes in the social formations and the nature of interaction between these communities. What the actual motifs meant is difficult to assess. The finding of a modeled clay head with dentate markings on its nose and an impressed circle on its forehead from the Lapita site of Kamgot, Anir (Summerhayes, 1998), and similar finds from Boduna Island (Torrence and White 2001) strengthens the argument that dentate-stamping originated in tattooing (see Green 1979b: 16; Kirch 1997: 142). Green had previously found a human figurine with dentate-stamped motifs on the buttocks that he suggested represented tattooing (Green, 1979a: 16). Spriggs (1990: 110) has noted that the motif found on the buttocks is similar to the “sun motif” earplug emblems found on sherds from Watom, Reef/Santa Cruz and Tonga. He suggests that the earplug emblem might represent “group (clan?) affiliation” (Spriggs, 1990: 119). Within many Island Melanesian societies, each motif used forfacial tattooing was a social signifier such as a clan mark (see Parkinson, 1907; Fox, 1925: 16, 296, 348; Ivens, 1927: 82–6). Whether dentate-stamped motifs represented clan markings in the Lapita period cannot be known for certain, but all the evidence indicates their significance as social markings/signifiers of some sort from their first appearances in the Bismarck Archipelago.
There is now general agreement that the initial colonizers of Remote Oceania came from communities in the Bismarck Archipelago, spoke a Proto-Oceanic language, and made Lapita pottery (Summerhayes, 2001a; Pawley, 2003). One last intriguing point remains concerning the origins of Lapita. Pottery of a similar nature, but without the dentate-stamping, is found in Island Southeast Asian sites in assemblages contemporary with the earliest Lapita sites in the Bismarck Archipelago, but not earlier. Red slipped pottery with identical shapes to the plain ware component of the Lapita assemblages is found in Uattamdi shelter on Kayoa Island, west of Halmahera (Bellwood, 1998: 960). These sherds were found in association with shell ornaments, polished stone adzes, domesticated pig, and dog, similar to what is found in Lapita sites.
Nature of Interaction between Lapita Communities: Obsidian Although obsidian had been exploited for some 17,000 years previously in the Bismarck Archipelago, during the Lapita period the extent of its distribution and the nature of its extraction and technology were unlike anything detected previously. Obsidian from the Bismarck Archipelago is found in Lapita assemblages as far to the east as Fiji, and as far west as Sabah (Bukit Tengkorak) dated to 3,000 BP (Bellwood, personal communication), making a total distribution extent of 6,500 km (Summerhayes, 2003a; 2003b; 2004). Second, the nature of obsidian selection changes. Prior to the Lapita period, obsidian was selected and distributed from a number of obsidian flows: Kutau, Baki, Gulu and Mopir, which was 60 km away, next to Mt Witori. After the eruption of Mt Witori at 3600 BP, New Britain obsidian was exclusively extracted from the Kutau source for export out of the region. Mopir would have been covered by the eruption and made inaccessible. However, both Gulu and Baki would have been just as accessible as Kutau, yet they were not exploited for external export. Locally, Kutau dominated in all the Talasea region assemblages. Obsidian from the Admiralty Islands was also exported for the first time out of its immediate region and out into the Pacific with Lapita assemblages. Third, the obsidian technology changes. Prior to the Witori eruption, stemmed tool production was predominant. Yet within Lapita assemblages, the stemmed tools disappear, retouched artifacts decrease, systematic quarrying ceased, and flakes were expediently flaked with no systematic planning from prepared cores (Torrence et al., 2000). Richard Fullagar (1992) has demonstrated that obsidian tools produced in Lapita assemblages had evidence of being used for more activities than obsidian
25
the framework
from earlier pre-Lapita assemblages. He argued that this equates with more long-term residence. In fact, Torrence found that Lapita sites were repeatedly occupied, unlike earlier sites, indicating a more intensive system of land use. Thus the fundamental nature of obsidian extraction and distribution changed with Lapita settlements, making a discontinuity with what went before. The distribution of obsidian as far as Sabah, in addition to the Lapita-like non-dentate pottery assemblages in that region, points to an even greater degree of interaction in a wider universe than many would accept.
Lapita Site Location and Economy The settlement locations of Early Lapita sites within the Bismarck Archipelago also marked a clear shift from what had gone before. Early Lapita settlements were coastal, on or near beaches. Excavations at Apalo and Adwe in the Arawe Islands (Gosden and Webb, 1994) and Talepakemalai on Eloaua in the Mussau Islands (Kirch, 1988, 2001) also showed that some of these settlements consisted of stilt houses over the reef platform. No beach settlement has been identified for any period of the previous 37,000 years of occupation. PreLapita sites were either open or cave/rock shelters away from the beach. The coastal location of Lapita settlement was reflected in the economy. From the Lapita site of Kamgot for instance, fish bone dominates the faunal assemblage. The most popular fish were inshore varieties such as Scaridae and Diontidae. Most of the fishing was inshore or from the reef, although shark, tuna, dolphin, turtle, and barracuda were also found. Yet the midden remains also have high land mammal content as well, including Phalanger, Thylogale browni, pig, chicken, dog, and Rattus exulans. The last four species are Southeast Asian introductions and apart from pig, have not been found in mainland New Guinea in earlier periods. Pig, chicken, and dog are all good indicators of agriculture. From Adwe and Apalo, Gosden and Webb noted that massive soil erosion covered the reef platform soon after the initial Lapita settlement. They hypothesized that the erosion was caused by the creation of gardens and the cutting of forest cover on the top of these raised limestone islands (Gosden and Webb, 1994). See Kirch (1997) and Spriggs (1997) for a further discussion on Lapita swidden (slash and burn). Independent evidence for Lapita agriculture has recently been provided by studies using either residue, phytolith, and pollen studies on Lapita pottery and/or associated sediments. From the sites of Kamgot (Crowther, 2005), Watom (Lentfer and Green, 2004) and Uripiv Island (Vanuatu) (Horrocks and Bedford, 2005), there is now evidence for the presence of Colocasia Esculenta, banana (Eusma) and araceae (palms), respectively.
26
There is some indication that the new forms of agriculture impacted those inhabitants in mainland New Guinea. Bayliss-Smith argues that Austronesian/Lapita influence was felt fairly quickly in the highlands. He writes that at 2,500 years before the present, there is a second energy threshold in the highlands that may be related to the introduction of Austronesian speakers on the coast: “a faint echo of these coastal developments, we can hypothesize in the Highlands zone ... the development of a new, more intensive form of dry land agriculture using introduced alata and esculenta yams” (Bayliss-Smith, 1996: 517). Bayliss-Smith thinks that these changes led to more productive use of grasslands, leading to increased populations, and “greater intensity in exchange networks.” He argues that “to grow yams properly requires soil tillage, and this change has been noted in the stratigraphy at Kuk at c. 2,500 years ago with the deposition of soil aggregates derived from more intensive use of soils in the catchment.”
Summary The Lapita peoples who developed within the Bismarck Archipelago and colonized Remote Oceania approximately 3,000 years ago, brought with them a “colonizing” package that incorporated an adaptation to coastal resources, an introduced subsistence regime including domesticates needed to survive on small and far flung islands, and a social system that ensured interaction between other communities needed for the survival and success of the colonization process. These Lapita settlements were socially related groups with strong communication ties. This does not necessarily mean that the frequency of contact was high. The nature of the data allows the identification of continued interaction, but not its frequency. Parallel changes in the Lapita decorative system occurred in the west and east that were not the result of pottery exchange but rather information exchange that implicates the movement of people. Communication was ongoing, indicating a more socially interactive network over a 1,000-year period.
Post-Lapita Connections and Barriers The Lapita interaction network did not last forever and soon splintered into smaller regional networks by 2,000 years ago (see Spriggs, 1997). Dentate-stamped vessels disappear, which may reflect a breakdown in social networks. The movements of peoples, however, did not stop, and eventually most island groups in Remote Oceania were colonized, including New Zealand. The following section reviews regional movements of people within the last 2,000 years and also the evidence of interactions between major regions.
Island Melanesian Pasts: A View from Archeology
Post-Lapita Connections in Island Melanesia Although having the same founder populations as the islands of western Polynesia, the Melanesian islands of western Remote Oceania had “genetic and cultural inputs from the west,” a point made by Kirch (2000: 156) for Fiji, and applied equally to New Caledonia and Vanuatu as well. Archeologically there is some evidence for continued interaction with these islands and those to the west in Near Oceania. For instance, by 2,000 years ago new vessel forms replace the remnants of the Lapita-derived pottery traditions in Manus, Buka, Vanuatu, and New Caledonia. The appearance of incurved pots, a form rarely found in prior assemblages (e.g. Lapita ware in the Bismarck Archipelago, Koné ware in New Caledonia, Erueti ware and Arapus ware in Vanuatu), and its dominance over complex globular pots and carinated jars, marks a change with what went before. These changes occurred simultaneously from Manus to Buka (Sohano ware), to Vanuatu (Mangaasi ware), and New Caledonia (Oundja and Naia Period pottery) at roughly the same time, suggesting some form of widespread interaction, whatever that may be. Spriggs (1997: 158–61) outlines the arguments for some “secondary migrations” from island Near Oceania to the Melanesian islands of Remote Oceania (see also Bellwood, 1978: 255–70). He makes the important point that any migrations from the Bismarck Archipelago were from peoples who would be “mixed with the original
Bismarck’s inhabitants than the previous Lapita spread, providing a more ‘Melanesian’ phenotype that is found in Vanuatu, New Caledonia and to a lesser extent in Fiji” (Spriggs, 1997: 159). The movements of people, however, were not just from west to east. Some Oceanic-speaking groups made their way back to what we call “Polynesian Outliers.” The Polynesian back colonization of small islands includes those off the main islands of Vanuatu (Mele, Emae, Fila, Aniwa, west Futuna), New Caledonia (West Uvea), the Solomons (Tikopia, Anuta, Taumako, Sikaiana; Bellona, Rennel, Ontong Java, Nukumanu, Takuu), and New Ireland (Nuguria). Indeed, Polynesian Outliers are found in Micronesia (Kapingamarangi and Nukuoro). This back migration from an east to west direction, although only starting about 700 years ago, was part of a process that began three millennia earlier.
South and North Papuan Coasts Descendants of the original Lapita populations colonized the south coast of New Guinea in this more recent period (figure 2.10). Only one occupation site (Kukuba Cave, on the mainland near Yule Island, at 4,000 BP) predates 2,000 years (Vanderwal, 1973, 1978). Judging from adze style similarities, the Oceanic-speaking peoples who entered the region after 2,000 BP came from areas
Figure 2.10 Archeological sites on the south coast of Papua.
27
the framework
in eastern Island Melanesia or western Polynesia (Vanderwal, 1973). Three major archeological sites over a 650-km stretch of coastline provide evidence for this settlement: Oposisi on Yule Island (Vanderwal, 1978), Nebira (Allen, 1972; Bulmer, 1982) near Port Moresby, and Mailu (Irwin, 1985). All three have pottery made with incised and shell impressed decoration, some of which is red slipped and similar to the later Lapita pottery from Bismarck Archipelago assemblages. At each site, the pottery was made locally, although a small amount of obsidian sourced to Fergusson Island, over 150 km east of Mailu, indicates external contacts (see Irwin and Holdaway, 1996). At each site, the obsidian was eventually replaced by locally available chert. By 800 BP, the pottery sequences from these south coast sites start to diverge into localized sequences with the development of specialized trading systems as seen in the assemblage at Motupore Island, which is located 15 km east of Port Moresby. Its prehistoric inhabitants made and exported pottery and shell beads to the Papuan Gulf over 350 km to the west in return for food (Allen, 1977; 1984). These trading systems mark the beginning of the famous Motu Hiri trade systems recorded at European contact. On the north coast of Papua, there is also evidence of a Lapita tradition and extensive trading networks. From Wanigela in Collingwood Bay, Brian Egloff recorded a pedestaled bowl from an undated context that is identical to vessel forms found in Bismarck Archipelago Lapita sites (Egloff, 1978; 1979). He showed that pottery trade was more extensive then than in the famous kula trade cycle recorded by anthropologists early in the 20th century. Prior to 500 years ago, the mainland potters from Collingwood Bay and Goodenough Island supplied pots to the region as a whole, whereas only the Amphlett Island potters supplied the kula exchange system. Thus, both the Trobriand Islands kula ring and the Motu Hiri trade system further west derive from older interaction systems operating in the previous millennia.
Other Connections Within New Guinea, there were extensive trade and exchange networks, as well as some population movements. Only a few of the major points of interactions between and along land masses will be outlined here.
However, evidence for the beginning of the current trade network linking the western tip of New Britain and New Guinea (as seen in Harding, 1967) only dates to 300 years ago (Lilley, 2004: 91). It has generally been argued that the scattered Oceanic speakers on the islands off the north coasts of New Guinea represent a western back migration from communities from the Vitiaz Strait–west New Britain area sometime between 2,000 and 500 years ago (Lilley, 1999: 28). The evidence is primarily linguistic (Ross, 1988), although there are archeological indications as well (see Lilley, 2000, for details).
Connections between Manus and the North Coast of New Guinea The presence of bronze in archeological contexts on Manus dated to 2,000 years ago attests to some connection to the west at this time (Ambrose, 1998), probably via the north coast of New Guinea. Manus influences on the north coast of New Guinea can be seen at least 1,300 years years ago, as evident by the importation of Admiralty obsidian to Tumleo Island, located off Aitape in Sandaun Province (Summerhayes, 2003b). There are references to Manus in the oral history of Koil, in which two brothers from Manus play an ancestral role to the current population. In the colonial period, the German trader Richard Parkinson noted Manus traders sailing to Wewak (straight line distance of 400 km). In 1897, he saw two Admiralty canoes from Mbuke Island on Jacqinot Island in the Schouten group (Parkinson, 1907). Having seen the photographs I have no doubt it was taken on Koil Island, 60 km off Wewak.
Connections between New Ireland and Bougainville Trade networks also existed between the islands off New Ireland and north Solomons (see Parkinson, 1907; Specht, 1974). The archeological evidence, however, suggests the beginning of this network about 700 years ago, with the presence of Buka Island pottery found on Anir and southern New Ireland. These types of exchange did not involve the long-distance movement of peoples, although exchange partners did intermarry between selected clans from each pair of island groups (Nissan to Buka, Anir to Nissan, Tanga to Anir, and so on).
Connections between New Britain and New Ireland. Connections between New Britain and Mainland New Guinea Sporadic interactions between New Britain and New Guinea have been recorded for the last 1,700 years. Type X pottery originating from the Huon Gulf has been found in a number of sites in west New Britain (Lilley, 2004).
28
The best evidence of exchange during the past 2,000 years is the presence of obsidian sourced to west New Britain in New Ireland archeological contexts. The Tolai migrations from southern New Ireland into the Gazelle Peninsula of east New Britain about 750 years ago are well known and important, as well (see Summerhayes, 2004).
Island Melanesian Pasts: A View from Archeology
Discussion These examples suggest both small-scale population migrations and the development of exchange networks. Yet, the fact that New Guinea alone has nearly 1,000 languages is an indication of regional isolation among groups. The populations of the central highland of Papua New Guinea, for instance, were by and large isolated from populations from the lowland regions up until the late 1930s. There were trade networks bringing up salt water shells from the coastal regions (see Hughes, 1977: 188), but the nature of exchange was of a filtering type where goods moved from group to group. By the time the seas shells reached the central highlands, their origins were unknown. In sum, there have been a number of more recent interactions that had to have influenced the genetic composition of Northern Island Melanesian populations. The Polynesian back migrations to the “outliers” are one. The influx of European whalers that entered the area beginning in the early 19th century is another. This had an impact on the populations mostly on the offshore islands of New Ireland and Buka. Here the whalers bartered what they had for food, and also had their share of deserters (see Gray, 1999, for more details). Other immigrants in the form of Polynesian lay priests with various missions arrived in the late 19th century and intermarried with local populations. From Micronesia, the remnants of the Sohks population from Pohnpei were shipped to Rabaul after their rebellion failed and the ringleaders executed by the Germans. Immigrants also include the Samoan connection of Queen Emma and her relations, not to mention the influx of the Germans, Australians, Japanese, Chinese, and Americans. No coastal area of this archipelago I have visited has been without occasional people of mixed ancestry from these countries. The interior areas of larger islands such as New Britain, however, are different and indeed remain isolated even today.
Conclusions Archeologists have often thought that since the initial occupation of New Guinea over 40,000 years ago, the region was relatively isolated until the arrival at 3,300 years ago of Austronesian settlers in the Bismarck Archipelago, who quickly went on to colonize Remote Oceania and other regions. However, it is clear that connections between groups of peoples in Near Oceania developed over many millennia, before and after Lapita, and both within and outside of the region. The idea of spheres of interaction between Southeast Asia and New Guinea is not new (see Allen and Gosden, 1996: 184), although it has rarely been spelled out in
any detail. The primary problem lies in our general ignorance of the archeology of Southeast Asia. We have little idea of the connections and interactions between the landmass of Sahul and Southeast Asia in the remote past. After the first colonization of Sahul at about 40,000 years ago, the next clear archeological evidence of an Asian introduction was the presence of pig in early Holocene contexts. The following major introductions from the west are found post mid-Holocene, with the introduction of domesticates such as chicken and dog, and perhaps the re-introduction of pig, with a new intense form of agriculture, and a new language (protoOceanic), and at least some people as well. While a good bit is now known about the prehistory of Northern Island Melanesia, the archeological record for the New Guinea mainland and Island Southeast Asia in the Pleistocene and Holocene is sparse. Firmly establishing “Past Connections” across the entire region is an endeavor that remains incomplete.
Acknowledgments I wish to thank Pam Swadling, Jim Specht, Chris Gosden, and Jim Allen for their important work in all facets of New Guinea archeology. I wish also to thank Robert Mondol and Baiva Ivuyo, both deceased, for their dedication and hard work in bringing to life Papua New Guinea’s past. Special thanks for Roger Green, Andrew Pawley, Pam Swadling, and Jonathan Friedlaender for their comments on this chapter.
Note 1. Events described as “years ago” incorporate age estimates calibrated to calendar years, while those described as “BP” use uncalibrated radiocarbon determinations. “BP” stands for years before 1950.
References Allen J. 1972. Nebira 4: An early Austronesian site in Central Papua, Archaeology and Physical Anthropology in Oceania 7: 92–124. Allen J. 1977. Sea traffic, trade and expanding horizons. In: Allen J, Golson J, Jones R, editors. Sunda and Sahul: Prehistoric studies in Island Southeast Asia, Melanesia and Australia. London: Academic Press. pp 387–417. Allen J. 1984. Pots and poor princes: A multidimensional approach to the study of pottery trading in coastal Papua. In: van der Leeuw S, Pritchard A, editors. The many dimensions of Papua: Ceramics in archaeology and anthropology. Amsterdam: Universiteit van Amsterdam. pp 407–63.
29
the framework
Allen J. 1991. Introduction. In: Allen J, Gosden C, editors. Report of the Lapita Homeland Project. Canberra: Occasional Papers in Prehistory 20, Department of Prehistory, Research School of Pacific Studies, Australian National University. pp 1–7. Allen J. 1996. The pre-Austronesian settlement of Island Melanesia: Implications for Lapita archaeology. In: Goodenough WG, editor. Prehistoric settlement of the Pacific, Philadelphia: Transactions of the American Philosophical Society 86. pp 11–27. Allen J. 2000. From beach to beach: The development of maritime economies in prehistoric Melanesia. Modern Quaternary Research in Southeast Asia 16: 137–76. Allen J. 2003. Discovering the Pleistocene in Island Melanesia, in Sand C, editor. Pacific archaeology: Assessments and prospects. Proceedings of the International Conference for the 50th anniversary of the first Lapita excavation July 1952, Koné-Nouméa. Nouméa: Le Cahiers de l’Archéologie en NouvelleCalédonie 15. pp 33–42. Allen J, Gosden C. 1996. Spheres of interaction and integration: Modelling the culture history of the Bismarck Archipelago. In: Davidson J, Irwin G, Leach F, Pawley A, Brown D, editors. Oceanic culture history: Essays in honour of Roger Green. Dunedin. North: New Zealand Journal of Archaeology Special Publications. pp 183–97. Allen J, White JP. 1989. The Lapita Homeland: Some new data and an interpretation. Journal of the Polynesian Society 98: 129–46. Allen J, Gosden, C White JP. 1989. Human Pleistocene adaptations in the tropical island Pacific: Recent evidence from New Ireland, a greater Australian outlier. Antiquity 63: 548–61. Ambrose W. 1988. An early bronze artefact from Papua New Guinea. Antiquity 62: 483–91. Ambrose W. 2002. From the very old to new, obsidian artefacts in the Admiralty Islands. In: Kaufmann C, Kocher Schmid C, Opnemus S, editors. Admiralty Islands: Art from the South Seas. Zurich: Museum Rietberg. pp 67–72 Anderson A, Bedford S, Lilley I, Sand C, Summerhayes G, Torrence R. 2001. An inventory of Lapita sites containing dentate-stamped pottery. In: Clark GR, Anderson A, Vunidilo T, editors. The archaeology of Lapita dispersal in Oceania. Canberra: Terra Australis 17, Pandanus Books. pp 1–14. Anderson D. 1990. Lang Rongrien Rockshelter. Philadelphia: University of Pennsylvania Museum. Anson D. 1983. Lapita pottery of the Bismarck Archipelago and its affinities. PhD dissertation. University of Sydney. Araho N, Torrence R, White JP. 2002. Valuable and useful: mid-Holocene stemmed obsidian artefacts from West New Britain, Papua New Guinea, Proceedings of the Prehistoric Society 68: 61–81. Bailey RC, Head G, Jenike M, Owen B, Rechtman R, Zechenter E. 1989. Hunting and gathering in tropical
30
rain forest: Is it possible? American Anthropologist 91: 59–82. Barker G, Badang D, Barton H, Beavitt P, Bird M, Daly P, Doherty C, Gilbertson D, Glover I, Hunt C, Manser J, McLaren S, Paz V, Pyatt B, Reynolds T, Rose J, Rushworth G, Step M. 2001. The Niah Cave Project: The second season of fieldwork. Sarawak Museum Journal 56: 37–119. Barton H, White JP. 1993. Use of stone and shell artefacts at Balof 2, New Ireland, Papua New Guinea. Asian Perspectives 32, 169–81. Bayliss-Smith T. 1996. People–plant interaction in the New Guinea highlands: Agricultural heartland or horticultural backwater? In: Harris DR, editor. The origins and spread of agriculture and pastoralism in Eurasia. Washington DC Smithsonian Institute Press. pp 499–523. Bellwood P. 1978. Man’s conquest of the Pacific: The prehistory of Southeast Asia and Oceania. Auckland: Collins. Bellwood P. 1998. From Bird’s Head to bird’s eye view; Long term structures and trends in Indo-Pacific prehistory. In: Miedema J, Ode C., Dam RAC, editors. Perspectives in the Bird’s Head of Irian Jaya, Indonesia. Amsterdam: Rodopoi. pp 951–75. Bellwood P, Nitihaminoto G, Irwin G, Gunadi, Waluyo A, Tanudirjo D. 1998. 35,000 years of prehistory in the northern Moluccas. Modern Quaternary Research in Southeast Asia 15: 233–73. Bird MI, Hope G, Taylor D. 2004. Populating PEP II: The dispersal of humans and agriculture through AustralAsia and Oceania. Quaternary International 118–19: 145–63. Bulmer S. 1975. Settlement and economy in prehistoric Papua New Guinea. Journal de la Société des Océanistes 31: 7–75. Bulmer S. 1977. Between the mountain and the plain. Prehistoric settlement and environment in the Kaironk Valley. In: Winslow JH, editor. The Melanesian environment. Canberra: Australian National University Press. pp 61–73. Bulmer S. 1982. Human ecology and cultural variation in prehistoric New Guinea. In: Gressitt J, editor. Biogeography and ecology of New Guinea. The Hague: Dr W.Junk. pp 169–206. Bulmer S. 1991. Variation and change in stone tools in the Highlands of Papua New Guinea: The witness of Wanlek. In: Pawley A, editor. Man and a half: Essays in Pacific anthropology and ethnobiology in honour of Ralph Bulmer. Auckland: Polynesian Society. pp 470–8. Chappell J. 2000. Pleistocene seedbeds of western Pacific maritime cultures. Modern Quaternary Research in Southeast Asia 16: 77–98. Christensen O. 1975. Hunters and horticulturalists: A preliminary report of the 1972–1974 excavations in the Manim Valley, PNG. Mankind 10: 24–36. Crowther A. 2005. Starch residues on undecorated Lapita pottery from Anir, New Ireland. Archaeology in Oceania 40: 62–6.
Island Melanesian Pasts: A View from Archeology
Denham T, Haberle SG, Lentfer C, Fullagar R, Field J, Therin M, Porch N, Winsborough B. 2003. Origins of agriculture at Kuk Swamp in the Highlands of New Guinea. Science 301: 189–93. Egloff B. 1978. The Kula before Malinowski: A changing configuration. Mankind 11(3): 429–35. Egloff B. 1979. Recent prehistory in Southeast Papua. Canberra: Terra Australis 4, Australian National University. Fairbairn A, Swadling P. 2005. Re-dating mid-Holocene betelnut (Areca Carechu L.) and other plant use at Dongan, Papua New Guinea. Radiocarbon 47: 377–82. Flannery TF. 1995. Mammals of the South-West Pacific and Moluccan Islands. Sydney: Reed Books. Flannery TF, White JP. 1991. Animal translocations. National Geographic Research and Exploration 7(1): 96–113. Fox CE. 1925. The threshold of the Pacific. London: Kegan Paul, Trubner and Co. Friedlaender JS, Schurr TG, Gentz F, Koki G., Friedlaender FR, Horvat G., Babb P, Cerchio S, Kaestle F, Schanfield M, Deka R, Yanagihara R, Merriwether DA. 2005a. Expanding Southwest Pacific mitochondrial haplogroups P and Q. Molecular Biology and Evolution 22 (6) 1506–17. Friedlaender JS, Gentz F, Friedlaender FR, Kaestle F, Schurr TG, Koki G, Schanfield M, McDonough J, Smith L, Cerchio S, Mgone C, Merriwether DA. 2005b. Mitochondrial genetic diversity and its determinants in Island Melanesia. In: Pawley A, Attenborough R, Golson J, Hyde R, editors. Papuan pasts: Studies in the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra. Pacific Linguistics. pp 693–716. Fullagar R. 1992. Lithically Lapita. Functional analysis flaked stone assemblages from West New Britain Province, Papua New Guinea. In: Galipaud JC, editor. Poterie Lapita et Peuplement. Noumea: ORSTOM. pp 135–43. Fullagar R. 1993. Flakes stone tools and plant food production: A preliminary report on obsidian tools from Talasea, West New Britain, PNG. Traces et Fonction: Les Gestes Retrouvés, Colloque international de Liège editions ERAUL, 50: 331–7. Gillieson D, Mountain M-J. 1983. Environmental history of Nombe rockshelter, Papua New Guinea. Archaeology in Oceania 18: 53–62. Glover I. 1986. Archaeology in Eastern Timor, 1966–67, Canberra: Terra Australia 11. Department of Prehistory, Research School of Pacific Studies, Australian National University. Golson J. 1991. Bulmer Phase II: Early agriculture in the New Guinea Highlands. In: Pawley A, editor. Man and a half: Essays in Pacific anthropology and ethnobiology in honour of Ralph Bulmer. Auckland: Polynesian Society. pp 484–91. Golson J. 1997. From horticulture to agriculture in the New Guinea Highlands: A case study of people and their environments. In: Kirch PV, Hunt TL, editors.
Historical ecology in the Pacific Islands: Prehistoric environmental and landscape change. New Haven: Yale University Press. pp 39–50. Golson J. Gardner D. 1990. Agriculture and sociopolitical organization in New Guinea Highlands prehistory. Annual Review of Anthropology 19: 395–417. Gorecki P. 1992. A Lapita smokescreen? In: Galipaud JC, editor. Poterie Lapita et Peuplement. Noumea: ORSTOM. pp 27–47. Gorecki P, Mabin M, Campbell J. l991. Archaeology and geomorphology of the Vanimo Coast, Papua New Guinea: Preliminary results. Archaeology in Oceania 26: 119–22. Gosden C. 1993. Understanding the settlement of the Pacific Islands in the Pleistocene. In: Smith MA, Spriggs M, Fankhauser B, editors. Sahul in review: Pleistocene archaeology in Australia and Island Melanesia. Canberra: Occasional Papers in Prehistory 24, Department of Prehistory, RSPAS, Australian National University. pp 131–6. Gosden C, Webb J. 1994. The creation of a Papua New Guinean landscape: Archaeological and geomorphological evidence. Journal of Field Archaeology 21: 29–51. Gray A. 1999. Trading contacts in the Bismarck Archipelago during the whaling era, 1799–1884. The Journal of Pacific History 34: 23–44. Green RC. 1979a. Early Lapita art from Polynesia and Island Melanesia. In: Mead S, editor. Exploring the visual art of Oceania. Honolulu: University of Hawaii Press. pp 13–31. Green RC. 1979b. Lapita. In: Jennings J, editor. The Prehistory of Polynesia. Canberra: Australian National University Press. pp 27–60. Green RC. 1991. The Lapita cultural complex: Current evidence and proposed models. Bulletin of the IndoPacific Prehistory Association 11: 295–305. Green RC. 2003. The Lapita horizon and traditions— Signature for one set of oceanic migrations. In: Sand C, editor. Pacific archaeology: Assessments and prospects. Proceedings of the International Conference for the 50th anniversary of the first Lapita excavation July 1952, Koné-Nouméa. Nouméa: Le Cahiers de l’Archéologie en Nouvelle-Calédonie 15. pp 95–120. Groube L. 1989. The taming of the rain forests: A model for late Pleistocene forest exploitation in New Guinea. In: Harris DR, Hillman GC, editors. Foraging and farming. London: Unwin Hyman. pp 292–304. Groube L, Chappell J, Muke J, Price D. 1986. A 40,000 year old occupation site at Huon Peninsula, Papua New Guinea. Nature 324: 453–5. Haberle SG, David B. 2004. Climates of change: Human dimensions of Holocene environmental change in low latitudes of the PEPII transect. Quaternary International, 118–19: 165–79. Haberle S, Hope GS, DeFretes Y. 1990. Environmental change in the Baliem Valley, montane Irian Jaya, Republic of Indonesia. Journal of Biogeography, 18: 25–40.
31
the framework
Harding T. 1967. Voyagers of the Vitiaz Strait: A study of a New Guinea trade system. Seattle. University of Washington Press. Hay A. 1990. Aroids of New Guinea. Christenson Research Institute. Madang: Kristen Press Inc. Hayden B. 1983. Social characteristics of early Austronesian colonisers. Bulletin of the Indo-Pacific Association 4: 123–34. Hedges REM, Housley RA, Bronk Ramsey C, van Klinken GJ. 1995. Radiocarbon dates from the Oxford AMS system: Archaeometry data list 20. Archaeometry 37: 417–30. Heinsohn T. 2003. Animal translocations: Long-term human influences on the vertebrate zoogeography of Australasia natural dispersal versus ethnophoresy. Australian Zoologist 32: 351–76. Heinsohn T. 2004a. Ecological variability in the common spott(ed) cuscus Spilocuscus maculatus in the Australasian Archipelago—a review. In: Goldingray RL, Jackson SM, editors. The biology of Australian possums and gliders. Sydney: Surrey Beatty and Sons. pp 527–38. Heinsohn T. 2004b. Phalangeriods as ethnotramps: A brief history of possums and gliders as introduced sprecies. In: Goldingray RL, Jackson SM, editors. The biology of Australian possums and gliders. Sydney: Surrey Beatty and Sons. pp 506–26. Hill JN. 1985. Style: A conceptual evolutionary framework. In: Nelson B, editor. Decoding prehistoric ceramics. Carbondale: Southern Illinois University Press. pp 362–85. Hope GS, Golson J. 1995. Late Quaternary change in the mountains of New Guinea. Antiquity 265: 818–30. Hope GS, Golson J, Allen J. 1983. Palaeoecology and prehistory in New Guinea. Journal of Human Evolution 12: 37–60. Horrocks M, Bedford S. 2005. Microfossil analysis of Lapita deposits in Vanuatu reveals introduced Araceae (ariods). Archaeology in Oceania 39: 67–74. Hughes I. 1977. New Guinea Stone Age trade. Terra Australis 3. Canberra: Department of Prehistory, Research School of Pacific and Asian Studies, The Australian National University. Irwin GJ. 1985. The emergence of Mailu. Terra Australis 10, Canberra: Department of Prehistory, Research School of Pacific Studies, Australian National University. Irwin GJ. 1992. The prehistoric exploration and colonisation of the Pacific. Melbourne: Cambridge University Press. Irwin GJ, Holdaway S. 1996. Colonisation, trade and exchange: From Papua to Lapita. In: Davidson J, Irwin GJ, Leach F, Pawley A, Brown D, editors. Oceanic culture history: Essays in honour of Roger Green. Dunedin North: New Zealand Journal of Archaeology Special Publication. pp 225–35. Ivens WG. 1927. Melanesians of the southeast Solomon Islands. London: Trench Trubner and Co. Kennedy J. 1983. On the prehistory of western Melanesia, the significance of new data from the Admiralties. Australian Archaeology 16: 115–22.
32
Kennedy J. 2002. Manus from the beginning: An archaeological overview. In: Kaufmann C, Kocher Schmid C, Ohnemus S, editors. Admiralty art: Art from the South Seas. Zurich: Museum Rietberg. pp 17–28. Kennedy J, Clarke W. 2004. Cultivated landscapes of the southwest Pacific. Canberra: Resource Management in Asia-Pacific Program Working Paper No. 50. Resource Management in Asia-Pacific Program, Research School of Pacific and Asian Studies, Australian National University http://rspas.anu.(ed)u.au/papers/rmap/ Wpapers/rmap_wp50.pdf Kirch PV. 1988. The Talapakemalai Lapita site and Oceanic prehistory. National Geographic Research 4: 328–42. Kirch PV. 1997. The Lapita peoples: Ancestors of the Oceanic World. Oxford: Blackwells. Kirch PV. 2000. On the Road of the Winds: An archaeological history of the Pacific Islands before European contact. Berkeley: University of California Press. Kirch PV. 2001. Lapita and its transformations in Near Oceania: Archaeological investigations in the Mussau Islands, Papua New Guinea, 1985–1988. Berkeley: Archaeological Research Facility Contribution No. 59., University of California at Berkeley. Kirch PV, Hunt TL. 1988. The spatial and temporal boundaries of Lapita. In: Kirch PV, Hunt TL, editors. Archaeology of the Lapita Cultural Complex: A critical review. Seattle: Thomas Burke Memorial Washington State Museum Research Report No. 5, Burke Museum. pp 9–31. Kirch PV, Allen MS, Butler VL, Hunt TL. 1987. Is there an early Far Western Lapita province? Sample size effects and new evidence from Eloaua Island. Archaeology in Oceania 22: 123–7. Leavesley M, Allen J. 1998. Dates, disturbance and artefact distributions: Another analysis of Buang Merabak, a Pleistocene site on New Ireland, Papua New Guinea. Archaeology in Oceania 33: 63–82. Leavesley M, Chapell J. 2004. Buang Merabak: Additional early radiocarbon evidence of the colonisation of the Bismarck Archipelago, Papua New Guinea. Antiquity 78, No. 301. Leavesley MG, Bird MI, Fifield LK, Hausladen PA, Santos GM, di Tada ML. 2002. Buang Merabak: Early evidence for human occupation in the Bismarck Archipelago: Papua New Guinea. Australian Archaeology 54: 55–7. Lentfer CJ, Green R. 2004. Phytoliths and the evidence for banana cultivation at the Lapita Reber-Rakival Site on Watom Island, Papua New Guinea. Records of the Australian Museum Supplement 29: 75–88. Lilley I. 1999. To good to be true? Post-Lapita scenarios for language and archaeology in West New Britain–North New Guinea. Indo-Pacific Prehistory Association Bulletin 18: 25–34. Lilley I. 2000. Migration and ethnicity in the evolution of Lapita and post-Lapita maritime societies in northwest Melanesia. Modern Quaternary Research in Southeast Asia 16: 177–95.
Island Melanesian Pasts: A View from Archeology
Lilley I. 2004. Trade and culture history across the Vitiaz Strait, Papua New Guinea: The emerging post Lapita coastal sequence. Records of the Australian Museum, Supplement 29: 89–96. Loy TH, Spriggs M, Wickler S. 1992. Direct evidence for the human use of plants 28,000 years ago: Starch residues on stone artefacts from the Northern Solomon Islands. Antiquity 66: 898–912. Marshall B, Allen J. 1991. Excavations at Panakiwuk Cave, New Ireland. In: Allen J, Gosden C, editors. Report of the Lapita Homeland Project. Canberra: Occasional Papers in Prehistory 20, Department of Prehistory, Research School of Pacific Studies. pp 59–91. Mountain M-J. 1991. Bulmer Phase I: Environmental change and human activity through the late Pleistocene into the Holocene in the Highlands of New Guinea: A scenario. In: Pawley A, editor. Man and a half: Essays in Pacific anthropology and ethnobiology in honour of Ralph Bulmer. Auckland: Polynesian Society. pp 510–20. O’Connell JF, Allen J. 2004. Dating the colonization of Sahul (Pleistocene Australia-New Guinea): A review of recent research. Journal of Archaeological Science 31: 835–53. O’Connor S, Spriggs M, Veth P. 2002. Excavation at Lene Hara Cave establishes occupation in East Timor at least 30,000–35,000 years ago. Antiquity 76: 46–9. Parkinson R. 1907. Dreizig Jahre in der Südsee. Stuttgart: Strecker and Schröder. Pavlides C. 1993. New archaeological research at Yombon, West New Britain, Papua New Guinea. Archaeology in Oceania 28: 55–9. Pavlides C, Gosden C. 1994. 35,000 year old sites in the rainforests of west New Britain. Antiquity 68: 604–10. Pasveer J. 2004. The Djief Hunters: 26,000 years of rainforest exploitation on the Bird’s Head of Papua, Indonesia. Modern Quaternary Research in Southeast Asia 17. London: AA. Balkema Publishers. Pawley A. 2003. Locating Proto Oceanic. In: Ross M, Pawley A, Osmond M, editors. The lexicon of Proto Oceanic: The culture and environment of ancestral Oceanic Society. Canberra: Pacific Linguistics 545, Research School of Pacific and Asian Studies, Australian National University. pp 17–34. Powell J. 1977. Plants, man and environment in the Island of Papua New Guinea. In: Winslow JH, editor. The Melanesian environment. Canberra: The Australian National University Press. pp 11–20. Rathje WL. 1978. Melanesian and Australian exchange systems. A view from Mesoamerica. Mankind 11: 165–74. Ross M. 1988. Proto Oceanic and the Austronesian languages of Western Melanesia. Pacific Linguistics C-98. Canberra: Australian National University. Sillitoe P. 2002. Always been farmer-foragers? Hunting and gathering in the Papua New Guinea Highlands. Anthropological Forum 12: 45–76. Specht J. 1974. Of Menak and man: Trade and the distribution of resources on Buka Island, Papua New Guinea. Ethnology 12: 225–37.
Specht J. 2005. Revisiting the Bismarcks: Some alternative views. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts: Studies in the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra: Pacific Linguistics, Research School of Pacific and Asian Studies, Australian National University. pp 235–88. Spriggs M. 1989. The dating of the Island Neolithic: An attempt at chronometric hygiene and linguistic correlation. Antiquity 63: 587–613. Spriggs M. 1990. The changing face of Lapita: Transformation of a design. In: Spriggs M, editor. Lapita design, form and composition. Canberra: Occasional Papers in Prehistory 19, Department of Prehistory, Research School of Pacific Studies, Australian National University. pp 83–122. Spriggs M. 1991. Nissan, The island in the middle. Summary report on excavations at the north end of the Solomons and south end of the Bismarcks. In: Allen J, Gosden C, editors. The report of the Lapita Homeland Project. Canberra: Occasional Papers in Prehistory 20, Department of Prehistory, Research School of Pacific Studies, Australian National University. pp 222–43. Spriggs M. 1996. What is southeast Asian about Lapita?, In: Akazaqa T,. Szathmary EJ, editors. Prehistoric Mongoloid dispersals. Tokyo: Oxford University Press. pp 324–48. Spriggs M. 1997. Island Melanesians. Oxford: Blackwell. Spriggs M. 1998. The archaeology of the Bird’s Head in its Pacific and Southeast Asian context. In: Miedema J, Ode C, Dam RAC, editors. Perspectives on the Bird’s Head of Irian Jaya, Indonesia. Amsterdam: Rodopoi. pp 931–9. Spriggs M. 2000. Can hunter-gatherers live in tropical rain forests? In: Schweitzer A, Biesele M, Hitchcock RK, editors. Hunters and gatherers in the modern world. New York: Berghan Books. pp 287–304. Summerhayes GR. 1998. The face of Lapita. Archaeology in Oceania 33: 100. Summerhayes GR. 2000a. Lapita interaction. Canberra: Terra Australis No.15, Centre of Archaeology, Australian National University. Summerhayes GR. 2000b. Far western, western and eastern Lapita—A re-evaluation. Asian Perspectives 39: 109–38. Summerhayes GR. 2000c. What’s in a pot? In: Anderson AJ, Murray T, editors. Australian Archaeologist: Collected papers in honour of Jim Allen. Canberra: Coombs Academic Publishing, Australian National University. pp 291–307. Summerhayes GR. 2001a. Lapita in the far west: Recent developments. Archaeology in Oceania 36: 53–64. Summerhayes GR. 2001b. Defining the chronology of Lapita in the Bismarck Archipelago. In: Clark GR, Anderson AJ, Vunidilo T, editors. The archaeology of Lapita dispersal in Oceania. Canberra: Terra Australis 17 Pandanus Books. pp 25–38. Summerhayes GR. 2003a. The rocky road; the selection and transport of Admiralties obsidian to Lapita communities. Australian Archaeology 57: 135–42.
33
the framework
Summerhayes GR. 2003b. Modelling differences between Lapita obsidian and pottery distribution patterns in the Bismarck Archipelago. In: Sand C, editor. Pacific archaeology: Assessments and prospects. Proceedings of the International Conference for the 50th anniversary of the first Lapita excavation July 1952, KonéNouméa. Nouméa: Le Cahiers de l’Archéologie en Nouvelle-Calédonie 15. pp 139–49. Summerhayes GR. 2004. The nature of prehistoric obsidian importation to Anir and the development of a 3,000 year old regional picture of obsidian exchange within the Bismarck Archipelago, Papua New Guinea. In: Attenbrow VJ, Fullagar R, editors. Archaeologist and anthropologist in the Western Pacific: Essays in honour of Jim Specht. Sydney: The Records of the Australian Museum Supplement 29. pp 145–56. Summerhayes GR, Allen J. 1993. The transport of Mopir obsidian to Late Pleistocene New Ireland. Archaeology in Oceania 28: 145–9. Summerhayes GR, Bird R, Fullagar R, Gosden C, Specht J, Torrence R. 1998. Application of PIXE-PIGME to archaeological analysis of changing patterns of obsidian use in West New Britain, Papua New Guinea. In: Shackley S, editor. Advances in Archaeological Volcanic Glass Studies. New York: Plenum Press. pp 129–58. Swadling P. 1990. Sepik prehistory. In: Lukethaus N, Kaufmann C, Mitchell WE, Newton D, Osmundsen L, Schuster M, editors. Sepik heritage: Tradition and change in Papua New Guinea. Bathurst: Crawford House. pp 71–86. Swadling P. 1997. Changing shorelines and cultural orientations in the Sepik-Ramu, Papua New Guinea: Implications for Pacific prehistory. World Archaeology 29: 1–14. Swadling P. 2005. The Huon Gulf and its hinterland: A long-term view of coastal–highlands interaction. In: Gross C, Lyons H and Counts D, editors. A polymath anthropologist: Essays in honour of Ann Chowning. Research in Anthropology and Linguistics Monograph Series Vol. 6, Department of Anthropology, University of Auckland. pp 1–14. Swadling P, Hide R. 2005. Changing landscape and social interaction, looking at agricultural history from a SepikRamu perspective. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan Pasts: Studies in the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra: Pacific Linguistics, Research School of Pacific and Asian Studies, Australian National University. pp 289–327. Swadling P, Chappell J, Francis G, Araho N, Ivuyo B. 1989. A Late Quaternary inland sea and early pottery in Papua New Guinea. Archaeology in Oceania 24:106–9. Swadling P, Araho N, Ivuyo B. 1991. Settlements associated with the inland Sepik-Ramu Sea. Indo-Pacific Prehistory Association Bulletin 11: 92–112. Torrence R. 2004. Now you see it now you don’t: Changing obsidian source use in the Willaumez Peninsula, Papua New Guinea. In: Cherry J, Scarre C, Shennan S, editors.
34
Explaining social change: Studies in honour of Colin Renfrew. McDonald Institute Monographs, Cambridge. pp 115–25. Torrence R, Summerhayes GR. 1997. Sociality and the short distance trader: Intra-regional obsidian exchange in the Willaumez Peninsula, Papua New Guinea. Archaeology in Oceania 32: 74–84. Torrence, R, White JP. 2001. Tattooed faces from Boduna Island, Papua New Guinea. In: Clark, G, Anderson AJ and Vunidilo Y, editors. The archaeology of Lapita dispersal in Oceania, Canberra: Terra Australis 17, Pandanus Books. pp 135–40. Torrence R, Pavlides C, Jackson P, Webb J. 2000. Volcanic disasters and cultural discontinuities in Holocene time, in West New Britain, Papua New Guinea. In: McGuire WJ, Griffiths DR, Hancock PL, Stewart IS, editors. The archaeology of geological catastrophes. London: The Geological Society. Special Publication 171. pp 225–44. Torrence R, Neall R, Doelman T, Rhodes E, McKee C, Davies H, Bonetti R, Gugliemetti A, Manzoni A, Oddone M, Parr J, Wallace C. 2004. Pleistocene colonisation of the Bismarck Archipelago: New evidence from West New Britain. Archaeology in Oceania 39: 101–30. Vanderwal R. 1973. Prehistoric studies in Central Coastal Papua, Canberra: Unpublished PhD thesis, Australian National University. Vanderwal R. 1978. Exchange in prehistoric Coastal Papua. Mankind 11(3): 416–25. Veth P, O’Connor S, Spriggs M. 1998. After Wallace: Preliminary results of the first season’s excavation of Liang Lemdubu, Aru Islands, Maluka. In: Klokke MJ, de Bruijn T, editors. Proceedings of the 6th International Conference of the European Association of Southeast Asian Archaeologists Leiden. Hull: Special Issue, Centre for South East Asian Studies Hull: University of Hull. pp 75–85. White JP. 1972. Ol Tumbuna: Archaeological excavations in the eastern Central Highlands, Papua New Guinea. Canberra: Terra Australis 2. The Australian National University. White JP. 2004. Where the wild things are: Prehistoric animal translocation in the Circum New Guinea Archipelago. In: Fitzpatrick SM, editor. Voyages of discovery: The archaeology of islands. London. Praeger. pp 147–64. White JP, Allen J. 1980. Melanesian prehistory: Some recent advances. Science 207: 728–34. White JP, O’Connell J. 1982. A Prehistory of Australia, New Guinea and Sahul. Sydney: Academic Press. White JP, Crook KAW, Ruxton BP. 1970. Kosipe: A Pleistocene site in the Papua Highlands. Proceedings of the Prehistoric Society 36: 152–70. White JP, Flannery TF, O’Brien R, Hancock RV, Pavlish L. 1991. The Balof Shelters, New Ireland. In: Allen J, Gosden C, editors. Report of the Lapita Homeland Project. Canberra: Occasional Papers in Prehistory 20, Department of Prehistory, Research School of Pacific Studies. pp 46–58.
Island Melanesian Pasts: A View from Archeology
White JP, Clark G, Bedford S. 2000. Distribution, present and past, of Rattus praetor in the Pacific and its implications. Pacific Science 54: 105–17. Wickler S. 2003. The prehistory of Buka: A stepping stone island in the Northern Solomons, Canberra: Terra Australia 16, Centre of Archaeology and the Department of Natural History, The Australian National University. Wickler S, Spriggs M. 1988. Pleistocene human occupation of the Solomon Islands, Melanesia. Antiquity 62: 703–6. Yen D. 1982. The history of cultivated plants. In: May R, Nelson H, editors. Melanesia: Beyond diversity,
Vol. 1. Canberra: Research School of Pacific and Asian Studies, The Australian National University. pp 281–95. Yen D. 1990. Environment, agriculture and the colonization of the Pacific. In: Yen D, Mummery JMJ, editors. Pacific production systems: Approaches to economic history. Canberra: Department of Prehistory, Research School of Pacific and Asian Studies, The Australian National University. pp 258–77. Yen D. 1995. The development of Sahul agriculture with Australia as bystander. Antiquity 69: 831–47.
35
3 Recent Research on the Historical Relationships of the Papuan Languages, or, What Does Linguistics Say about the Prehistory of Melanesia? Andrew Pawley
Introduction My purpose here is to respond to the question: What does comparative linguistics tell us about the history of the Papuan-speaking peoples? To this end I will review research on the history of the Papuan languages and ask to what extent does linguistic evidence corroborate, contradict, or add to the evidence of archeology and other historical disciplines, and vice versa? I will, however, leave to others better qualified than me to consider the evidence of biological anthropology. Linguists have applied the term ‘Papuan’ to a number of language families and isolates that have in common three things: (a) they are indigenous to New Guinea and nearby island groups; (b) they do not belong to the vast Austronesian family; and (c) unlike Austronesian, they have no relatives outside of the Melanesia–East Indonesia region. However, the various Papuan families are not known to be related to each other. ‘Papuan’ is merely a useful label for those genetically diverse non-Austronesian languages that are sandwiched between Australia to the south and the vast Austronesianspeaking area to the west, north, and east. The various Papuan families almost certainly represent continuations of linguistic stocks that have been in New Guinea and Island Melanesia for millennia before Austronesian speakers arrived there some 3,000 years ago. The hub of the Papuan-speaking region is the island of New Guinea. This island, 2,400 km long, has an area
36
not much larger than France but contains some 900 mutually unintelligible languages. About 750 of these are Papuan. Another 50 or so Papuan languages are spoken outside of New Guinea. The westernmost Papuan languages are spoken in Timor and nearby Alor and Pantar, and in Halmahera. The easternmost ones are in the Solomon Islands. A number are spoken north and northeast of New Guinea in New Britain, New Ireland, and Bougainville. It will be useful here to distinguish the following biogeographical regions, whose boundaries have played an important role in limiting the movement of plants, animals, and people (see figure 3.1): (i) Near Oceania, which encompasses New Guinea, the Admiralties, New Britain, New Ireland, and the main Solomons chain as far east as Makira (San Cristobal), in contrast to Remote Oceania, consisting of the rest of the Pacific Islands; (ii) Northern Island Melanesia, which is all of Near Oceania except for New Guinea; (iii) Greater Bougainville, consisting of Bougainville and extensions during the last glacial maximum, which reached as far as present-day Nggela (Florida) in the central Solomons; (iv) Sundaland, that part of today’s IndoMalaysian archipelago which was an extension of the SE Asian continent when sea levels were at their lowest in the Pleistocene, including present day Sumatra, Borneo, Java, and Bali and Palawan but not the main Philippines group, Sulawesi or the Lesser Sundas; (v) Sahul, the Greater Australia continent that included
Recent Research on the Historical Relationships of the Papuan Languages
Figure 3.1 Major biogeographic regions of Island SE Asia and the Pacific: Sundaland, Wallacea, Near Oceania, and Remote Oceania.
New Guinea when sea-levels were lower; and (vi) Wallacea, the islands between Sundaland and Sahul, which during the Pleistocene were always separated from both continents by ocean gaps.
Reconstructing the Prehistory of Near Oceania from Archeology and Associated Disciplines Here I summarize those parts of the archeological and geomorphological record that can usefully be compared
with the evidence from historical linguistics. For a detailed review of the archeological record for Near Oceania the reader should consult chapter 2. Based on changes in technology, trade patterns, patterns of mobility and settlement, and other variables, archeologists have found it useful to distinguish between several periods of prehistory in Near Oceania. These periods are roughly pre-20,000 BP, 20–10,000 BP, 10–6,000, 6–3,000 BP and 3,000 BP to the time of first written records.
37
the framework
Humans entered Sahul more than 40,000 years ago. To get there from Sundaland people had to cross Wallacea, making a number of ocean voyages some of more than 70 km. Two sites from Australia have been dated at between 50 and 66,000 BP (Roberts et al., 1990; Thorne et al., 1999) although these dates are disputed. Other early dates cluster around 40,000 BP. The earliest dates for New Guinea are from two sites on the north coast—one from uplifted coral terraces on the Huon Peninsula, dated at between 40 and 50,000 BP (Groube, 1986; Groube et al., 1986; Chappell et al., 1994) and the other at Lachitu, near Vanimo, dated to 39,000 BP. By 39,000 BP people had made the crossing from New Guinea to New Britain, the nearest part of Island Melanesia, which requires a 90 km direct voyage from the Huon Peninsula or shorter steps by island hopping, and had reached New Ireland (Allen and Gosden, 1996; Pavlides and Gosden, 1994; Specht, 2005). By 29–28,000 BP people had made the 180 km crossing from New Ireland to the northern end of Bougainville (Spriggs 1997). By at least 21,000 BP people were in Manus (Specht 2005), a crossing that required a voyage of about 200 km from New Guinea or on the westflowing current from New Ireland. It is inconceivable that such a series of ocean crossings could have been made without seaworthy craft. Even if their longer voyages were unintentional these early colonizers of Near Oceania must have been competent makers of craft designed for short inter-island crossings, probably rafts. However, the voyaging capacities of these Pleistocene sailors were limited (Anderson, 2000). During the Pleistocene interaction between the Bismarcks and Bougainville seems to have been minimal (Specht, 2005; Spriggs, 2000; Summerhayes, 2000a,b), and the initial phase of expansion into the southwest Pacific got no further than the main Solomons chain. It was not until late in the 2nd millennium BC that people settled any part of Remote Oceania. To reach the nearest islands of Remote Oceania from the main Solomons chain one must make ocean crossings of 350 km and more to the Santa Cruz-Reefs group and the Vanuatu archipelago. Such voyages evidently did not happen until outriggers with sails came on the scene. What little that survives of the tool kits and habitation sites from the pre-20,000 BP Pleistocene sites in Near Oceania indicates that the people were broad-spectrum foragers, hunting and gathering a range of animals and plants. The basic social groups must have been small, mobile bands of close kin who ranged over a territory. There were no truly sedentary settlements, only camps and seasonal bases. It is likely that the larger islands of Near Oceania remained very sparsely peopled during the period when people were primarily foragers. Population increase must have been slow, limited by the birth-spacing of four to five years needed by mothers in foraging societies.
38
If Australian Aboriginal and Bushman models are an indication, language communities would seldom have exceeded a few hundred speakers. Even in the 1970s, when almost all communities in Near Oceania were sedentary farmers, the mean size of language communities was about 2000–3000 speakers (Wurm and Hattori, 1981–83). The earliest settlers probably relied heavily on the rich resources to be found on the seashore (Allen, 2000; Gosden, 1992; Gosden and Robertson, 1991) but people did not remain confined to the coast. During the late Pleistocene, when temperatures were about 4 degrees cooler than today, the valleys in the central Highlands of New Guinea contained extensive grasslands which were home to a considerable mega fauna. Rock shelter and open sites in the central Highlands show human occupation by 30–26,000 BP. People were evidently seasonal visitors to the uplands, not permanent inhabitants, coming to hunt game and harvest pandanus nuts (Evans and Mountain, 2005). After about 18,000 BP, as the climate warmed by several degrees, the upland grasslands were replaced by dense forests, dominated by Nothofagus, making the region less penetrable. Dryland rainforests are not friendly places to foragers, though the forest fringe can be a productive source of food plants and game and therein lies a motive for clearing and burning. At different times in the history of New Guinea possible directions of population spread have been constrained by geographic factors, including changes in climate and sea level. Since the height of the last ice age, around 21,000–18,000 years ago, the coastlines of New Guinea, particularly on the southern side, have fluctuated (Chappell, 2005). The Sahul-Armature shelf, which linked Australia and New Guinea, was gradually flooded by rising seas, with the last land connections (through the Torres Straits) inundated shortly before 8,000 BP. By 6,000 BP rapid changes to the New Guinea coastline ceased, with changes since then largely confined to the progradation or coastal plains and deltas. The low-lying Digul River region was invaded by the sea and inundated at 6,000 BP, as was the delta and narrow floodplain of the Fly River. It appears that most of the swampy Digul lowland has been established over the past six millennia as a result of slow isostatic emergence. These changes have implications for understanding the current distribution of linguistic groups. Early in the post-glacial period climate changes brought major shifts in patterns of vegetation, in sea levels, and available resources. In the New Guinea Highlands landscapes begin to be modified by humans at a number of sites after 10,000 BP with a marked increase from about 5,000 years ago (Hope and Haberle, 2005). There is increasingly strong evidence for some form of agriculture as early as 10,000 BP at Kuk in the Upper
Recent Research on the Historical Relationships of the Papuan Languages
Wahgi Valley (Denham, 2005; Denham et al., 2003; Golson, 1977; Golson et al., f.c.). The main cultivated plants are thought to have been Colocasia taro and bananas. Taro is a lowlands plant but Denham (2002) argues that it had spread naturally into the Highlands by 10,000 BP. The shift from a primarily foraging to a primarily agricultural economy at Kuk may have taken place over many millennia. As to how fast and far agriculture spread in New Guinea the archeological evidence at present says little. There are several sites in the Upper Wahgi Valley with well-dated drainage systems older than 3,000 BP. These remain the only New Guinea sites of this kind with secure dates although there is another early site at Yeni swamp in the lower Jimi Valley with signs of drainage structures at 5,000 BP (Gorecki and Gillieson 1989). Pollen analysis shows that reduction in forests due to burning had also taken place in the Kelala swamp in the Baliem Valley by 7,800 BP, although in the Tari Basin in the southern Highlands of Papua New Guinea it is first evident only at 1,700 BP (Hope and Golson, 1995; Hope and Haberle, 2005). There is as yet no direct evidence that such burning was associated with agriculture. However, in the case of the Baliem Valley, Golson (1991: 487) observes that pollen cores record an almost “continuous vegetation history from beyond 7000 bp to the present, reflecting progressive human impact by way of agriculture through the increasing representation of secondary forest taxa and associated changes … This new evidence from the Baliem is the strongest independent support for the claims of 9000 year old agriculture based on Kuk.” Where the shift to intensive agriculture did occur it must have brought radical changes in patterns of social organization and material culture. Agriculturalists are sedentary, tied to the land they have cleared, tilled, planted, and fallowed. There is potential for faster population growth, larger social units and social hierarchy and for the making of ‘heavy’ artifacts, such as substantial houses, elaborate carvings, and large containers. Language populations tend to become larger and this in turn must have allowed more marriage within the language community (Friedlaender, 1975; 2005). Ethnographic evidence suggests that the shift to intensive agriculture would have occurred faster in certain regions than others, the broad, fertile highland valley floors being among the first. Not all New Guinea societies were farmers even in historic times. Roscoe (2005) points to another route to sedentism and larger social units. New Guinea foragers who occupy favorable aquatic environments where it is possible to gather sago to supply carbohydrates, and to get fish, shellfish, and crustaceans from tidal rivers, lakes, and swamps, tend to live in medium to large villages and to show a degree of hierarchical structure and elaborate visual art forms. Some of the Asmat and Mimika communities
of the southwest coast of New Guinea and the Murik communities of the Sepik basin are examples. None of the domestic animals that were important in Near Oceania at first contact—pigs, dogs, and chickens— were native to the region. They were, however, all part of the Austronesian cultural package in Island SE Asia. There has been vigorous debate over the antiquity of the pig in New Guinea. Did it predate the arrival of Austronesian speakers? Bulmer (1975; 1982) and Allen (1993) report evidence of pig teeth from several sites associated with pre-Lapita dates. The majority view is that the evidence is unconvincing. A parallel debate has taken place over the antiquity of pottery in New Guinea. Pottery sherds in archeological sites in the Sepik-Ramu and Simbai areas that may predate the Lapita horizon are reported by Bulmer (1982) and Swadling et al. (1989). The dating of these materials remains controversial (Spriggs, 1997). The prehistory of Northern Island Melanesia seems to have followed a rather different course from that of New Guinea during the Upper Pleistocene and early Holocene. The large waisted axes found in a number of New Guinea sites were virtually absent from sites in Northern Island Melanesia. Although many of the plants, small mammals, birds, and aquatic animals were the same on both sides of the Vitiaz Straits, New Britain, New Ireland, and Manus had no counterpart to New Guinea’s Pleistocene mega fauna. With no large land animals present hunters must have found it hard going. By about 23,000 BP the Gray Cuscus (Phalanger orientalis) had been imported from New Guinea into the Bismarcks, where it became a significant food source. Spriggs (1997: 62 ff) questions the tenuous arguments advanced by Allen (1993) that early agriculture may have spread to the Bismarck Archipelago in the mid-Holocene. Sweeping changes in technology and life styles in Northern Island Melanesia began late in the 2nd millennium BC. The main catalyst was the arrival of people bearing a distinctive Neolithic culture whose origins lay in Island Southeast Asia. The Lapita cultural complex which first appears in the Bismarck Archipelago around 3,500–3,300 BP (Green, 2003; Kirch, 1997; 2000; Spriggs,1997; Summerhayes, 2000a,b) is very plainly the archeological footprint of Austronesian speakers coming from Taiwan though eastern Indonesia (Bellwood 1997; Bellwood and Dizon, in press; Blust, 1978; 1995a; Pawley, 2002). Carried by their outrigger canoes with sails, and with an economy based on fishing, agriculture, arboriculture, and domestic animals, the Austronesian-speaking Lapita peoples in the Bismarck Archipelago rapidly explored the islands and reef systems of the region. At first they settled mainly on small islands that provided favorable habitats and for several generations maintained a network of social and economic relationships between
39
the framework
widely scattered communities (Summerhayes, 2000a,b). Lapita people appear to have followed this strategy in their initial dispersal throughout Near Oceania. Over the next three millennia, however, their cultural descendants occupied parts of the main islands in the Bismarcks and pockets of the New Guinea mainland. Contact between Austronesian and Papuan speakers transformed the ways of life of all the Papuan-speaking peoples of Northern Island Melanesia and Wallacea, and of some of those in New Guinea (Spriggs, 1997). There was a rise in population densities, a dramatic change in settlement patterns, with large villages appearing, and a very sharp increase in intensity of interaction between regions.
Some Unresolved Questions There are many questions about the prehistory of Near Oceania that are not fully resolved by the archeological record as it stands. Among those for which we might seek relevant evidence in the linguistic record are the following: 1. Archeologists read the archeological record for Near Oceania as indicating continuity of occupation by descendants of the founding populations during the long period between first settlement around 50,000–40,000 years ago and the Austronesian colonization around 3,000 years ago. Is the linguistic evidence compatible with this view or does it indicate that new languages entered Near Oceania from Wallacea during that period? 2. The archeological record suggests that, after the first settlement of Greater Bougainville by about 30,000 BP, there was little or no contact between this region and the Bismarck Archipelago. Is the linguistic record consistent with this view? 3. Archeological evidence is equivocal regarding the extent of contact between New Guinea and the Bismarcks in the late Pleistocene and early Holocene. Are there linguistic traces of population movements in that period? 4. In which directions and when did full-scale agriculture first spread in New Guinea? 5. Was the initial spread of agriculture accomplished mainly by the expansion of farming societies who carried their languages with them or was it mainly a movement of ideas and technology between existing populations? 6. Did agriculture spread from New Guinea to Northern Island Melanesia before the Lapita period? 7. Were pigs, dogs, and pottery present in New Guinea before the Lapita period? 8. What kinds and intensity of interactions typically took place between Austronesian- and Papuanspeaking neighbors, with what cultural, linguistic, and genetic consequences?
40
On the Methods of Historical Linguistics Linguists have used various methods to draw historical inferences about the Papuan languages. The methods are not all of equal reliability. It may be useful to sketch and evaluate the most widely used methods before we begin to discuss these historical inferences.
The Comparative Method The principal method of historical linguistics is simply called ‘the comparative method’. At one level the comparative method is a set of procedures for (i) identifying linguistic residues shared by related languages, that is, cognate elements, retained from the common ancestor and (ii) drawing historical inferences from these residues. However, ‘the comparative method’ is not just a set of procedures. It is also a theory of how particular resemblances and differences among languages come about. Central to the theory is the genealogical (or family tree) model—the assumption that certain languages belong to families that trace descent from a common ancestor. This assumption rests on the fact that languages are typically fairly stable codes, each language being learnt by successive generations of native speakers with gradual change. We can speak of genealogical continuity so long as the line of native speaker transmission is unbroken. Linguistic splitting occurs when a population speaking the same language becomes sharply separated by geographic or social barriers and the isolated daughter communities undergo independent changes, leading eventually to mutual unintelligibility. Successive splits yield a family of related languages. There are certain peculiar facts of language change that make it possible to identify cognate elements and to distinguish these from resemblances that are due to chance or borrowing: (i) sound change (change in the pronunciation of words) is more or less regular across the lexicon of a language; (ii) sound changes are highly constrained (only certain kinds of changes are possible and among these some are rare); and (iii) regular sound changes are irreversible. Over a century of work on a number of language families has shown that related languages typically exhibit a high degree of regularity in sound correspondences. Many of these correspondences reflect structural changes in certain languages, such as the loss of particular phonemes (distinctive sounds) in some or all positions, or the merger of two phonemes in some or all positions, for example, earlier h and s may merge as h, or l and r as r. Many changes are simply phonetic (without changing the number of phonemic contrasts), for example, p may change to f, s to h, t to ts before i, ai to e and au to o. The existence of regular sound correspondences is one of the strongest proofs of genetic relationship.
Recent Research on the Historical Relationships of the Papuan Languages
The sounds that reflect systematic correspondences across languages, and with earlier stages, and the mutations they undergo are broadly comparable to the kinds of genetic markers used by population geneticists. Regular sound correspondences provide a principled basis for reconstructing the sound system, and as much of the lexicon and morphology of the common ancestor as is represented by cognate material in daughter languages. Reconstruction of cognate morphological paradigms (such as systems of personal pronouns, articles, tenseaspect affixes) in turn provide a powerful confirmation of genetic relationship. Subgrouping (determining the sequence of linguistic splits) in a family is done with reference to shared innovations. If a subset of languages shares changes in its sound system, lexicon or morphology apart from other languages this is evidence that they fall in a subgroup. How strong the evidence is depends on how unusual the innovations are and how many of them there are. A family tree diagram schematically represents a sequence of periods of unified development and separate development. Like most models in science, family tree representations oversimplify real events. The genealogical model is not intended to account for all historical resemblances between languages. For instance, it is well known that speakers of a language will take words, phrases, and conceptual structures from other languages. Borrowings can spread over a large area. However, even heavy borrowing usually does not completely obscure the line of genetic continuity in a language. The Germanic heritage of English remains obvious in spite of the huge overlay of loans from French. Historical linguistics has other frameworks besides the comparative method for dealing with the fact that innovations may spread unevenly through a language community.
Dialect Geography Languages are not homogeneous. What may be loosely termed a single language will vary across geographic and social space. Dialect geography is a method for mapping the distribution of linguistic variables and for drawing historical inferences from these distributions and from any geographic or social correlates that may be evident. Certain distributional patterns, for instance, indicate sharp dialect boundaries, others indicate a continuum of intergrading dialects. Other patterns provide evidence of diffusion centers, regions from which innovations spread, or transition areas receiving innovations from two or more different diffusion centers, or relic areas, usually geographically marginal regions that preserve older features.
Lexicostatistics and Glottochronology To apply the comparative method thoroughly, one needs fairly extensive and good-quality language descriptions and takes a long time. Linguists lacking the necessary data and time have often resorted to other methods, faster but less sound, to assign languages to families and to arrive at subgroupings. Some 50 years ago linguists developed a quick and dirty method of subgrouping using a small sample of words, representing either 100 or 200 putatively universal, ‘basic’ concepts. This method was based on the assumption that lexical replacement in such a body of vocabulary will occur at a roughly constant rate over long periods. It was but a small step from lexicostatistical subgrouping to glottochronology, a method for estimating absolute dates for linguistic splits. Estimates of replacement rates were based on a number of cases where written records exist, in some cases extending over 3500 years, for Indo-European, Semitic, and Chinese languages. These indicated a mean replacement rate of about 14 percent per 1000 years in the 100-item list and 19 percent in the 200-item list (Swadesh, 1952; Lees, 1953). Various refinements to the methods of lexicostatistics and glottochronology have been proposed as information about variability has accumulated. Cases were soon found where individual languages show much faster or slower replacement rates than the mean. However, in a large matrix of comparisons strongly deviant languages can generally be detected. Detection is easier when a subgrouping has been arrived at independently by the comparative method. The main value of glottochronology is to give very rough dates for linguistic splits in those cases where the order of splits has been independently determined by the comparative method. Even linguists who decry glottochronology sometimes make use of non-quantitative comparisons which are, in effect, a kind of impressionistic lexicostatistics, when they say, for example, “this family (or subgroup) is about as diverse as Romance (or Germanic or Slavic).”
Difficulties in Tracing Common Origins at Extreme Time Depths As the residue of cognate material diminishes over time the comparative method loses it force. It is no accident that the proto-languages for which detailed reconstructions have been made were all probably spoken less than 7,000 years ago and in most cases less than 5,000 years ago. However, lexical traces of common origin may persist for much longer. Words for certain kinds of concepts in the 100- and 200-item lists tend to be far more persistent than others. A study of Indo-European languages by Kruskal et al. (1971) found that most words on the 200-list
41
the framework
have a half-life of between 1,000 and 2,000 years, that is, they have a 50 percent chance of being replaced by a noncognate word in that time. But 28 words (14%) on the list have a half-life of more than 5,000 years. (These are cases where early Indo-European etyma have been retained as the most common forms to denote their original meaning in about half the languages in the sample.) It follows that members of this set of 28 words have a 25% chance of persisting for over 10,000 years. Fifteen words in the IndoEuropean data were estimated to have half-lives of at least 13,000 years. These consisted of several pronouns, several numerals plus the question words for ‘who’, ‘what’, and ‘how’, plus ‘name’, ‘tongue’, and ‘new’. Seven words were estimated to have half-lives greater than 21,000 years: the numerals ‘two’ to ‘five’, plus ‘I’, ‘we’, and ‘who’. Critics have rightly objected that claims for universals in rates of lexical replacement need to be tested against a range of language families. So far such testing has proved difficult to do satisfactorily. However, it is noteworthy that Austronesian and Trans New Guinea languages show quite close agreement with IndoEuropean in the nature of the words that are most stable. The most stable 20 words in Austronesian (Dyen et al., 1967) include forms for the numerals ‘one’ to ‘five’, the pronouns ‘we’, ‘thou’, and ‘ye’, and ‘name’, and ‘new’, while ‘tongue’ and ‘what’ are in the top 40. The TNG family shows just one important difference: numerals do not figure among the most stable words. If the Indo-European pattern applies to all language families under all social conditions it follows that a significant residue of the original common vocabulary should remain after 10,000 years, in those favorable cases where (a) a proto-language has left many surviving daughter languages belonging to several different high-order subgroups, as with Indo-European, Niger Congo, or Austronesian, and (b) the cognate lexicon has not suffered too much phonological change. On average, about 8 percent of the words in the 200-word list should remain on the list in a single language after 10,000 years. Because languages in different subgroups will retain a somewhat different set of words, the total residue of the 200-word list in such cases should amount to more than 100 words. Indeed, some slight lexical residue should remain even after 20,000 or 30,000 years. In practice however it will usually be very hard to detect residues at time depths of 10,000 years or more. Even if the cognates are recognizable the number of cognate sets will probably be too few for regular sound correspondences to be determined. Thus it may be impossible to distinguish genuine cognates from chance resemblances and borrowings. Statistical methods have also been developed for detecting shared structural residues at great times depths, with the implication that the shared material reflects
42
either common origin or ancient contact (Nichols, 1992). Such a method has recently been applied to Papuan languages of the Northern Melanesia and the Solomons (Dunn et al., 2002; 2005).
Papuan Historical Linguistics: 1950–1986 Papuan historical linguistics is still at a very rudimentary stage of development compared with, say, studies on the Indo-European or Austronesian families. The reasons are several. First, Papuan languages are not one family but many. Second, few historical linguists have sought to apply the comparative method rigorously to any of the families. Third, fewer than 10 percent of Papuan languages are reasonably well described in terms of extensive dictionaries and grammars. Following pioneering studies by Capell in the 1940s (Capell, 1944; 1948) wide-ranging comparative studies of Papuan languages were undertaken by Stephen Wurm and his staff and students in the Australian National University’s Research School of Pacific Studies. Between 1958 and the mid 1970s they collected elementary data on hundreds of languages with the aim of clarifying their genealogical relations. Around the same time the Dutch linguists Anceaux, Cowan, and Voorhoeve began comparative research in Irian Jaya. This research yielded a series of preliminary classifications, which identified some 60 or more distinct Papuan families. Then in the 1970s a number of bold proposals were advanced for lumping many of these diverse small groups into a few large families, using lexicostatistical and typological arguments. The major synthesis was presented in Wurm (ed. 1975). There the number of Papuan families was reduced to 10, with another 10 isolates recognized. The authors generally use the term ‘phylum’ to refer to the highest-order genetic groups. These terms form part of a set used in lexicostatistical classifications to rank subgroups according to percentages of shared cognates; here we will replace ‘phylum’ with ‘family’. The 1975 classification is best known to the wider, interdisciplinary world through two derivative works, the two-volume Atlas of Languages of the Pacific (Wurm and Hattori, 1981–83, Wurm, 1982). The classification in Wurm (ed. 1975) and Wurm and Hattori (1981–83) included two particularly striking claims. One was that almost 500 Papuan languages can be assigned to a single genetic unit, the ‘Trans New Guinea phylum’. If accepted, this would make Trans New Guinea the third largest established language family in the world, after Niger-Congo and Austronesian. The other was that there is an ‘East Papuan’ phylum consisting of all the surviving non-Austronesian languages of Island Melanesia plus Yeli Dnye of the Louisiade Archipelago.
Recent Research on the Historical Relationships of the Papuan Languages
A good many scholars outside the field accepted these proposals uncritically. However, they were not well received by specialists. The evidence adduced for the East Papuan phylum was considered much too flimsy to be accepted. The TNG hypothesis fared somewhat better. All the main specialist reviewers of Wurm (ed. 1975) regarded the hypothesis as unproven, but not without promise (Foley, 1986; Haiman, 1979; Heeschen, 1978; Lang, 1976). Serious reservations were entered by two of the principal contributors to Wurm (ed. 1975), namely McElhanon (1975) and Z’graggen (1975). In terms of number of languages, though not in geographic expanse, the second largest Papuan family proposed in the 1970s was the Sepik-Ramu phylum, a grouping of more than 90 languages spoken in and around the Sepik-Ramu basin put forward by Laycock and Z’graggen (1975). The SepikRamu proposal fell into the same basket at the TNG hypothesis. The critics considered it promising but unproven. The most ambitious lumper in the Papuan field has been the American linguist, Joseph Greenberg, who put forward the Indo-Pacific hypothesis (Greenberg, 1971). This not only assigns all 750 or so Papuan languages to a single genetic stock, but also includes in it the Tasmanian and Andaman Island languages while excluding those of mainland Australia. The Indo-Pacific hypothesis is based on very flimsy evidence and I know of no Papuan specialist who accepts it. However, Greenberg’s list of resemblant forms did contain some material that points to a more restricted grouping of several hundred languages of New Guinea and the Timor region. It is unfortunate he did not separate this material from the dross.
Critiques of the Case for TNG The critics’ objections to the case for the TNG hypothesis can be summarized under several headings. I have discussed these objections in some detail elsewhere (Pawley, 1998; 2005a) and will touch on them only briefly here. (i) Lexicostatistical evidence was considered unconvincing as a basis for positing the TNG family or indeed for positing high-order subgroups in it. Cognate percentages between distant branches of TNG are very low. Wurm (1971: 585) says that the languages in the East New Guinea Highlands stock, the Huon stock, the Central and South New Guinea stock, and the West New Guinea Highlands phylum show an average of 3–7 percent cognation, while these groups and the Southeast Papuan phylum show an average of 2–3 percent. Certain objections can be raised against such low agreements as grounds for claiming genetic relations: We are not dealing with established cognates here but with ‘resemblant forms’, which are possibly cognate. Some or all of the resemblant forms may be due to chance or borrowing.
(ii) Undue weight was given to structural resemblances. It is generally agreed in historical linguistic circles that many kinds of structural features can be quite readily borrowed or lost and thus do not constitute strong evidence for genetic relationship unless they are also associated with cognate morphemes. (iii) The comparative method was not properly applied. Nowhere in Wurm (ed. 1975) is a systematic attempt made to establish regular sound correspondences for cognate sets attributed to proto TNG. Such work is needed to underpin the reconstruction of a phonological system for proto TNG, and so permit specific lexical forms to be reconstructed. (iv) Accumulated borrowing and relatively rapid lexical replacement in Papuan languages may have made it impossible to use the comparative method effectively to determine deep genetic relationships. Foley refers to demographic and social factors that appear to make Papuan languages particularly difficult subjects for comparative lexical studies. Papuan language families are small and are generally spoken in small areas. The languages are usually contiguous, and have been so for millennia. None of the particular historical and geographical patterns necessary for the smooth application of the com parative method obtain in Papuan languages. Rather, Papuan languages normally exhibit a pattern of enormous cross-influence in all areas; so in no sense can the assumption that the daughter languages develop independently be taken as viable in this context. As the comparative method, with its sorting of cognates from borrowing, is deeply grounded in the family tree model, its application to Papuan languages is no mean problem …. (Foley, 1986: 209–10) While ‘basic’ vocabulary, in general, is less prone to borrowing than ‘cultural’ vocabulary, the difference is only one of degree. Comrie (1986; 1990), among others, has documented extensive borrowing in basic vocabulary among Papuan languages that are at best only very distantly related. Foley questions the feasibility of applying the comparative method to the lexicon of Papuan languages except when the languages are quite closely related, say, at the level of the members of the Germanic family, or the Romance family—with a common ancestor spoken no more than about 2,000 years ago. After comparing basic vocabulary lists from languages representing several Highlands groups: Gorokan, Kainantu, Huon, and Engan from Papua New Guinea, and Dani and Ekagi from Irian (Papua), Foley (1986: 262) concludes that “[i]n general, where such distant relationships are concerned, the number of cognates is wholly insufficient to establish genetic affiliation.”
43
the framework
(v) There was a failure to exploit morphological paradigms. Wurm and his ANU associates were holding an ace but they did not play it effectively. The ANU researchers noted that widespread cognates in certain independent pronoun forms allow the reconstruction of a near complete paradigm (Voorhoeve, 1975; Wurm, 1975). But although approximate reconstructions of several forms in this paradigm were made the supporting evidence was neither assembled nor analyzed in a systematic fashion. Foley (1986: 258–61) found a small number of formal resemblances among personal pronouns and verbal affixes shared by languages of the well-defined Gorokan and Huon groups with the geographically Dani and Ekagi languages, though he concluded that Engan languages show no convincing resemblances to the other languages in these respects. (vi) Is the family tree model an appropriate means of representing historical relations among many Papuan languages? Some of the key chapters in Wurm (ed. 1975) took an equivocal stance on this point. The following statement indicates that the authors regarded some languages—those that are typologically and lexically most divergent from the rest—as being members of the family only in a secondary sense, as a result of non-native speakers adopting a pidgin form of TNG in place of their mother tongue. It appears that much of the Trans New Guinea Phylum area may have originally been occupied by a number of probably unrelated earlier languages, and that the inter-relationship of many of the presentday Trans-New Guinea Phylum languages is, in a way, secondary, or partial and fractional, in nature. The presence of the older, different languages upon which the Trans-New Guinea Phylum languages appear to have been superimposed is noticeable in the form of substrata of varying strength throughout the greater part of the Trans-New Guinea Phylum. (Wurm, Voorhoeve and McElhanon, 1975: 300; my italics) In his review of Wurm (ed. 1975), Ranier Lang was sharply critical of the weight given to substratum influence as an explanation of diversity within TNG languages. He argued that: what evidence we have of population movements in Papua New Guinea is of a kind that does not allow for substrata. Populations have been displaced in recent history … through either of two events (or a combination of the two): (a) natural disasters such as volcanic eruption, an earthquake, or drought and/or frost have driven populations from their home ground; (b) warfare has had the same effect. When they have left their home ground they have
44
either moved into virgin bush to carve out an entirely new existence for themselves ... or they have taken refuge with allies, in which case they have been absorbed into the host group, thus giving up their language and adopting that of their hosts. It would seem that the natural fragmentation of the country and the social conditions (partly brought about by geographical factors) would be much stronger determinants of linguistic diversity than substratum influence. But how the social conditions bring about linguistic changes, of this we know precious little in the New Guinea area. The sophisticated sociolinguistic research has just not yet been carried out. (Lang, 1976: 77–8) The critics were well justified in being highly skeptical of the case for the TNG hypothesis presented in Wurm (ed. 1975). The arguments were marred by serious methodological flaws. Having assembled various bits of suggestive evidence the research group failed to carry the job through by systematically applying the comparative method. In fact, they showed a mysterious reluctance to do so, preferring to rely heavily on typological criteria.
Recent Research in Papuan Historical Linguistics For a couple of decades after 1975 little new work was done on Papuan historical linguistics. An exception was Foley’s influential general survey of Papuan linguistics (Foley, 1986), which contains a summary of his pioneering application of the comparative method to two Papuan groups: Lower Sepik and Eastern Highlands. That book and a later short survey (Foley, 1992) contained the sobering message that we need to recognize some 40–60 separate Papuan families that have not been convincingly shown to be related by the comparative method—in other words, back to the pre-1970 situation. In the mid 1990s researchers began to return to this daunting field. At the Australian National University, Malcolm Ross, Meredith Osmond and I began to sift through the growing body of descriptive data on two of the most extensive families proposed by Wurm and his associates: the ‘Trans New Guinea phylum’ and the ‘East Papuan phylum’. At the University of Sydney, William Foley and Mark Donohue and their students undertook comparative studies of several language families of the Sepik-Ramu basin and the north coast of New Guinea, building on work Foley had begun in the 1980s. More recently, at the Max-Planck-Institute in Nijmegen Michael Dunn, Angela Terrill, and Ger Reesink have begun a historical study of the Papuan languages of Northern Island Melanesia. And there have been other comparative projects elsewhere.1
Recent Research on the Historical Relationships of the Papuan Languages
Identifying Families The most extensive classification of Papuan languages based on a single class of evidence is that of Ross (2000; 2001; 2005). Pronoun paradigms are widely regarded as among the most reliable diagnostics of genetic relatedness. For this reason and because data on free pronouns are available for almost all Papuan languages Ross compared 605 languages, using cognation in pronoun forms as the main basis for recognizing language families and for subgrouping. He concluded that the pronouns indicate these languages fall into some 23 families that cannot on present evidence be related to each other, plus 9 or 10 isolates. For each of the larger families he sought to determine a sequence of innovations in pronoun forms and categories that would yield subgroups. Ross’s study thus indicates that the Papuan languages show more deep genetic diversity than was recognized by Wurm (ed. 1975) and Wurm and Hattori (1981–83) but less than was proposed by Foley (1986).
Northern Island Melanesia (figure 3.2) Ross’s pronoun study gives no support for Wurm’s East Papuan phylum. Instead he finds eight distinct genetic units, including five families, which show a few noteworthy typological similarities, such as a masculine/feminine distinction in 3rd person pronouns (Ross, 2001; Terrill, 2002; Dunn et al., 2002; Wurm, 1982). The Papuan
languages of New Britain are divided into an East New Britain family (the close-knit Baining group, Taulil, and Butam), a West New Britain family (Anêm and Ata), and two isolates, Sulka and Kol. Only one Papuan language, Kuot, survives in New Ireland, although in their phonological systems some neighboring Austronesian languages show what seems to be a Kuot-like substratum. Ross noted a couple of suggestive resemblances between the pronouns of West New Britain family and those of Yeli Dnye, of the Louisiade Archipelago. The eight Papuan languages spoken in Bougainville fall into two families, North Bougainville (including Rotokas and Konua) and South Bougainville (Nasioi, Nagovisi, Motuna, and Buin). Four Papuan languages (Bilua, Baniata, Lavukaleve, and Savosavo) are found in the central Solomons, each separated from its nearest Papuan neighbor by a number of Oceanic Austronesian languages. There is some evidence indicating that these four languages are distantly related to one another, forming a Solomons family. In the Santa Cruz group, in the eastern Solomons, there are three languages whose status, as Austronesian or Papuan, has been disputed (Lincoln, 1978; Wurm, 1978). Recent work has yielded evidence favoring the view that they are Austronesian (M. Ross, p.c.). In any event, given strong evidence that the Santa Cruz group was not settled until about 3,000 years ago when bearers of the Lapita culture arrived, the ancestors of the Santa Cruz languages must have arrived either with or later than the Lapita colonists.
Figure 3.2 East Papuan languages.
45
the framework
Figure 3.3 Map of the Papuan language families of New Guinea other than Trans New Guinea (see table 3.1 for language names).
New Guinea (figure 3.3 and table 3.1) There are 18 Papuan families represented on the New Guinea mainland. The most spectacular linguistic diversity, unsurpassed anywhere else in the world, is found in north New Guinea, from the Bird’s Head to Madang Province. I will refer to this region as ‘north New Guinea’. It contains no fewer than 15 families plus several isolates. In Papua New Guinea the densest concentration of families is found in a continuous area spanning three provinces: Sandaun, East Sepik, and Madang. In West Papua several families are located on the Bird’s Head and, east of the Bird’s Head, to the north of the main highlands range, from Cenderawasih Bay to the Mamberamo River. The putative Sepik-Ramu family of more than 90 languages (Laycock and Z’graggen, 1975) is not supported by William Foley’s recent comparative study (Foley, 2005) or by Ross’s study of pronouns. Foley assigns these languages to three unrelated groups: Sepik, Lower SepikRamu, and Yuat. He argues that the large Sepik family of nearly 50 languages has its greatest diversity, and therefore its original dispersal center, upriver from Ambunti. Ross (2005) also recognizes the Sepik and Yuat groups but divides Lower Sepik-Ramu into two possibly unrelated groups: Lower Sepik and Ramu, as well as treating Taiap as an isolate. However, he follows Foley in finding some slight evidence for uniting Lower Sepik and Ramu. Ross concludes that the distribution of the Ramu and Lower Sepik
46
Table 3.1 A Tentative Listing of Papuan Families in New Guinea (see figure 3.3) 1 ‘extended West Papuan’ (?) 1a West Papuan 1b Bird’s Head, Sentani, Burmeso, Tause 1c Yava 2 Mairasi 3 East Cenderwasih Bay 4 Lakes Plain 5 Orya–Mawes–Tor–Kwerba 6 Nimboran 7 Skou 8 Border 9 Left May–Kwomtari 9a Kwomtari 9b Left May 10 Senagi 11 Torricelli (three separate areas) 12 Sepik 13 Ramu–Lower Sepik 13a Lower Sepik 13b Ramu 14 Yuat 15 Piawi 16 South–Central Papuan 16a Yelmek–Maklew 16b Morehead-Upper Maro 16c Pahothri 17 Eastern Trans–Fly
Recent Research on the Historical Relationships of the Papuan Languages
languages indicates that their diversification predated the regression of the inland sea some 5,000 years ago. As the silt from the Sepik delta filled up this sea Lower Sepik speakers progressively followed the river to the coast. The Torricelli family proposed by Laycock (1973a; 1975) is supported. It consists of about 47 languages. Most are located in the Torricelli and Prince Alexander Ranges between the Sepik River and the north coast. A second group of Torricelli languages lies to the west and south of the Murik Lakes and a third, small group is spoken on the coast in NW Madang Province, separated from its sisters by languages of the Ndu branch of the Sepik family. Ross, following Layock (1965) and Foley (1986) concludes that Ndu languages have expanded north from around the Chambri Lakes and driven a wedge into the Torricelli family. Several smaller families, each with fewer than 10 languages, have been identified. These include Sko (spoken on the north coast around the Papua New Guinea–Papua border), which may in fact be two families, Kwomtari (northwest part of Sandaun or West Sepik Province), Left May (south of Kwomtari around the May River, a tributary of the Sepik), and Amto-Musian (between Kwomtari and Left May). The West Papuan family comprises about 24 languages spoken at the western end of New Guinea, on the northern part of the Birds’ Head, on Yapen, and in North Halmahera. Slight evidence is found for extending the West Papuan group to include the East Bird’s Head, Yawa, the Sentani group, and a few other isolates from northwest New Guinea. At the western end of New Guinea there is the small Geelvink Bay family, spoken on the coast of Cenderawasih (formerly Geelvink) Bay and the small East Bird’s Head family. In addition there are several other very small groups and isolates scattered about New Guinea, especially in the Sepik area. If much of the diversity has developed in situ it follows that many of the North New Guinea families are related but diverged so long ago that the signals have faded. There are a few tantalizing lexical resemblances among certain non-TNG families of North New Guinea, and between certain of these and TNG. Such resemblances may be interpreted as due to chance or diffusion or as the faint signal of very remote common origin (Reesink, 2005). Here is a matter for future research.
The Trans New Guinea Family Ross and I have examined a range of evidence to do with the TNG hypothesis. We conclude that the hypothesis is valid (Pawley, 1995; 1998; 2001; 2005a; Ross, 1995; 2000; 2005), although the membership is a bit smaller than that
posited in Wurm (ed. 1975). The main evidence for TNG is as follows: (i) Some 200 putative cognate sets, nearly all denoting so-called basic vocabulary, which are represented in two or more major subgroups (Pawley, 2005a, n.d.). This is a very small number compared to what has been compiled for proto Austronesian (between 1,000 and 2,000) or its firstorder branch proto Malayo-Polynesian (over 4,000) (Blust, 1995a, 1995b). The modest size of the proto TNG list can be attributed in part to the shortage of good dictionaries of TNG languages. However, I believe it is also due in part to the very considerable age of the family. (ii) A body of regular sound correspondences, based on (i), which has allowed a good part of the proto TNG sound system and its development in a sample of daughter languages to be reconstructed (Pawley, 1995; 1998; 2001). (iii) Systematic form-meaning correspondences in the personal pronouns, permitting reconstruction of virtually a complete paradigm. In Ross’ sample of 605 Papuan languages 311 showed one or more reflexes of proto TNG pronouns and another 36 were assignable to TNG on other grounds. (iv) Widespread resemblances in fragments of certain other grammatical paradigms (Pawley, 2005a, n.d.; Suter, 1997). In addition, the distribution of certain striking structural resemblances noted by Wurm, Voorhoeve and McElhanon, 1975) has been more precisely charted, and shown to correlate rather closely with the distribution of TNG languages. While such structural evidence cannot be primary grounds for positing a genetic stock it carries some weight as corroborative evidence. In the account of TNG given in Wurm, Voorhoeve and McElhanon. (1975) some 256 languages were accepted as core TNG and another 235 were assigned to the family as ‘marginal’ members. Ross accepts several groups of these ‘marginal’ languages as members of TNG, on the grounds that they meet criterion (iii), e.g. the Papuan languages of Timor, Alor and Pantar, West Bomberai, Gogodala-Suki, and Kairi. Although the Marind, Kiwai, and Inanwatan groups are provisionally assigned to the TNG family on lexical grounds, the evidence at present is slender. In two areas of mainland New Guinea there is a concentration of small groups and isolates whose TNG status remains unclear. These are (1) the area north of the central highlands from the Mamberamo River and the eastern border of the Geelvink Bay family in Papua (West Irian) to the western part of Sandaun Province in Papua New Guinea (this area contains such problematic groups as Kwerba, Lake Plains-Tor,
47
the framework
Border, Nimboran, Sentani, and Kaure); (2) the Gulf of Papua area, covering most of the Gulf Province and the adjacent coastal part of the Western Province of PNG. Clouse (1997) has given strong arguments for excluding the Lakes Plain languages of Irian Jaya from TNG, reassigning them instead to the Geelvink Bay family. In summary, there are secure grounds for identifying a core of TNG languages numbering over 350 with another 100 or so languages less securely assigned to TNG. Strictly speaking this should be called TNG IV, to distinguish it from three earlier proposed groupings with the same name. Grouping of Trans New Guinea Languages (figure 3.4). I turn now to the internal classification of TNG. At present the family tree has a structure that is consistent with a fairly rapid initial dispersal of TNG languages. Some 20 or so subgroups have been identified which cannot be assigned to any grouping lower than Trans New Guinea itself. Most of these have fewer than 15 member languages. Only a few large, internally diverse subgroups are supported. The most secure of these are Madang (about 100 languages) and Finisterre-Huon (about 70 languages). Less secure are the putative Southeast Papuan, West Papuan, and West Bomberai-Timor groups. There is a fair number of subgroups that are roughly comparable in
internal diversity to the Germanic and Romance groups of Indo-European, for example, Chimbu-Wahgi, Engan, Goroka, Kainantu, Binandere, Ok, Dani, Awyu-Dumut, and Asmat-Kamoro. These appear to be the result of dispersals some 2,000 to 3,000 years ago and may reflect population expansions due to some technological advances in that period that enabled people to occupy new habitation zones. The following is a selection of subgroups that seem reasonably secure. Much of the evidence for these is based on innovations in the personal pronouns. Madang With about 100 members, Madang is by far the largest subgroup of TNG that can be justified in terms of shared innovations. The group is defined by a number of innovations, the most important being the replacement of the proto TNG 1st, 2nd, and 3rd person singular pronouns *na, *ŋga and *ya by pMadang *ya-, *na-, and *nu(Pawley, 1998; Ross, 2000). The extreme structural and lexical diversity found across its major branches strongly suggests that the Madang group is at least four or five millennia old. It appears that the Madang group consists of four major branches: South Adelbert Range, Croiselles (corresponding in part to the ‘North Adelbert Range’ group posited in Z’graggen (1975)), Rai Coast, and Kalam-Kobon
Figure 3.4 Putative subgroups within the Trans New Guinea family.
48
Recent Research on the Historical Relationships of the Papuan Languages
(Ross, 2000; Pawley, 2005b) The Kalam-Kobon group and Gants had previously been included in a subgroup with the East New Guinea Highlands micro-phylum (Wurm ed., 1975). Isabi is now excluded, reassigned to the Goroka group. The reassignment of Kalam-Kobon and Gants to the Madang subgroup is of special interest in that it points to two cases of languages moving from the Middle Ramu lowlands into the northern fringe of the Central Highlands. Finisterre-Huon The second largest TNG subgroup at present is Finisterre-Huon. This has perhaps 70 member languages, depending on how one deals with dialect chains (McElhanon, 1975). The group is internally very diverse, lexicostatistically; some of its members share less than 10 percent of basic vocabulary cognates. Ross (2000) notes a possible innovation in the 1, 2, 3 plural pronouns, all of which end in -n, marking plural number. It is likely that members of the Finisterre-Huon group will turn out to share several other distinctive innovations. Kainantu-Goroka Two well-defined groups are located in the Eastern Highlands, now generally termed the Kainantu and Goroka groups. These have long been tentatively placed together in a subgroup. Foley (1986) notes a close fit between Kainantu and Gorokan verbal affixes denoting undergoer and actor, and points to a probable pronominal innovation, in which earlier *ni ‘1st pl.’ is replaced by p-Kainantu-Gorokan *ta. Ross (2000) notes three probable shared innovations: (1) replacement of proto TNG *ni, *nu ‘1st pl.’ by *ta[za], (2) replacement of proto TNG *ŋgi or *ja ‘2nd pl.’ by *ta-na, (3) development of genitive pronouns ending in -i. Chimbu-Wahgi This is an uncontroversial group centered east and south of Mt Hagen, in the Wahgi, Nebilyer, and Kaugel Valleys, and extending north of the Sepik-Wahgi Divide into the Jimi Valley. Ross (2000) notes an innovation in the pronouns, the use of *im ‘inclusive marker’, obligatory in 1st inclusive and optional in 2nd and 3rd person plural. Chimbu-Wahgi contains perhaps 12 languages, although the situation is complicated by the fact that certain ‘languages’ are made up of extensive dialect chains. The best-known members of this group are probably Kuman, Middle Wahgi, Sinasina, and the Medlpa-Hagen dialect chain. The subgroup parent was spoken perhaps 2,000 to 3,000 years ago. Engan This is an uncontroversial group consisting of several languages spoken in and near the Mt Hagen Range west of Mt Hagen. There is a northern subgroup that includes Enga, Ipili, Iniai, and Lembena and a southern subgroup that includes Sau, Huli, Mendi, and Kewa. Ross finds that proto Engan replaced proto TNG *ŋga ‘2S’ by *ne(ke), and *ŋgi ‘2PL’ by *ni(a) and that Engan languages
have considerably modified the proto TNG pronoun paradigm in other ways. There is a probable subgroup that includes at least the Ok, Awyu-Dumut, and Asmat-Kamoro groups, which I will call Central and Southwest TNG. It is centered in the central ranges around the Papua–PNG border, including the Star Mountains, and Thurnwald and Victor Emmanual Ranges, and in the lowlands to the southwest of this. This group should not be confused with an entity called Central and South New Guinea, proposed by Voorhoeve (1968), who put it forward as a separate language family, not as a subgroup of TNG, which was not then recognized. Ross finds no evidence in the pronouns for such a group. Awyu-Dumut and Asmat do, however, share a rounding of the vowel in the proto TNG 1st and 2nd singular pronouns *na and *ŋga. It is not clear whether this change is independent or a retention from a common interstage. Ok and Marind both distinguish 3S masculine and feminine: proto Ok *ya, proto Marind -ye- ‘3sM’, proto Ok *u-, proto Marind *-u- 3sF. Is this a shared innovation or a retention of an old TNG feature that has been lost in most branches? Ross inclines to the latter view. Western New Guinea The Dani languages, Wano, and the Wissel Lakes languages share an innovation whereby proto TNG *na ‘1S’ is replaced by reflexes of an intermediate form *ani ‘1S’. They also reflex the *i grade of the nonsingular pronouns, i.e. *ni and *ŋgi. Ross tentatively groups them together under the heading ‘Western New Guinea’. West Bomberai-Timor The West Bomberai and TimorAlor-Pantar groups share two probable innovations in pronouns. Both reflect *bi (or *ba) ‘1P’, replacing proto TNG *ni or *nu. In W. Bomberai the reflexes denote 1st exclusive plural, in contrast to *in or *ni, 1st inclusive plural. In the Timor group the reflexes denote the opposite, 1st inclusive plural, while *in or *ni denotes 1st exclusive plural. Both groups also show metathesis of proto TNG *na ‘1s’ and *ni ‘1P’. Ross (2000) tentatively includes West Bomberai-Timor along with Western New Guinea in a West Trans New Guinea Linkage. Southeast Papuan Ross finds one piece of evidence for a widespread and internally diverse Southeast Papuan group, consisting of the Dagan, Mailuan, Yareba, Manubaran, Kwalean, and Koiari subgroups. This is the replacement of proto TNG *ŋgi ‘2 PL’ by *ya. The Binandere and Goilalan groups are excluded from Southeast Papuan. On the Location of Proto TNG. It is clear that TNG is predominantly a family of the central cordillera, which runs almost the full length of New Guinea. Can we say any more than that proto TNG was very probably spoken somewhere in these highlands? Measured in terms of the
49
the framework
density of established high-order subgroups, the region of greatest diversity in TNG is that region of Papua New Guinea between the Strickland River and the Eastern Highlands Province, together with Madang and Morobe Provinces. The center of gravity of this area is in the Eastern Highlands, Chimbu, and Western Highlands Provinces. It is safe to say that the area corresponding to these three provinces was a very early area of TNG expansion. Whether it was the original dispersal center is another matter. We do not know which of the many high-order subgroups are first-order subgroups. Without this knowledge we cannot rule out the possibility that the primary split occurred elsewhere, e.g. in the highlands of Papua (Irian), or in the Madang and Morobe Provinces. The high degree of diversity within the Madang subgroup points to a very early TNG presence in the eastern half of Madang Province. Movement into the Timor region was probably well after the initial breakup of proto TNG as the Timor-Alor-Pantar group appears to be reasonably homogeneous. On the Dating of the TNG Dispersal. Approximately when did proto TNG break up? The types of evidence that bear on this include: (1) the degree of diversity in basic vocabulary between high-order subgroups of TNG languages, (2) the dating of certain archeological or geomorphological events that can be correlated with linguistic events, (3) relative chronologies indicated by borrowing in relation to subgroups. It is instructive to compare the lexicostatistical diversity within TNG to that of Indo-European and Austronesian, two families whose chronologies are fairly well established. The major Western branches of Indo-European (Celtic, Romance, Germanic, and BaltoSlavic) are generally regarded as separating not later than 5,000 to 6,000 years ago. These subgroups generally converge at around 15–25 percent of cognates in the 200-item list. In the Austronesian family the center of genetic diversity is in Taiwan (Blust, 1999). The firstorder groups found in Taiwan generally converge with the first-order Malayo-Polynesian group at between 15 and 25 percent. Archeology dates the Austronesianassociated Neolithic expansion out of Taiwan and into Island SE Asia as probably beginning around 4,000 BP. The breakup of Proto Austronesian is likely to be no later than that date (Bellwood, 1997; Bellwood and Dizon, in press; Pawley, 2002). Lexicostatistical diversity in TNG is far greater than in either Indo-European or Austronesian. Indeed the largest branch of TNG so far identified, Madang, probably exhibits greater lexical diversity among its major subgroups than either Indo-European or Austronesian. As was noted earlier, pairs of languages from geographically widely separated subgroups of TNG generally show less
50
than 7 percent cognates in basic vocabulary and in some cases less than 3 percent, i.e. below the threshold of chance. If we take Indo-European and Austronesian as yardsticks, a date of between 7,000 and 12,000 BP date for the breakup of proto TNG is compatible with the degree of internal diversity found in TNG. The question arises whether all branches of TNG languages have consistently replaced basic vocabulary at a faster rate than most Indo-European and Austronesian languages have. It is reasonable to suppose that very small language communities will tend to change their basic vocabulary faster than large communities because innovations can more readily spread through a small community. However, I believe it is unlikely that all TNG branches have changed their vocabulary twice as fast as Indo-European or Austronesian. In the absence of written records our only checks on deep glottochronological dates are correlations with archeologically or geologically dated events. There is some tenuous evidence of type (2). The dates for early agriculture at Kuk are compatible with an early stage of TNG being spoken by the occupants of this site, a point we will return to later. The possible directions of spread of TNG languages were at different times constrained by a number of geographic factors, including sea level shifts, as noted earlier. TNG languages are now spoken in the Digul lowlands and in the Fly floodplain, regions that were flooded by the sea around 6,000 years ago. It is noteworthy that these languages belong to two subgroups, Asmat-Kamoro and Awyu-Dumut, that appear to be internally more homogeneous than groups of comparable extent in other regions. These groups probably represent relatively recent (within the past three millennia) expansions of TNG languages into areas of swampy land that were previously below sea level. There is a little evidence of type (3). In Madang Province we can infer that TNG speakers already occupied the coastal regions when Austronesian speakers arrived. Among TNG languages of this area there is evidence of early borrowings from Austronesian. Ross (1988: 21) has shown that some of these borrowings show phonological features that associate them with a stage of Oceanic that predates the more recent (1,000–2,000 BP) settlement of Oceanic languages along the north coast west of Vitiaz Straits. There are a few resemblances between TNG and certain non-TNG families of North New Guinea that may be interpreted either as diffusion or as the faint residue of very remote common origin. If much of the diversity has occurred in situ it follows that many of the North New Guinea families are related but that they diverged so long ago that the signals have faded.
Conclusions Let us return, finally, to the lists of questions that arose from a survey of the archeological literature on Near
Recent Research on the Historical Relationships of the Papuan Languages
Oceania and ask whether linguistics can confirm, contradict, or add to certain conclusions derived from the archeological evidence.2 1. In the period between initial settlement some 50,000–40,000 years ago and the Austronesian colonization around 3,000 years ago, did new languages and populations from Wallacea become established in New Guinea and Northern Island Melanesia, or vice versa? With two rather trivial exceptions none of the Papuan families have no known relatives outside Near Oceania. The exceptions are branches of the TNG and West Papuan families of New Guinea that came to be spoken in the Timor and Halmahera regions. Given the strong likelihood of language family extinctions in Wallacea after the Austronesian Diaspora, the lack of external relatives does not completely rule out the possibility of linguistic movements through Wallacea into Near Oceania after initial colonization. Indeed it would be surprising if there were none. However, the weight of the evidence strongly suggests that most or all of the surviving Papuan families are the outcome of in situ
diversification that began in Near Oceania in the late Pleistocene. This conclusion does not imply that all the Papuan families stem from a single movement into Near Oceania. There are at least two regions of extreme diversity, geographically sharply separated, and each could have had quite separate settlement histories. But within each region the most economical hypothesis is that the families diverged in situ. The northern third of New Guinea, from the Bird’s Head to the Sepik-Ramu basin, is probably the most linguistically diverse part of the planet (figure 3.5). According to Ross’ classification, no fewer than 16 unrelated language families, along with several isolates are found in this area, which is no larger than Great Britain. The south central region of New Guinea, which contains perhaps four families, is a smaller region of high diversity. The second most diverse region of Near Oceania is Northern Island Melanesia. Ross’s classification distinguishes six unrelated language families and about three
Figure 3.5 Shoreline of the Sepik-Ramu ‘inland sea’ ca. 6,000 BP superimposed on a modern language map. Sources: Swadling and Hide (2005), Ross (2005).
51
the framework
isolates, in a total land mass about the size of Ireland. And 3,000 years ago Northern Island Melanesia probably contained many more Papuan languages it does now. Whereas this region now harbors about 150 Austronesian languages (all belonging to the large Oceanic subgroup) only 20 or so Papuan languages survive there. None are present in the Admiralty Islands and only one in New Ireland. The Papuan languages of Northern Island Melanesia have probably had separate histories from those of New Guinea since the late Pleistocene and possibly since people first reached Northern Island Melanesia. 2. The archeological record suggests that, after the first settlement of Greater Bougainville by about 30,000 BP, there was little or no contact between this region and the Bismarck Archipelago. Is the linguistic record consistent with this view? It is The Papuan families of New Britain show no clear evidence of a genetic relationship with those of Bougainville and the Solomons. There are faint signals, in the form of grammatical structure, of a very distant relationship between most of the North New Guinea families (Dunn et al., 2005). As the time it takes for signs of common ancestry to fade out completely can vary from just a few millennia to perhaps 40,000 years, depending on the size of the family and other variables, it is impossible to say more than that the common origins, if such they be, probably go back to the late Pleistocene. 3. Archeological evidence is equivocal regarding the extent of movement or contact between New Guinea and the Bismarcks in the late Pleistocene and early Holocene. Are there linguistic traces of population movements in that period? One kind of evidence for such movement would be the existence of one or more families of Papuan languages having branches both in New Guinea and in Northern Island Melanesia. There were no such families at first European contact. One member of the TNG family is spoken on Umboi Island in the Vitiaz Straits. But none are spoken in New Britain itself. The absence of Papuan families straddling the Vitiaz Straits suggests that in preLapita times population movements across the Strait were in small numbers too small to allow linguistic colonies to become established. However, it must be conceded that language replacements could have occurred on either side of the Vitiaz Straits, obscuring earlier connections. 4. In which directions and when did full-scale agriculture first spread in New Guinea? Was the initial spread of agriculture accomplished mainly by the expansion of farming societies who carried their languages with them or was it mainly a movement of ideas and technology between existing populations? In New Guinea only one of the Papuan language families is very extensive. All the inhabited valleys of
52
the central highlands of New Guinea, from the neck of the Bird’s Head to Southeast Papua, are occupied by members of the Trans New Guinea family. TNG languages also dominate several smaller lowland and mountainous areas to the north and south of the central highlands. It is a reasonable supposition to link the initial dispersal of TNG languages with early agricultural systems of the central Highlands of PNG for four reasons: (a) the location is about right. The central highlands of PNG are (on present evidence) within the region where TNG has its greatest concentration of high-order subgroups. (b) The timing is about right. The TNG family looks to be somewhere between 7,000 and 10,000 years old. It was noted earlier that most of the highland valleys were heavily forested in the early Holocene and may have had no permanent populations, or only very small ones, before about 10,000 BP. (c) The TNG family is exceptional among the families of Papuan languages in its wide geographic distribution and large membership. (d) It seems unlikely that the TNG family would have achieved its present remarkable distribution unless its speakers possessed some cultural advantages that enabled them to build up populations that could (i) expand fairly rapidly along the central cordillera of New Guinea and (ii) maintain continuous habitation of the major highland valleys. Bellwood (1996; 1999; 2004) and Renfrew (1992; 2002) have argued forcefully that the spread of large language families cannot be accounted for simply in terms of the diffusion of a new technology across existing populations. Rather, the languages are carried by migrating populations, and successful, rapid, large-scale expansions are enabled by cultural advantages. It would be nice to be able to support these arguments with proto TNG lexical reconstructions associated with agriculture. At present a term for ‘taro’ (something like *ma) is about the only lexical reconstruction that can be tentatively attributed to early TNG because of its wide distribution. But as the term stands alone, instead of being embedded in a full terminology for parts of the plant and practices associated with its cultivation, diffusion cannot be ruled out. I know of no widely distributed cognate sets for other names of plants and their parts or for implements and processes associated with their cultivation. However, as yet no historical linguist has undertaken a thorough, New Guinea-wide search for cognates in cultural domains (however, see the work of an anthropologist (Hays, 2005)). Why did TNG languages come to dominate the Highlands but not the northern third of New Guinea west of the Ramu River? The linguistic evidence makes it plain that certain parts of north New Guinea had been continuously occupied by speakers of various non-TNG families since before the TNG expansion but linguistics
Recent Research on the Historical Relationships of the Papuan Languages
provides no clues as to whether speakers of these nonTNG languages had agriculture or some other cultural or genetic adaptations that allowed them to hold their ground. The consensus has been that taro and bananas were first domesticated in the lowlands (see Denham, 2002, for a different view) and it makes sense to posit a north New Guinea domestication center (Swadling and Hide, 2005). 5. Did people move beyond Greater Bougainville into the central and southeast Solomon Islands before 7,000–6,000 years ago? Linguistic evidence might bear on this question in the following way. Today only four Papuan languages are spoken in the main Solomons chain east of Bougainville (a fifth became extinct recently). If it could be shown that some or all of these are distantly related to languages spoken on Bougainville this would be evidence of population movements between Bougainville and the central Solomons a few thousand years ago. However, no convincing evidence for a genetic relationship has so far been forthcoming. The absence of a demonstrable relationship between any Bougainville and any Solomons language then is evidence that the Solomons languages have been there for long enough for all lexical traces of common origin with Bougainville to disappear. How many millennia would be needed for that? Given that we are dealing with very small families it may be that five or six millennia would be sufficient. Also relevant is the degree of diversity within the putative Solomons family. This appears to be so great that the existence of the family is controversial (Dunn et al., 2002; Ross, 2001; Todd, 1975)— a situation consistent with the assumption that languages of the Solomons family have been spoken moved east of Bougainville at least 6,000–7,000 years ago and possibly much longer. 6. Were pigs, dogs, and pottery present in New Guinea before the Lapita period? I have not examined the large electronic database of animal and plant names that Terence Hays has compiled for New Guinea languages. The linguistic evidence, in the form of widespread loans from Austronesian, is consistent with the view that pigs were not initially part of the TNG farming complex but spread after the Austronesians reached New Guinea some 3,000 years ago. In many TNG languages the primary term for ‘pig’ is a borrowing from an Austronesian source or sources. Reflexes of proto Oceanic *boRok ‘pig’ are found widely in Papuan languages of north New Guinea, from at least the Huon Gulf to Madang Province, and in the eastern Highlands. The widespread borrowing of the Austronesian name suggests that the animal itself was imported together with the name. The history of Oceanic (Austronesian) terms for ‘dog’ is discussed by Hudson (1991), who concludes that there is no secure proto Oceanic reconstruction possible and
that this is consistent with archeological indications that dogs arrived in Oceania only about 2,000 years ago. A contrary view is argued by Bulmer (1991) and KolerMatnick et al. (2003). Comparative study of potting terms in Papuanspeaking communities of north New Guinea who make pottery has not been undertaken. Proto Oceanic had several terms for kinds of earthenware pots (Ross, 1996a). 8. What kinds and intensity of interactions typically took place between Austronesian- and Papuan-speaking neighbors, with what cultural, linguistic, and genetic consequences? In New Guinea itself the overall impact of Austronesian languages has been relatively slight. Except in SE Papua, specifically the Massim area and Central Province, and in the Markham Valley, Austronesian languages have remained confined to coastal pockets and offshore islands. Dutton (1977; 1982) documents the diffusion of cultural vocabulary in SE Papua in fairly recent times. There have been other insightful case studies of the effects of bilingualism in Near Oceania (e.g. Bradshaw, 1997; Laycock, 1973b; Lincoln, 1976; Ross, 1996b; Thurston, 1987; 1994). In much of Northern Island Melanesia it seems that the interaction was of a kind that led to widespread language shift. With few exceptions the shifts appear to have been cases of communities that formerly spoke Papuan languages, adopting Austronesian languages while maintaining much of their biological and social distinctiveness. The Admiralties and New Ireland have become entirely Austronesian speaking (except for one language on New Ireland). On the larger islands, such as New Britain, New Ireland, and Bougainville, such language shifts probably occurred on a gradually expanding front. Indeed the shift never occurred among certain communities living in the interior of the large islands (especially in Bougainville) and on certain small islands in the central Solomons. As to the mechanisms of language shift, it seems likely that the spread of agriculture and pottery through Northern Island Melanesia occurred after the arrival of Lapita people and was associated with the spread of Austronesian languages. In the Moluccas and in the Timor region we find a pattern broadly similar to the large islands of Northern Island Melanesia. In most parts of these two regions Austronesian languages predominate, but sizeable areas are occupied by speakers of Papuan languages. The Austronesian and Papuan languages of both the Moluccas and the Timor region each bear the marks of intensive inter-family language contact. Acknowledgments I am grateful to Jack Golson for suggesting a number of improvements to a first draft and to Jonathan Friedlaender for his editorial work
53
the framework
(and patience). Thanks are also due to to Matthew Spriggs, Glenn Summerhayes and Pam Swadling for helpful comments on the archaeological sections of this chapter and to Mark Donohue for doing the same for the linguistics sections. A previous review of “Papuan languages and New Guinea prehistory” was given by William Foley in chapter 8 of Foley (1986).
Notes 1. For example, Tonya Stebbins, of La Trobe University, is currently investigating relationships among the Papuan languages of East New Britain and the Gazelle Peninsula. 2. I will make only passing reference here to the question of whether ancient linguistic connections between Aboriginal Australian and Papuan languages can be traced. The biological evidence suggests that some Australians and some New Guinea Highlanders share distinctive mutations in certain genetic subsystems. Foley (1986: 271–5) assembles some 15 sets of resemblant forms shared by some New Guinea Highland languages with forms attributed to ‘proto Australian’. Foley correctly observes that there are some grave methodological problems inherent in these comparisons. Most of the lexical resemblances are probably accidental. There are however a few tantalizing resemblances in grammatical structure.
References Allen J. 1993. Notions of the Pleistocene in Greater Australia. In: Smith MA, Spriggs M, Fankhauser B, editors. Sahul in review: Pleistocene Archaeology in Australia, New Guinea and Island Melanesia. Canberra, Australia: Department of Prehistory, Australian National University. pp 139–51. Allen J. 2000. From beach to beach: the development of maritime economies in prehistoric Melanesia. In: O’Connor S., Veth P, editors. East of Wallace’s Line: Studies of present and past maritime cultures of the Indo-Pacific region. Rotterdam: A.A. Balkema. pp 139–76. Allen J, Gosden C. 1996. Spheres of interaction and integration: Modelling the culture history of the Bismarck Archipelago. In: Davidson J, Irwin G, Leach F, Pawley A, Brown D, editors. Oceanic culture history: Essays in honour of Roger Green. Dunedin, New Zealand: NZ Archaeological Association. pp 183–97. Anderson A. 2000. Slow boats from China: Issues in the prehistory of Indo-Pacific seafaring. Modern Quaternary Research in Southeast Asia 16: 13–50. Bellwood P. 1996. The origins and spread of agriculture in the Indo-Pacific region: Gradualism and diffusion or revolution and colonization? In: Harris DR, Hillman GC, editors. Foraging and farming: The evolution of plant exploitation. London: Unwin Hyman. pp 465–98.
54
Bellwood P.1997. Prehistory of the Indo-Malaysian Archipelago. Honolulu, Hawaii: University of Hawaii Press. Bellwood P. 1999. The time depth of major language families: An archaeologist’s perspective. In: Renfrew C, McMahon A, Trask L, editors. Time depth in historical linguistics. Cambridge, UK: MacDonald Institute for Archaeological Research. pp 109–40. Bellwood P. 2004. The first farmers. Malden, Mass.: Blackwell. Bellwood, P, Dizon E, in press. Austronesian cultural origins: Out of Taiwan, via the Batanes Islands, and onwards to Western Polynesia. In Sanchez-Mazas A, Blench R, Ross MD, Peiros I, Lin M, editors. Past human migrations in East Asia: Matching archaeology, linguistics and genetics. London: RoutledgeCurzon. Blust R. 1978. Eastern Malayo-Polynesian: A subgrouping argument. In: Wurm SA, Carrington L, editors. Second International Conference on Austronesian Linguistics: Proceedings, Canberra, Australia: Pacific Linguistics C–61. pp 181–234. Blust R. 1995a. The prehistory of the Austronesian-speaking peoples: Aview from language. Journal of World Prehistory 9: 453–510. Blust R. 1995b. Austronesian Comparative Dictionary. Computer files, Department of Linguistics, University of Hawaii. Blust R. 1999. Subgrouping, circularity and extinction: Some issues in Austronesian comparative linguistics. In: Zeitoun E, Jen-kuei Li P, editors. Selected papers from the Eighth International Conference on Austronesian Linguistics. Symposium Series of the Institute of Linguistics (Preparatory Office), Academia Sinica 1. Taipei: Academia Sinica. pp 31–94. Bradshaw J. 1997. The population kaleidoscope: Another factor in the Melanesian diversity vs Polynesian homogeneity debate. Journal of the Polynesian Society 106: 222–49. Bulmer S. 1975. Settlement and economy in prehistoric Papua New Guinea; a review of the archaeological evidence. Journal de la Société des Océanistes 31: 7–75. Bulmer S. 1982. Human ecology and cultural variation in prehistoric New Guinea. Monographiae Biologicae vol. 42. The Hague: Dr W. Junk. pp 169–206. Bulmer S. 1991. Variation and change in stone tools in the Highlands of Papua New Guinea: The witness of Wañelek. In: Pawley A, editor. Man and a half: Essays in Pacific anthropology and ethnobiology in honour of Ralph Bulmer. Auckland, New Zealand: Polynesian Society. pp 470–8. Capell A. 1944. Peoples and languages of Timor. Oceania 14/3, 191–219, 14/4, 313–37, 15/1, 19–48. Capell A. 1948–49. Distribution of languages in the central Highlands, New Guinea. Oceania 19/2, 104–19, 19/3, 234–53, 19/4, 349–77. Chappell J. 2005. Geographic changes of coastal lowlands in the Papuan past. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural,
Recent Research on the Historical Relationships of the Papuan Languages
linguistic and biological histories of Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 525–40. Chappell J, Omura A, McCulloch M, Esat T, Ota Y, Pandolfi, J. 1994. Revised late Quaternary sea levels between 70 and 30 Ka from coral terraces at Huon Peninsula. In: Ota Y, editor. Study on coral reef terraces of the Huon Peninsula, Papua New Guinea. Yokohama, Japan: Dept. of Geography, Yokohama National University. pp 155–65. Clouse DA. 1997. Towards a reconstruction and reclassification of the Lakes plain languages. In: Franklin K, editor. Papers in Papuan linguistics No 2. Canberra, Australia: Pacific Linguistics A–85. pp 133–236. Comrie BA. 1986. Haruai verb structure and language classification in the Upper Yuat. Language and Linguistics in Melanesia 17: 140–60. Comrie BA. 1990. Lexical variation and genetic affiliation: The case of Haruai. In: Edmondson J, Feagin C, Mühlhausler P, editors. Development and diversity: Linguistic variation across time and space. Dallas/ Arlington: Summer Institute of Linguistics and University of Texas: pp 461–6. Denham T. 2002. Archaeological evidence for mid-Holocene agriculture in interior New Guinea: A critical review. Archaeology in Oceania 38(3): 159–76. Denham T. 2005. Agricultural origins and the emergence of rectilinear ditch networks in the highlands of New Guinea. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural, linguistic and biological histories of Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 329–62. Denham T, Haberle SG, Lentfer C, Fullagar T, Field J, Therin M, Porch N, Winsborough B. 2003. Origins of agriculture at Kuk Swamp in the Highlands of New Guinea. Science 201: 189–93. Dunn M, Reesink G, Terrill A. 2002. The East Papuan languages: A preliminary typological appraisal. Oceanic Linguistics 41: 28–62. Dunn M, Terrill A, Reesink G, Foley R, Levinson S. 2005. Structural phylogenetics and the reconstruction of ancient language history. Science 309: 2072–5. Dutton TE. 1977. The distribution of cultural vocabulary in Central Papua. In: Wurm SA, editor. New Guinea area languages. Vol. 3. Language, culture, society and the modern world. Canberra, Australia: Pacific Linguistics. pp 51–97. Dutton TE. 1982. Borrowing in Austronesian and nonAustronesian languages of coastal south-east mainland Papua New Guinea. In: Halim A, Wurm SA, Carrington L, editors. Papers from the Third International Conference on Austronesian Linguistics: Vol. 1. Currents in Oceanic linguistics. Canberra, Australia: Pacific Linguistics, C–74. pp 109–77. Dyen I, James AT, Cole JWL. 1967. Language divergence and estimated word retention rate. Language 43(1): 150–71. Evans B, Mountain M-J. 2005. Pasin bilong tumbuna: Archaeological evidence for early human activity in
the highlands of Papua New Guinea. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural, linguistic and biological histories of Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 363–86. Foley WA. 1986. The Papuan languages of New Guinea. Cambridge: Cambridge University Press. Foley WA. 1992. Language and identity in Papua New Guinea. In: Attenborough RD, Alpers MP, editors. Human biology in New Guinea. The small cosmos. Oxford: Clarendon Press. pp 136–49. Foley WA. 2005. Linguistic prehistory in the Sepik-Ramu Basin. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural, linguistic and biological histories of Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 109–44. Friedlaender JS. 1975. Patterns of human variation: The demography, genetics, and phenetics of Bougainville Islanders. Cambridge, MA: Harvard University Press. Friedlaender JS, Gentz F, Friedlaender F, Kaestle F, Schurr TG, Koki G, Mgone CS, McDonough J, Smith L, Merriwether DA. 2005. Mitochondrial genetic diversity and its determinants in Island Melanesia. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural, linguistic and biological histories of Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 693–716. Golson J. 1977. No room at the top: Agricultural intensification in the New Guinea Highlands. In: Allen J, Golson J, Jones R, editors. Sunda and Sahul: Prehistoric studies in Southeast Asia, Melanesia and Australia. London & New York: Academic Press. pp 601–38. Golson J. 1991. Bulmer phase II: Early agriculture in the New Guinea Highlands. In: Pawley A, editor. Man and a half: Essays in Pacific anthropology and ethnobiology in honour of Ralph Bulmer. Auckland, New Zealand: Polynesian Society. pp 484–91. Golson J, Gardner DS. 1990. Agriculture and social organisation in New Guinea Highlands prehistory. Annual Review of Anthropology 19: 395–417. Golson J, Denham T, Swadling P, Muke J. forthcoming. 10,000 years of gardening: Kuk and the archaeology of agriculture in Papua New Guinea. Gorecki P., Gillieson DS. 1989. A crack in the spine: Prehistory and ecology of the Jimi-Yuat Valley, Papua New Guinea. Townsville: Division of Archaeology and Anthropology, School of Behavioural Sciences, James Cook University of North Queensland. Gosden C. 1992. Production systems and the colonization of the Western Pacific. World Archaeology 24: 55–69. Gosden C, Robertson N. 1991. Models for Matenkupum: Interpreting a late Pleistocene site from Southern New Ireland, Papua New Guinea. In: Allen J, Gosden C, editors. Report of the Lapita Homeland Project. Canberra, Australia: Dept. Prehistory, Australian National University. pp 20–45.
55
the framework
Green RC. 2003. The Lapita horizon and traditions— Signature for one set of oceanic migrations. In Sand, C, editor. Pacific archeology: Assessments and prospects. Proceedings of the international conference for the 50th anniversary of the first Lapita excavation. Kone Noumea 2002. Noumea: Les Cahiers de l’archeologie en Nouvelle. Caledonie. pp 95–120. Greenberg JH. 1971. The Indo-Pacific hypothesis. In: Sebeok TA, editor. Current trends in linguistics 8: Linguistics in Oceania. The Hague: Mouton. pp 807–71. Groube LM. 1986. Waisted axes of Asia, Melanesia and Australia. In: Ward GK, editor. Archaeology at ANZAAS Canberra. Canberra, Australia: Dept of Prehistory and Anthropology, Australian National University. pp 168–77. Groube LM, Chappell J, Muke J, Price D. 1986. A 40,000 year old human occupation site at Huon Peninsula, Papua New Guinea. Nature 324: 453–5. Haiman J. 1979. Review of Wurm (ed.) 1975. Language 55(4): 894–903. Hays T. 2005. Vernacular names for staple foods in Irian Jaya: Implications for agricultural prehistory. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural, linguistic and biological histories of Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 625–71. Heeschen V. 1978. Review of Wurm (ed.) 1975. Anthropos 72: 978–80. Hope G., Golson J. 1995. In Allen J and O’Connell JF, editors. Transitions: Pleistocene to Holocene in Australia and Papua New Guinea. Antiquity 69, Special Number 265. pp 818–30. Hope G, Haberle S. 2005. The history of the human landscapes of New Guinea. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural, linguistic and biological histories of Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 541–54. Hudson E. 1991. All nggaun: Oceanic terms for ‘dog’. In: Harlow R, Hooper R, editors. VICAL I: Papers from the Fifth International Conference on Austr onesian Linguistics, part one. Auckland, New Zealand: Linguistic Society of New Zealand. pp 283–304. Kirch PV. 1997. The Lapita peoples. Ancestors of the Oceanic world. Oxford, UK: Blackwell. Kirch PV. 2000. On the road of the winds: An archaeological history of the Pacific Islands before European contact. Berkeley, CA: University of California Press. Koler-Matnick J, Brisbin IL Jr, Feinstein M, Bulmer S, 2003. An updated description of the New Guinea singing dog (Canis hallstromi, Troughton 1957). Journal of the Zoological Society of London 261: 109–18. Kruskal J, Dyen I, Black P. 1971. The vocabulary method of reconstructing family trees: Innovations and large scale applications. In: Hodson RF, Kendall DG, Tautu P, editors. Mathematics in the archaeological and historical sciences. Edinburgh: Edinburgh University Press. pp 361–80.
56
Lang R. 1976. Review of Wurm (ed.) 1975. Kivung. Journal of the Linguistic Society of Papua New Guinea 9(1): 72–80. Laycock DC. 1965. The Ndu language family (Sepik District, New Guinea). Canberra, Australia: Pacific Linguistics. Laycock DC. 1973a. Sepik languages—Checklist and preliminary classification. Pacific Linguistics B–25. Canberra, Australia: Pacific Linguistics. Laycock DC. 1973b. Sisssano, Warapu and Melanesian pidginization. Oceanic Linguistics. 12: 245–77. Laycock DC. 1975. The Torricelli phylum. In: Wurm SA, editor. New Guinea area languages and language study 1: Papuan languages and the New Guinea linguistic scene. Pacific Linguistics C–38. Canberra, Australia: Australian National University. pp 767–80. Laycock DC, Z’graggen J. 1975. The Sepik-Ramu phylum. In: Wurm SA, editor. New Guinea area languages and language study 1: Papuan languages and the New Guinea linguistic scene. Pacific Linguistics C–38. Canberra, Australia: Australian National University. pp 731–63. Lees RB. 1953. The basis of glottochronology. Language 29: 113–27. Lincoln PC. 1976. Banoni, Piva and Papuanization. Canberra, Australia: Pacific Linguistics A45: 77–105. Lincoln PC. 1978. Reef–Santa Cruz as Austronesian. In: Wurm SA, Carrington L, editors. Second International Conference on Austronesian Linguistics: Proceedings. Canberra, Australia: Pacific Linguistics, C–61. pp 929–67. McElhanon KA. 1975. North-Eastern Trans-New Guinea phylum languages. In: Wurm SA, editor. New Guinea area languages. Vol. 1. Papuan languages and the New Guinea linguistic scene. Canberra, Australia: Pacific Linguistics, C–38. pp 527–67. Nichols J. 1992. Linguistic diversity in space and time. Chicago: Chicago University Press. Pavlides C, Gosden C. 1994. 35000 year-old sites in the rainforests of West New Britain, Papua New Guinea. Antiquity 68: 604–10. Pawley A. 1995. C.L. Voorhoeve and the Trans New Guinea phylum hypothesis. In: Baak C, Bakker M, van der Meij D, editors. Tales from a concave world. Liber Amoricum Bert Voorhoeve. Leiden University, Projects Division, Department of Languages and Cultures of South-East Asia and Oceania. pp 83–122. Pawley A. 1998. The Trans New Guinea phylum hypothesis: A reassessment. In: Miedema J, Ode C, Dam RAC, editors. Perspectives on the Bird’s Head of Irian Jaya, Indonesia. Amsterdam, Holland: Editions Rodopi. pp 655–89. Pawley A. 2000. Some Trans New Guinea phylum cognate sets. Printout. Dept. Linguistics, RSPAS, ANU. Pawley A. 2001. The proto Trans New Guinea obstruents: Arguments from top-down reconstruction. In: Pawley A, Ross M, Tryon D, editors. The boy from Bundaberg: Studies in Melanesian linguistics in honour of Tom Dutton. Canberra, Australia: Pacific Linguistics. pp 261–300.
Recent Research on the Historical Relationships of the Papuan Languages
Pawley A. 2002. The Austronesian dispersal: People, language, technology. In: Bellwood P, Renfrew C, editors. Examining the language/farming dispersal hypothesis. Cambridge UK: McDonald Institute for Archaeological Research. pp 251–73. Pawley A. 2005a. The chequered career of the Trans New Guinea hypothesis: Recent research and its implications. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural, linguistic and biological histories of Papuan speaking-peoples. Canberra, Australia: Pacific Linguistics. pp 67–107. Pawley A. 2005b. Madang languages. In Brown K, editor, Encyclopaedia of Language and Linguistics, 2nd edition, vol. 7. Oxford: Elsevier. pp 429–432. Pawley A. n.d. Notes on the history of verb morphology in the Trans New Guinea phylum languages. Printout [2000] Dept. Linguistics, RSPAS, Australian National University. Pawley A, Ross M, Osmond M. forthcoming. Papuan languages and the Trans New Guinea family. Canberra, Australia: Pacific Linguistics. Reesink G. 2005. West Papuan languages: Roots and development. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural, linguistic and biological histories of Papuan-speaking peoples. Canberra, Australia: Pacific Linguistics. pp 185–219. Renfrew C. 1992. World languages and human dispersal: A minimalist view. In: Hall H, Jarvie JC, editors. Transition to modernity. Cambridge: Cambridge University Press. pp 11–68. Renfrew C. 2002. ‘The emerging synthesis’: The archaeogenetics of farming/language dispersals and other spread zones. In: Bellwood P, Renfrew C, editors. Examining the language/farming dispersal hypothesis. Cambridge UK: McDonald Institute for Archaeological Research. pp 3–16. Roberts RG, Jones R, Smith M. 1990. Thermoluminescence dating of a 50,000 year old human occupation site in northern Australia. Nature 345: 153–6. Roscoe P. 2005. Foraging, ethnographic analogy, and Papuan pasts: Contemporary models for the SepikRamu past. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural, linguistic and biological histories of Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 555–84. Ross MD. 1988. Proto Oceanic and the Austronesian languages of Western Melanesia. Canberra, Australia: Pacific Linguistics C–88. Ross MD. 1995. The great Papuan pronoun hunt: Recalibrating our sights. In: Baak C, Bakker M, van der Meij D, editors. Tales from a concave world. Liber Amoricum Bert Voorhoeve. Leiden University, Projects Division, Department of Languages and Cultures of South-East Asia and Oceania. pp 139–68. Ross MD. 1996a. Pottery terms in proto Oceanic. In: Davidson J, Irwin G, Leach F, Pawley A, Brown D, editors. Oceanic culture history: Essays in honour of Roger Green. Dunedin, New Zealand: NZ Archaeological Association. pp 67–82.
Ross MD. 1996b. Contact-induced change and the comparative method: Cases from Papua New Guinea. In: Durie M, Ross MD, editors. The comparative method reviewed: Irregularity and regularity in linguistic change. New York: Oxford University Press. Ross MD. 2000. Pronouns as preliminary evidence for grouping Papuan languages. Printout, RPAS. In: Pawley A, Ross M and Osmond M. (forthcoming). Ross MD. 2001. Is there an East Papuan phylum? Evidence from pronouns. In: Pawley A, Ross M, Tryon D, editors. The boy from Bundaberg: Studies in Melanesian linguistics in honour of Tom Dutton. Canberra, Australia: Pacific Linguistics. pp 301–32. Ross MD. 2005. Pronouns as a preliminary diagnostic for grouping Papuan languages. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural, linguistic and biological histories of Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 15–66. Specht J. 2005. Revisiting the Bismarcks: Some alternative views. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts. Cultural, linguistic and biological histories of Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 235–88. Spriggs M. 1997. The Island Melanesians. Oxford, UK: Blackwell. Summerhayes G. 2000a. Lapita interaction. No. 15, Terra Australia. Canberra, Australia: Archaeology and Natural History Publications and the Centre for Arch. Research, ANU. Summerhayes G. 2000b. Recent archaeological investigations in the Bismarck archipelago, Anir-New Ireland Province, Papua New Guinea. Indo-Pacific Prehistory Bulletin 19: 167–74. Suter E. 1997. A comparative look at the dual and plural forms of inflections and pronouns in Northeast New Guinea Papuan languages. Language and Linguistics in Melanesia 28: 17–68. Swadesh M. 1952. Lexico-statistical dating of prehistoric ethnic contacts, with special reference to North American Indians and Eskimos. Proceedings of the American Philosophical Society 96: 452–63. Swadling P, Hide R. 2005. Changing landscape and social interaction: Looking at agricultural history from a Sepik-Ramu perspective. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan Pasts. Cultural, linguistic and biological histories of Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 289–328. Swadling P, Chappell J, Francis G, Araho N, Ivuyo B. 1989. A late quaternary inland sea and early pottery in Papua New Guinea. Archaeology in Oceania, 24: 106–9. Terrill A. 2002. Systems of nominal classification in East Papuan languages. Oceanic Linguistics 41: 63–88. Thorne A, Grün R, Mortimer G, Spooner N, Simpson J, McCulloch M, Taylor L, Curnoe D. 1999. Australia’s oldest human remains: Age of the Lake Mungo 3 skeleton. Journal of Human Evolution 36: 591–612.
57
the framework
Thurston WR. 1987. Processes of Change in the Languages of North-western New Britain. Canberra, Australia: Pacific Linguistics. Thurston WR. 1994. Renovation and innovation in the languages of north-western New Britain. In: Dutton TE, Tryon DE, editors. Language contact and change in the Austronesian world. Berlin, Germany: Mouton de Gruyter. pp 530–610. Todd EM. 1975. The Solomon language family. In: Wurm SA, editor. New Guinea area languages and language study 1: Papuan languages and the New Guinea linguistic scene. Pacific Linguistics C–38. Canberra, Australia: Australian National University. pp 805–46. Voorhoeve CL. 1968. The Central and South New Guinea phylum: A report on the language situation in south New Guinea. In Papers in New Guinea Linguistics No. 8. Canberra, Australia: Pacific Linguistics. pp 1–17. Voorhoeve CL. 1975. Central and Western trans-New Guinea phylum languages. In: Wurm SA, editor. New Guinea area languages and language study 1: Papuan languages and the New Guinea linguistic scene. Pacific Linguistics C–38. Canberra, Australia: Australian National University. pp 344–459. Wurm SA. 1971. The Papuan linguistic situation. In: Sebeok TE, editor. Current trends in linguistics Vol. 8. Oceania. The Hague: Mouton. pp 541–657. Wurm SA (editor). 1975. New Guinea area languages and language study 1: Papuan languages and the New Guinea linguistic scene. Pacific Linguistics C–38. Canberra, Australia: Australian National University. Wurm SA. 1975. Personal pronouns. In: Wurm SA, editor. New Guinea area languages. Vol. 1. Papuan languages
58
and the New Guinea linguistic scene. Canberra, Australia: Pacific Linguistics, C–38. pp 191–217. Wurm SA. 1978. Reef–Santa Cruz: Austronesian, but … In: Wurm SA, Carrington L, editors. Second International Conference on Austronesian Linguistics: Proceedings, Canberra, Australia: Pacific Linguistics C–61. pp 969–1010. Wurm SA. 1982. Papuan languages of Oceania. Ars Linguistica 7. Tübingen: Gunter Narr. Wurm SA, Hattori S (editors). 1981–83. Language atlas of the Pacific area. (Vol. 1 1981, Vol. 2 1983.) Canberra, Australian Academy for the humanities in collaboration with the Japanese Academy. Canberra, Australia: Pacific Linguistics C–66. Wurm SA, McElhanon KA. 1975. Papuan language classification problems. In: Wurm SA, editor. New Guinea area languages and language study 1: Papuan languages and the New Guinea linguistic scene. Pacific Linguistics C–38. Canberra, Australia: Australian National University. pp 145–64. Wurm SA, Voorhoeve CL, McElhanon KA. 1975. The transNew Guinea phylum in general. In: Wurm SA, editor. New Guinea area languages and language study 1: Papuan languages and the New Guinea linguistic scene. Pacific Linguistics C–38. Canberra, Australia: Australian National University. pp 299–322. Z’graggen JA. 1975. The Madang-Adelbert range subphylum. In: Wurm SA, editor. New Guinea area languages and language study 1: Papuan languages and the New Guinea linguistic scene. Pacific Linguistics C–38. Canberra, Australia: Australian National University. pp 569–612.
part ii Core Studies in Northern Island Melanesia
This page intentionally left blank
4 Mitochondrial DNA Variation in Northern Island Melanesia Jonathan S. Friedlaender, Françoise R. Friedlaender, Jason A. Hodgson, Stacy McGrath, Matthew Stoltz, George Koki, Theodore G. Schurr, and D. Andrew Merriwether
The chapters in part II of this book analyze the truly remarkable variation in Northern Island Melanesia (as seen in both genetic and linguistic evidence). This first chapter, which focuses on mitochondrial DNA (mtDNA) variation, is particularly informative from the genetic side because it is based on an especially large and intensive sample set, and because the mtDNA itself is extremely variable and informative. We have found a series of mitochondrial variants that developed in this region that likely date from the initial settlement >40,000 years ago, and extend across the intervening periods up to ~3,500 years ago. Many of these variants developed in particular locales still closely associated with specific Papuan-speaking groups, and their genetic distributions and stratigraphies suggest a series of local population expansions over that long time span. This pattern sharply contradicts the scenario featuring a long period of inactivity following first settlement, terminated by a major intrusion from Southeast Asia/Taiwan at ~3,500 years ago. The implications of these findings for the modeling of the settlement and subsequent population history of Northern Island Melanesia are explored here.
Introduction A great deal of attention in population genetics has centered on variation in the mtDNA molecule. The very small
ring of 16,569 mtDNA base pairs is transmitted as a unit from mother to offspring, rather than experiencing the shuffling of recombination that affects genes carried on the paired autosomes in the nucleus. As a consequence, separate mtDNA lineages arise through the slow accumulation of mutations, with the more recent mutations occurring only on a subset of older variant backgrounds or haplotypes (they form nested sets). This mutational process allows for a straight-forward reconstruction of the step-wise sequence of ancient mutations, and the construction of a phylogenetic tree, which is otherwise almost impossible to accomplish for the recombining sections of chromosomes in the nuclear genome. Figure 4.1 shows the mitochondrial DNA genome and indicates some key ancient mutations that have defined major east Eurasian lineages, including those that first appeared in South Asia and the Southwest Pacific. Although the mtDNA is a relatively tiny section of the DNA in a cell, it is still laborious and costly to sequence all 16,569 bases or nucleotide sites (nts) in hundreds of samples for the typical population survey. A great deal of the mtDNA genome codes for important proteins and other molecules involved in cellular metabolism, and is largely invariant because of the selection pressures to maintain those processes. Therefore, those sections are of little use in studying distinctions among human populations. However, particular sites scattered around the coding portion of the mtDNA ring are more likely to
61
core studies in northern island melanesia
Figure 4.1 Schematic diagram of the mtDNA. Important nucleotide site markers for Southwest Pacific populations are noted. Segments denote different units of the molecule. Refer to the map at http://www.mitomap.org for details.
accumulate mutations because of certain position effects, making them important. Approximately 1000 base pairs at the apex called the control region contain hypervariable regions I and II (HVR I and II), roughly between nts 16,000 and 400. HVR I and II contain a number of sites that accumulate mutations at an especially rapid rate, and the usual analytic approach today is to sequence this region first. Sequencing the control region ordinarily provides enough identifying information for phylogenetic lineage (or haplogroup) assignments in most instances. Novel haplotypes are then selected for more extensive sequencing in the coding part of the mtDNA ring to tie them to appropriate branches of the mtDNA tree. For especially unusual haplotypes, one usually sequences the entire mtDNA ring. Figure 4.2 shows an abbreviated outline of the entire Eurasian mtDNA tree, which stems from the African L3 lineage, with those major clades or macrohaplogroups that occur in the Southwest Pacific shown in boldface. Their numerical representations in our sample series are shown above. While establishing the sequence of mutational events in different mtDNAs is a relatively straight-forward process, the absolute dating of those events is far more problematic and can never hope to attain the exactitude of carbon-14 dating in archeology. Their calibration depends either on (differing) chimpanzee/human divergence estimates (Ingman and Gyllensten, 2003; Mishmar et al., 2003), or on ancient human
62
population bifurcations, assumed to be undiluted by subsequent mixing (Forster et al., 1996; Saillard et al., 2000). However, for the past few years, there has been a developing critical literature that suggests major problems with the molecular clock. There are clear problems with rate heterogeneity among haplogroup lineages (Howell et al., 2004), as well as problems with establishing proper calibration points or splits, especially for population data. Even more troubling, it has been argued that the molecular clock is particularly unsuitable when applied to young, and particularly expanding, populations. Ho et al. (2005) think mutation rates estimated from old divergences are much slower than those from more recent divergences. They believe there is a substitution rate decay that should be corrected for when calculating absolute dates, especially for recent ages (less than 1 million years). They suggest the causes could be saturation at hotspots and purifying selection on slightly detrimental mutations, along with drift. Others have recently suggested that mutations occurring during a range expansion can get driven to high frequency or even fixation by “surfing” on the wave front of expansion by benefiting from the repeated bottlenecking that characterizes expansions (Edmonds et al., 2004; Klopfstein et al., 2006). The result of this is that under expansion conditions the substitution rate can be elevated. In this chapter, we will follow the “standard” molecular dating practices, but the derived absolute dates are to be used in relative terms only. More on these issues will follow in the Discussion.
Mitochondrial DNA Variation in Northern Island Melanesia
Figure 4.2 mtDNA schematic tree for Eurasian haplotypes (note L3 origin). Haplotypes found in our series are in bold, with number of occurrences shown above.
Also, the mtDNA only represents the female side of the ancestral equation, since it is only passed down from mothers to their children. This is the reason the combination of mtDNA and Y patterns should provide a more complete picture. In fact, mtDNA variation in today’s populations represents a small fraction of ancient haplotypes that once existed, since many must have been lost to the normal process of extinction through genetic drift. Nevertheless, linking the phylogeny of mtDNA mutation history with the geographical distributions of the different lineages (referred to as phylogeography— Avise et al., 1987) has provided compelling evidence for the recent origin of modern humans in Africa, and their spread throughout Eurasia sometime during the last 150,000–50,000 years.
Tables 4.1, 4.2 and figure 4.3 show the frequencies of the major mtDNA haplogroups in Island Southeast Asia and the Pacific. This distribution reveals the dramatic shift in their occurrence across the region of the Wallace Line (or Wallacea). There are many mtDNA haplogroups present in Asia but missing in the Pacific (e.g., many branches of M, as well as R, U, W, B4c, and B5), and also a constellation of haplogroups that are specific to the Southwest Pacific to the exclusion of Asia (P1, P2, P4, Q1, Q2, Q3, M27, M28, and M29). This pattern provides clear evidence of the relative isolation of the populations that entered this region. Certain mtDNA lineages do occur in appreciable frequency in both Southeast Asia and Near and Remote Oceania, and these are the candidates for being introduced from Southeast Asia more recently during the Holocene. Among these are B4a1a, E, and also the less common haplogroups B4b, F, M7, and Y.
An Overview of Regional mtDNA Haplotype Distributions The outline of the Eurasian mtDNA phylogenetic tree (figure 4.2) indicates that two primary lineages, N and M, branched from the African L3 root at a time estimated approximately 60,000 years ago. Representatives of N and M appeared soon afterwards in the Southwest Pacific, specifically the ancient continent of Sahul and Northern Island Melanesia, with their first settlement some time before 40,000 years ago.
MtDNA Phylogeny in Near Oceania and Island Melanesia The mtDNA variants found in Near Oceania and Island Melanesia fall into groups according to their ages and geographical distributions. They will be discussed in order from the most ancient to the most recent. The oldest appear to have developed in ancient Sahul around the time of its initial settlement, but a second old set appears
63
64 Table 4.1
mtDNA Lineage Frequencies of Island Melanesia and Nearby Regions Haplogroup Frequencies (%) B
Population
N
B*
B4a
Vietnam Thailand Moken Thai Thailand Urak Lawoi Malaysia Malaysia Melayu Orang Asli Sabah Taiwan Amis Bunun Atayal Paiwan Puyuma Rukai Tsou Saisiat Yami Indonesia Borneo Java Sumatra Indonesia Philippines
59
2
7
3
3 2
1
8 36 527 10 52 7 260 37
5
B4c
B5
8
8
2
11 5
E*
E1a
E1b
E2
M8
Z
C
6
2 14
1 8 42 5 24 5 11 11 5 35
3 33 6
3 1
6 1
12
5 3 1
5 18 1 9
5 5 8 10 14
1 2 8
7 9 4
8 14 7 11
10 1 8 7 15 12 1 11 16 8 8 9
D
G
2
2
3 4
11 14 20
4
6
20
125 116 136 80 74 70 80 83 84 19 19 99 259 119
B4b
E
10
6
19 11 2 4
4
1 17 1 4 5 1 6 27
6 6 4
8 5 16 4 19 1
2 4
2
11
3
5 5 6 12
5 1 3 1
1 4
5 5
F
M*
M7
15
8
14
22 22
88 17 22 20
11 3
12
12
8
11 5
26 3
3 19
8 27 30 30 43 43 23 13 20 16 5 18 16 8
3 3
5 16 7 13 2
15 4 37 10 23 17 3 28 29 11 26 12 7 18
N
P
Q
R
U
Y
Other
Source1
7
5
19
a,b
1
3 5
13 17 5 40
a a,c a,b,c,d a
31 43 16 19
e f f g
2 14 7
4
2 14 30 14
3
6 1 1
2
1
1 1 1
21 10
1
1
3
5 3 2 2
1
10 2 3
37 16 10 7 14
h,e,k h,e,k h,e,k h,e,k h,e h,e h,e h,e h,e a a f a,e,i,I a,g,e
Data sources a: Lum et al. (2000), b: Oota (2001), c: Yao et al. (2003), d: Fucharoen et al. (2001), e: Tajima et al.(2004), f: Macaulay et al. (2005), g: Sykes (1995), h:Trejaut et al. (2005), i: Redd (1999), k: Melton (1998), l: Cox (2003).
Mitochondrial DNA Variation in Northern Island Melanesia
Figure 4.3 Regional distribution of major haplogroup frequencies in Asia and Oceania (see tables 4.1 and 4.2). See color insert.
to have developed in Northern Island Melanesia not appreciably later. Other branches developed subsequently in the Upper Pleistocene. The most recent set spread from Island Southeast Asia during the Holocene with some haplogroups dating to ~8,000 years before present (YBP), and the most recent dating approximately to the time of the development of the Lapita Cultural Complex.
The Oldest Haplogroups Common to New Guinea Haplogroup P P is clearly the oldest branch of R in the Southwest Pacific. The different branches of P with their defining mutations at different nts are shown in figure 4.4 and table 4.3. Table 4.4a gives the associated age estimates from the coding region for P and its branches. In figure 4.4, nts in bold indicate control region variants, which generally are not used in age estimations because of their high rates of mutation and recurrence. The extensive branches
of P are united by only one mutation at nts 15,607. Some of the links between branches in figure 4.4 are not entirely certain, as there appear to have been some back mutations at particular nts, especially at the base of P4. West of the Wallace Line, P has only been found in three samples from Indonesia (table 4.1). P is particularly interesting because, while a number of branches are apparently specific to either Near Oceania (P1 and P2) or to aboriginal Australia (see Friedlaender et al., 2005b), certain branches (P3 and P4) do show some distant and very old connections between Near Oceania and Australia. As shown in table 4.2 and figure 4.5, P is most common in New Guinea, especially the highlands, is less common in New Britain, and is increasingly rare in New Ireland, Bougainville, and locations further southeast. Table 4.2 shows that P1 is its most common and widespread representative in Near Oceania. P2 and P4 are also most common in New Guinea as well, although P4 is also found in Vanuatu (Cox, 2003). The age estimates for P and its common branches in Near Oceania are in table 4.4a. The founder ages, i.e., when the branches first split from R,
65
core studies in northern island melanesia
Table 4.2
mtDNA Lineage Occurrences in Island Melanesia (Individuals) B
Island
Region
Population
New Guinea
West New Guinea
Southwest Riverine Lowland Riverine Markham North Coast Rigo Eastern Highlands Fringe Highlands Morobe Highlands Western Highlands Misima Rossel Manus Kove Anêm Mangseng Mamusi Nakanai Nakanai (Loso) Mengen Melamela Ata Kol Sulka Tolai Mali (Baining) Kaket (Baining) Mussau Lavongai Tigak Nalik Notsi Madak Kuot Saposa Teop Buka Aita Rotokas Eivo Simeku Nasioi Nagovisi Siwai Torau Solomons Baegu Lau Kwaio
PNG Coast
PNG Highlands
PNG Island
PNG Island
Manus New Britain
Manus West New Britain
East New Britain
Mussau New Hanover New Ireland
Bougainville
Mussau Lavongai North New Ireland AN
New Ireland PAP North Bougainville
Central Bougainville
South Bougainville
Solomon Islands
Solomons Malaita
Santa Cruz Vanuatu New Caledonia Fiji Micronesia Polynesia
66
N
B4a1
33 27 67 14 18 4 18 20 18 7 5 2 20 19 17 63 64 17 23 23 58 57 28 78 58 59 16 18 27 24 23 31 62 25 20 15 54 19 19 7 33 16 19 5 27 103 23 37 69 23 25 15 47 9 1628
1
P B4b1
P1 7 3 9 6 1 2 4 5 10 4 3
22 4 10
7
11 11 4 1 23
2 1
P2
Q P3
P4
Q1
1
1 7
19 22 12 4
6 5
14 3 13
8 12 23 15 19 15 59 22 10 9 2
9 7 1 2
1
1
13 3 16 16 14 4 18 78 16 20 23 8 6 9 32 9 612
3 2 7
1 4 2 3 1 2
Q3
10
6
1 7 2
Q2
1 3 1 7
2 1 6 4 20 9 8 6 1 2 7 5 4 1
1 2
5 2 6 1 2
18 6
21 2 8
1 1 1
1
1 7 1 5
2 2 4 42 8 2
8
5
1
1
5 4 2
1 8 1
3 3 6 12
4
2
7 2 1 1
237
106
4 8
99
38
1
22
7
Mitochondrial DNA Variation in Northern Island Melanesia
E E1a
E1b
M27 E2
a
b
M28 c
a1
a2
a3
M29 a4
b1
b2
a
b
F
Y
M7
2 1
1 1
1 1
3
1
1
2 19
17 3 1
2 5
21
6
3 5
23
14 7 3
4 3
1 13
1
14 6 1 10 24
2
6
1 5 6 13 1
7 23 28
22
1
1
4 2 2 1
9
5 1
1 1
1
1
3 1 2 11 2 2 8
1 1 4
1 1
5 1 2 1 5
2 18 1
1 1 5
1
1
1
1
1
18 1 2 1
1 1 3
2 3
1
1
1 1
2
1 11
8
75
2
41
37
15
47
104
28
10
59
14
24
16
4
1
13 67
core studies in northern island melanesia
Figure 4.4 Phylogenetic tree of macrohaplogroup P branches found in Near Oceania, with shared Australian brances shown (for additional branches, see Friedlaender et al., 2005b). GenBank accession numbers are listed. Control region mutations are in bold; back mutations are in italics; suffix letters are transversions; underline indicates recurrence in this tree.
are particularly ancient, and could be taken to indicate the initial branching of P occurred prior to the initial settlement of Sahul.
Macrohaplogroup M Many deep branches of M have been found throughout Asia, especially India (Kivisild et al., 1999, 2002; Palanichamy et al., 2004; Macaulay et al., 2005; Metspalu et al., 2005; Thangaraj et al., 2005; Sun et al., 2006). As mentioned, the generally accepted interpretation of this very deep tree is that it was caused by a single ancient expansion out of Africa in the vicinity of 60,000 years ago. Figure 4.6 shows the main branches of macrohaplogroup M that occur in Near Oceania as we now understand them. None of them can be tied to any particular branches of M in Asia. In fact, it is truly remarkable how many deep and old 68
branches of M are restricted to this particular region. At least at this stage in our analyses, there are even fewer suspected M links between Aboriginal Australia and Near Oceania than for macrohaplogroup P, although this pattern may change as more whole-genome sequencing of Australian Aboriginal M variants is conducted. A good deal of research has shown that the New Guinea and Northern Island Melanesian branches of M developed around the time of initial settlement, beginning approximately 40,000 years ago, although some may well have diverged at even earlier dates (Forster et al., 2001; Huoponen et al., 2001; Ingman and Gyllensten, 2003; Friedlaender et al., 2005a, 2005b; Merriwether et al., 2005). Our age estimates for these ancient M haplogroups and their branches are given in table 4.4b, again using the standard techniques. Overall, the founder ages appear to be equivalent to those for P. However, as mentioned,
Mitochondrial DNA Variation in Northern Island Melanesia
Table 4.3
Defining Mutations for mtDNA Haplogroups Found in Island Melanesia
Haplogroup
HVS 1
HVS 2
Coding Region
B4a1a1 B4a1a1a
16189, 16217, 16261 16189, 16217, 16261, 16247
146 146
9bp del, 10398 9bp del, 10398
B4b1
16189, 16217, 16136
207
9bp del
P1 P2a P2b
16357, 16176, 16266 16278, 16497 16184, 16256
212 143
10398, 15607, 6077 10398, 15607, 8572 10398, 15607, 8572
P4
16319
35, 36, 146, 152
Q1 Q2 Q3
16223, 16129, 16241, 16311, 16265C, 16144, 16148, 16343 16223, 16129, 16241, 16066 16223, 16129, 16241, 16311
89, 146, 92 228T, 195 143
E1
M28
M29
10400, 4177, 12940 10400, 4177, 12940 10400, 4177, 12940
a b
16223, 16362, 16390, 16291 16223, 16362, 16390, 16261 16223, 16362, 16390, 16051
a b c
16223, 16048, 16077T, 16172, 16311, 16320 16209, 16299, 16390 16223, 16301, 16304
195, 234, 228 146, 186
10400, 5375, 9201, 12538 10400, 5375, 9201, 12538 10400, 5375, 9201, 12538
a1 a2 a3 a4
16223, 16148, 16468, 16362, 16086, 16129, 16320 16223, 16148, 16468, 16362, 16086, 16129, 16429 16223, 16148, 16468, 16362, 16086, 16129, 16189, 16209 16223, 16148, 16468, 16362, 16086, 16129, 16051
152, 195 152, 195 152 195 279 152 195 198
10400 10400 10400 10400
b1 b2
16223, 16148, 16468, 16362, 16318C 16223, 16148, 16468, 16362, 16318T
152, 94 152, 94
10400 10400
a b
16223, 16189, 16311 16223, 16189, 16294
211, 310 211, 200
10400 10400
E2 M27
15607
10400, 4491, 7598 10400, 4491, 7598 10400, 4491, 7598
Bold-faced nucleotide sites (nts) denote key mutations.
Table 4.4a
Coding Region Age Estimates for P Haplogroups P1-P3 in Near Oceania
Haplogroup
N
ρ
σ
TMRCA, Years
SD
Founder age, Years
SD
P1 P2 P3
6 5 5
8.333 4.800 11.600
1.354 1.131 1.811
42,800 24,700 59,600
7,000 5,800 9,300
53,000 60,600 85,300
10,000 14,800 14,800
Table 4.4b
Coding Region Age Estimates for M Haplogroups in Near Oceania N
ρ
σ
TMRCA, Years
SD
7 2 3 2
13.430 1.500 0.667 0.500
2.176 0.866 0.667 0.500
69,000 7,700 3,400 2,600
11,200 4,500 3,400 2,600
84,400
14,300
M27a M27b M27c
4 2 2
7.000 3.000 8.000
1.458 1.225 2.000
36,000 15,400 41,000
7,500 6,300 10,300
61,700
13,700
M28a M28b
3
0.667
0.471
3,400
2,400
65,100
18,000
12 6 3 3
8.667 3.833 4.667 7.333
1.434 0.833 1.247 1.886
44,500 19,700 24,000 37,000
7,400 4,300 6,400 9,700
70,200
13,700
Haplogroup M27
M28
M29 Q Q1 Q2 Q3
Founder Age, Years
SD
69
core studies in northern island melanesia
Figure 4.5 Summary view of mtDNA haplogroup frequencies as reported in table 4.2. See color insert.
there are serious questions about absolute dates using the molecular clock. Haplogroup Q Q is the most common branch of M in New Guinea. As shown in figure 4.6 and table 4.3, Q has a number of defining mutations, as well as very long internal branches. No Q branches have been identified in Australia thus far, and it is much more common than P in most regions of Near Oceania (table 4.2). The Q1 branch is especially common in West New Guinea, in the Markham Valley, throughout New Britain, and north Bougainville, especially among the Aita (refer to table 4.2 and figure 4.7, which show the population break-down of the haplogroups). Conversely, haplogroup Q2 is rare in New Guinea except for the Markham Valley, but is most common among certain inland Papuan groups of New Britain (Mali Baining and Ata). It also occurs in New Ireland, but is absent in Bougainville. Both Q1 and Q2 are found in Malaita, Santa Cruz, Vanuatu, and less commonly in Fiji (Cox, 2003). We identified only seven Q3 samples: two from the highlands (also reported in Redd and Stoneking, 1999; Ingman and Gyllensten, 2003) and five from the Kove in West New Britain.
Oldest mtDNA Haplogroups in Northern Island Melanesia In addition to Q2, we have identified a number of deep branches of macrohaplogroup M that, from their current
70
distributions, seem to have first appeared in different sections of Northern Island Melanesia and not in ancient Sahul.
Haplogroup M27 As shown in table 4.3 and figure 4.6, this ancient haplogroup has three very long branches that are linked together by three mutations in the coding region. The branches (M27a, b, and c) show rather different distributions but have not been found in New Guinea (table 4.2 and figures 4.5 and 4.7). M27a, which is the most common branch, is centered on Bougainville, and especially the Rotokas Papuan speakers of North Bougainville, but also was detected in Malaita. M27c has a more scattered distribution, identified in New Ireland, Malaita, as well as Bougainville. M27b was only identified in one Bougainville sample, but was relatively common in the Tolai of New Britain, as well as Malaita once again. Beyond this region of Near Oceania, we have identified one M27b individual in New Caledonia, and there may be a couple of possible occurrences of M27a and M27c in Vanuatu (Cox, 2003). To judge from its divergence age estimate (table 4.4b), M27 is particularly ancient (perhaps ~84,400 YBP).
Haplogroup M28 This second old Island Melanesian haplogroup has six different branches (defined in table 4.3). Only four have been entirely sequenced, and these are shown in figure 4.6
Mitochondrial DNA Variation in Northern Island Melanesia
Figure 4.6 Phylogenetic tree of older macrohaplogroup M branches found in Northern Oceania (E is not shown). See figure 4.4 legend for details of notation.
(the two remaining are branches M28a3 and M28a4). M28 appears to have diverged from M during the same early time frame as M27 and Q (table 4.4b). It has not been found west of New Britain, except for a single Misima Island sample, and more than 80% of the M28 samples we identified were from New Britain (figure 4.7). Clearly, this is its origin. Even so, the branches have very heterogeneous distributions within New Britain populations. The center of M28 diversity (and most common occurrence) is in the Baining vicinity, but it spills over into the Tolai, Sulka, Kol, Ata, Nakanai, and Mamusi as well. However, M28a1 is most common among the Mamusi and Ata, and we only found the M28a4 branches in the Melamela in New Britain,
plus scattered individuals from Bougainville, Malaita, and New Caledonia. As shown in table 4.2, these various branches of M28 are rare in New Ireland, Bougainville, and the central Solomons, but occur rather commonly in Santa Cruz, Vanuatu (33%), New Caledonia, and as far into the Central Pacific as Fiji. These haplotypes clearly spread there from New Britain, although exactly when during the last 3,200 years cannot be ascertained.
Haplogroup M29 This haplogroup also apparently has its origin in Northern Island Melanesia. Although its divergence
71
core studies in northern island melanesia
Figure 4.7 Detailed mtDNA haplogroup frequencies by population in Northern Island Melanesia (see table 4.2). See color insert.
from M is certainly ancient, our survey suggests there are only a few (recent) internal branches, at least for M29a. It is most common among the Tolai of East New Britain, who are known to have migrated there from southern New Ireland (an area we did not cover in our survey). We therefore believe that it might have developed in southern New Ireland. We detected it in two other New Britain populations, a north New Ireland population, and in single cases from Vanuatu and Solomon Islands. However, a complete sequencing of M29b will be necessary to determine when it separated from M29a.
Recently Introduced mtDNA Haplogroups from Island Southeast Asia A third set of mtDNA haplogroups is clearly younger than the previously described M, P, and Q branches, dating to the late Pleistocene or Holocene. Their distributions also differ from those just discussed, as they occur not only in Island Melanesia (and other parts of Oceania), but also in Island Southeast Asia. However, our interpretation of
72
these distributions differs from the currently most widely accepted scenario.
Haplogroup B4a1a1 (also known as the “Polynesian Motif”—see figure 4.8) For over a decade, this haplogroup has been widely accepted to have been directly linked with the Austronesian expansion out of Taiwan that led to the development of the Lapita Cultural Complex, and ultimately the settlement of Polynesia and Micronesia (Redd et al., 1995; Lum et al., 1998; Melton, 1998, among others), although there have been some contrary voices suggesting an origin in, and dispersal from, Wallacea (Richards et al., 1998; Oppenheimer and Richards, 2001). Recently, more whole mtDNA sequencing has identified the immediate precursor to the full motif in Taiwan Aboriginal groups (Trejaut et al., 2005). As shown in figure 4.8, many Taiwan aboriginals (particularly the Amis) have haplogroup B4a1a, with the suite of coding region transitions at 15,746, 12,239, and 6,719. Its coalescence age has been
Mitochondrial DNA Variation in Northern Island Melanesia
Figure 4.8 Phylogenetic trees of recently introduced mtDNA haplogroups in Northern Island Melanesia. See figure 4.4 legend for details of notation. Some estimated dates as shown in table 4.5 are indicated.
estimated from a short segment of the control region as 12,200 ± 4,700 YBP, and from the coding region, 13,169 ± 3,800 YBP—a good agreement. Its descendant, haplotype B4a1a1 with the transition at 14,022, the full “Polynesian Motif,” has not been identified in Taiwan. Note that although 14,022 was not ordinarily sequenced in earlier studies (since it lies outside the control region), B4a1a1 was defined then by the presence of transitions at 16,261 and 16,247, which we found to be hypermutable in our series (and therefore unreliable), as discussed below. In spite of this question mark, it would appear from earlier surveys that B4a1a1 was identified sporadically in central and eastern Indonesia (Wallacea), Madagascar (in descendants of Austronesian speakers from insular Southeast Asia, most probably from the East Barito region of Borneo), and more frequently in Near Oceania. It is extremely common in some, but not all, Polynesian populations. In Near Oceania, this
haplogroup was not observed in the New Guinea highlands (Redd et al., 1995), or in certain Papuan-speaking areas of Northern Island Melanesia, namely among the north Bougainville Aita/Rotokas, and among the Baining and Ata of New Britain (table 4.2 and figures 4.5 and 4.7). These distributions were taken to be consistent with the hypothesis that the haplotype was a marker for an Oceanic/Lapita intrusion. There are, however, several facts that would seem to beg the very close association of the Lapita Cultural Complex with the appearance of this mtDNA variant at ~3,300 YBP. First, its associated age estimates with the “standard” molecular clock methods considerably predate the appearance of Lapita. The coalescence of the fully developed “Polynesian Motif” in Near Oceania (B4a1a1) was estimated by Trejaut et al. to be 9,300 ± 2,600 YBP, and by us at 8,700 ± 2,100 YBP with the inclusion of three additional sequences (in very close agreement),
73
core studies in northern island melanesia
and significantly older than the appearance of Lapita at 3,300 YBP. The restricted Taiwanese precursor (B4a1a) is dated to ∼12,000 YBP (Trejaut et al., 2005), which is again uncomfortably old. Second, its distribution in Northern Island Melanesia does not fit a neat association with Oceanic populations there particularly well at all. While it is absent in some inland Papuan-speaking groups, as mentioned above, its frequency approaches 100% in other Papuan-speaking groups. Nowhere in New Britain, the putative home of Lapita, is this haplogroup nearly as common as in New Ireland and south Bougainville (in far western New Britain, it attained frequencies around 50% in the Kove and the Anêm, another Papuan group). It approaches or supersedes 50% again in parts of Malaita, is less frequent in Vanuatu, and is higher to the east. Dating of the expansion of B4a1a1 is simply not satisfactory at present. Figure 4.9 shows the median joining network for B4a1a in our series, including our small sample of Polynesian and Micronesian (and New Guinea) samples.
The associated expansion ages, using the standard techniques, are given in figure 4.5a, broken down by island and also by method (from the control or coding region). The sample sizes from Remote Oceania are small and unreliable, and more comprehensive coverage would be very helpful in Polynesia and Micronesia, as well as in Indonesia and the Philippines. This haplogroup is the most common in our series, constituting over 30% of the total. As shown, the great majority of samples within this family are the full-blown “Motif” at the core of the median joining network, but a large number of different variants lie a mutation or two away: the perfect signal of a population expansion. There are quite a few haplotypes and samples with a back mutation at nts 16,247 and also 16,261. Both of these were previously considered to be defining sites for the development of the “Polynesian Motif,” but our dataset shows these are not that stable, at least in this region. Because of this, we calculated rho values with and without these two
Figure 4.9 Haplogroup B4a1a1 median joining network. Sample sizes of each haplotype are proportional to their frequencies, as indicated in the upper left corner. Key back mutations are shown. See color insert.
74
Mitochondrial DNA Variation in Northern Island Melanesia
recurrent mutations. All of the estimates are still considerably older than the development of Lapita at 3,300 YBP, whether one includes the most mutable sites or not. From the archeological sequence (chapter 2), one might expect that the estimated expansion dates from Polynesia and Micronesia would be younger than 3,200 YBP, or at least younger than those in Near Oceania, but this is not the case (see table 4.5a, especially highlighted similar dates for Near Oceania and Melanesia/Polynesia). What, then, is a plausible scenario for the appearance and distribution of this B4a1a1 haplotype? Because of the identification of its Taiwanese Aboriginal precursor, its origin clearly is in the Island Southeast Asia/Taiwan area, and its estimated date using “standard” techniques is between 16,000 and 10,000 YBP (B4a1a), and it spread eastward to Wallacea/Near Oceania, where the fully developed Motif (B4a1a1, with 14,022) apparently appeared by at least 8,600 YBP. While this sequence of events clearly indicates B4a1a and the full Motif developed only within Austronesian-speaking groups, the estimated absolute dates are so ancient that they appear unreasonable. Either way, the Motif clearly has had enough time depth in Northern Island Melanesia that it developed a number of distinctive local variants and attained surprisingly high frequencies in some Papuan-speaking groups. Population movements within Northern Island Melanesia subsequent to the development of Lapita may well have altered earlier distributions and associations, so ancient mtDNA information might well help to resolve some of Table 4.5a
these issues. To date, there are no published early Lapita skeletal ancient DNA results that indicate the presence of haplogroup B, although there is a site in Efate, Vanuatu, that holds considerable promise. Finally, the B4a1a1 results (their very high frequency and apparent link between Taiwan/Island Southeast Asia and Polynesia via Northern Island Melanesia) are very much at odds with the Y chromosome results (presented in chapter 5) as well as some other genetic data, the alpha thalassemias data in particular, as described below.
Haplogroup E This haplogroup, which is a subdivision of M9, has been reported sporadically but rarely in Asia. As shown in table 4.1 and 4.2, its various branches have all been found in Taiwan Aboriginal populations, with the notable exception of E1b, which had only been reported prior to our work in ten Indonesian samples, and has not been identified in Southeast Asia. We were surprised to find 85 E samples in our series, 75 of them the rare E1b. Most of these were from New Britain (55 individuals), with 26 from two Papuan groups (Ata and Sulka) and the rest from a selection of Oceanic groups in New Britain, New Ireland, and north Bougainville. As shown in table 4.1, E1a is fairly widespread, whereas E2 has been identified in only Taiwan Aboriginals and Filipinos. The ten E1a and E2 samples that we identified were not found further east than New Britain, in both Papuan and Oceanic groups. Intriguingly, in our small
B4a1a1 Haplogroup Age Estimates from Rho Statistics Ignoring 16247 & 16261
Population
Nf
Bougainville Micronesia New Britain New Guineab New Ireland Polynesiac Solomons South Island Melanesia All samples Near Oceania Melanesia/Polynesiad Melanesia/Polynesiae
93 31 85 51 128 25 154 23 590 10 13
No. Haplotypes 34 11 21 19 19 31 15 123
Method Useda HVS1 HVS1 HVS1 HVS1 HVS1 HVS1 HVS1 HVS1 HVS1 CR CR
Rho
Sigma
0.591 0.484 0.360 0.680 0.711 0.600 0.571 0.522 0.582
0.230 0.264 0.152 0.226 0.325 0.174 0.250 0.222 0.192
Age Estimate (Years)
S.D. (Years)
11,900 9,800 7,300 13,700 14,300 12,100 11,500 10,500 11,700 9,300 8,700
4,600 5,300 3,100 4,600 6,600 3,500 5,000 4,500 3,900 2,500 2,100
Age Estimate (Years)
S.D. (Years)
6,500 7,800 5,400 10,500 7,300 12,100 7,100 7,000 7,800
2,700 5,100 2,800 3,300 3,300 3,500 3,300 2,800 2,000
a
Method used: HVS1—Saillard’s method, with HVS1 (16090-16365), no transversions, mutation rate of 20,180 years per mutation (Forster 1996); CR—Saillard’s method, with coding region (577-16023) with a mutation rate of 5,139 years per mutation (Mishmar 2003). b Published samples (7) were added (Ingman 2003). c Published samples (16) were added (Redd). d Trejaut (2005). e Trejaut (2005) plus 1 Nasioi (Macaulay 2005), and 2 from this study. f Partial sequences are excluded.
75
core studies in northern island melanesia
Table 4.5b
E1b Haplogroup Age Estimates from Rho Statistics
Population
N
No. Haplotypes
All samples Near Oceania All samples Near Oceania
75 4
12
Method Useda
Rho
Sigma
HVS1 CR
0.26
0.129
Age Estimate (Years)
S.D. (Years)
5,200 10,300
2,600 4,800
a
Method used: HVS1—Saillard’s method, with HVS1 (16090-16365), no transversions, mutation rate of 20,180 years per mutation (Forster 1996); CR—Saillard’s method, with coding region (577-16023) with a mutation rate of 5,139 years per mutation (Mishmar 2003).
Rossel Island sample of five (a Papuan-speaking group), both E1a and E2 were represented. E appears to be another relatively late arrival in the region, specifically in New Britain. Age estimates for this lineage based on both the coding region (figure 4.8) and the control region (table 4.5b) are consistent. If anything, it appears younger than the “Polynesian Motif.” Its distribution in Vanuatu and out into Remote Oceania is unclear, since its identification depends on sequencing the HVS1 beyond position 16,390, which was often not included in most HVS1 sequencing. It would appear not to have been carried there, however.
Rare Haplogroups Haplogroups B4b, Y, and M7 are all rare in our series although they are all found west of the Wallace Line in different locales (refer to tables 4.1 and 4.2). All of the M7 samples in our series were from Ontong Java, a “Polynesian Outlier” with heavy Micronesian influence. While nothing can be concluded in terms of their distributions, the presence of these lineages indicates the problems in equating specific populations with specific haplogroup distributions.
Northern Island Melanesia Population Comparisons To analyze the population structure of the mtDNA results, an analysis of molecular variance (AMOVA) was Table 4.6
performed on the 28 groups in Northern Island Melanesia (results shown in table 4.6). The among-group variance represented a very large proportion of the total—almost 30%—reflecting the remarkable population structure in this region. The variation among islands was almost as great as the variation among populations within islands. However, partitioning the variance by the two major language groups (Oceanic vs. Papuan) produced an insignificant between-language group statistic. To visualize the significant distinctions in population relationships, we performed a non-parametric multidimensional scaling on their pairwise FST values (shown in figure 4.10). This two-dimensional plot provides a very good representation of the overall population differences and accounts for 98% of the overall dispersion. While the general island-by-island clustering of the populations is apparent (with New Britain populations clustered to the left top portion, New Ireland to the upper right, and Bougainville populations generally to the lower portion of the plot), the other noteworthy trend is for the Papuan groups to occupy the more extreme or outlier positions in the plot in almost all instances. The New Britain Ata and Baining (Mali and Kaket) form one extreme cluster that contrasts with the New Ireland Kuot and south Bougainville Nagovisi at the other extreme of the first axis; the second axis contrasts the same New Britain Papuan cluster with the north Bougainville Aita and Rotokas. The exceptional population that does not fit these generalizations is the Anêm, who, while Papuanspeaking, fall towards the middle of the distribution.
AMOVA Based on mtDNA HVS1 and HVS2 Variance Components (%)
Grouping No grouping (28 populations) Geography (3 Islands)a Language (2 groups)b Oceanic (16 populations) Papuan (12 populations)
N 941 941 941 467 474
No. of Populations 28 28 28 16 12
Non-significant values are shown in italic. a Groups for geography: New Ireland, New Britain, and Bougainville. b Linguistic groups: Papuans, Oceanic.
76
No. of Groups
Between Groups
Within Groups
Within Populations
1 3 2 1 1
… 14.5 0.2 … …
28.1 18 28 14.8 38.9
71.9 67.5 71.8 85.2 61.1
Mitochondrial DNA Variation in Northern Island Melanesia
Figure 4.10 First two dimensions of Multidimensional Scaling plot, generated from pairwise FST values. See color insert.
One way to understand the MDS distribution is by correlating the population scores on the two MDS axes by the population haplogroup frequencies. The first axis is strongly negatively correlated with the frequency of haplogroup B4a1a1 (r = –0.95) and positively correlated with frequencies of M28 (r = 0.70), which explains the contrast between the New Britain and Kuot/Nagovisi Papuan contrast. Population scores on the second axis are most strongly correlated with Q1 frequencies (r = 0.83), which set the Aita and Rotokas off from the Baining and Ata cluster. The next strongest correlations for the second axis scores are with M27 (r = 0.50) and M28 (r = –0.49).
Discussion and Conclusions Intensive population sampling, combined with both sequencing of the entire hypervariable region and selective whole genome sequencing for newly identified haplogroups, have produced a remarkable description of (maternally mediated) population structure within this small region of Northern Island Melanesia. While the amount of among-group mtDNA variation is very high,
this variation is structured in a fashion that can now be understood. It is the difference among the more remote Papuan-speaking clusters on different islands that drives the pattern of overall mitochondrial variation in the entire region. A triangular contrast between (a) the New Britain Baining and Ata, (b) the north Bougainville Aita and Rotokas, and (c) the New Ireland Kuot and south Bougainville Nagovisi is most apparent. The Oceanicspeaking groups of these islands tend to fall towards the middle of all haplogroup distributions, although there is some remaining distinction between them on an islandby-island basis. The major Papuan distinctions are driven by a combination of very old and somewhat newer haplogroup contrasts. The New Britain Papuan cluster is the center of haplogroup M28; the north Bougainville cluster is the center of M27 and Q1 in this region; and the Kuot/Nagovisi cluster is the center of (the more recently introduced) B4a1a1, and specifically on those haplotypes having the back mutation 16,261. This finding is particularly intriguing, as it suggests that the introduction and divergence of haplogroup B4a1a1 might have occurred thousands of years prior to the appearance of Lapita.
77
core studies in northern island melanesia
Furthermore, these major axes of regional haplogroup variation exist within a context of substantial within-island and also between-island distinctions. As is clear from table 4.2, while there are centers of concentration for each haplogroup, in each case, neighboring populations also show some lower level of these variants, as might be expected with low levels of intermarriage among groups. The major distinctions among the groups relate directly to the distribution of (mostly) very old haplogroups that have not been found to the west of New Britain. Specifically, haplogroups M27, M28, and M29 and a number of specific Q haplotypes appear to have developed in this region well before 20,000 years ago. This pattern is especially apparent with Q2, which has its highest frequencies among the Ata and Baining Mali of New Britain. As argued before, we emphasize that no particular haplogroup can be completely identified with a particular population, with the possible exception of the Nagovisi/Kuot and their variants of haplogroup B4a1a1. For example, while the Baining/Ata cluster is characterized by both high concentrations and diversity of haplogroup M28 variants, Q2 is also common in two of those populations. In addition, both Q1 and M27 characterize the north Bougainville cluster. These two population clusters are clearly quite old, judging from the age estimates associated with the coalescence of their characteristic haplogroups. By contrast, haplogroup B4a1a1 clearly arose in Near Oceania among Proto-Oceanic-speaking populations or their immediate descendants. This inescapable association, along with the firm dating of the appearance of the Lapita Cultural Complex, suggests that the critics of molecular clocks may well be correct. The dates for the entire B4a1 tree could be too old by a factor of at least two, and perhaps more. This could also be true for the older haplogroups in Near Oceania as well. Another possible explanation is that the full Motif and its precursors actually spread into the region considerably before the development of Lapita. These issues will be discussed further in chapter 16. Is there any clear remaining signature of an Austronesian/Proto-Oceanic intrusion? The primary mtDNA characteristic of the Oceanic-speaking populations in this region is that they are less variable and more intermixed than the (generally more interior-distributed) Papuan groups. As shown, while the frequency of B4a1a1 is very high in the region generally (∼30%) it is not relatively high in many contemporary Oceanic groups here, especially in New Britain. It is particularly common today and heterogeneous in New Ireland, south Bougainville, and along the north New Guinea coast, and is less common (and less variable) in New Britain. It has become most diversified in certain Papuan-speaking populations in different islands, which is something of a puzzle.
78
The association of B4a1a1 with the initial settlement of Remote Oceania from Northern Island Melanesia contradicts some other genetic evidence, but is in accord with other sets. It will be shown in chapter 5 that the Y chromosome evidence suggests a vanishingly small Southeast Asian (male) contribution to Remote Oceanic/Polynesian settling populations. Also, alpha thalassemia variation suggests that Polynesians have a preponderance of Near Oceanic ancestry (Martinson et al., 1994). On the other hand, evidence from the human leukocyte antigens (HLAs) has generally been interpreted as suggesting a strong Polynesian tie to Southeast Asia with little Melanesian input (Serjeantson et al., 1982; Serjeantson, 1985, 1989; Mack et al., 2000). Craniometric studies also suggest a strong Polynesian–Asian tie (Pietrusewsky, 2005). We hope to address this issue in a more comprehensive way with a much larger set of variants from across the genome in the near future. Taken together, the mtDNA variation in Northern Island Melanesia reflects the very ancient settlement of the region; the subsequent isolation of many inland populations; some subsequent internal population expansions; the introduction of at least two haplogroups and populations during the Holocene; and considerable intermixture among many groups, especially those living along the shorelines. Because the mtDNA only reflects a very small (exclusively maternal) fraction of the heritage of an individual or population, it may yield a biased result, but this survey also shows its considerable power.
Acknowledgments Gisele Horvat has greatly assisted the mtDNA analysis. The late John McDonough was essential to the early analysis phase. Special thanks also to Fred Gentz, Salvatore Cerchio, Lydia Smith, Frederika Kaestle, and Paul Babb. Financial support was provided by grants from the National Science Foundation, the Wenner-Gren Foundation for Anthropological Research, and the National Geographic Society Exploration Fund.
References Avise JC, Arnold J, Ball RM, Bermingham E, Lamb T, Neigel JE, Reeb CA, Saunders NC. 1987. Intraspecific phylogeography: The mitochondrial DNA bridge between population genetics and systematics. Annual Review of Ecological Systematics 18: 489–522. Cox MP. 2003. Genetic patterning at Austronesian contact zones. Anthropology Department, University of Otago. Dunedin, New Zealand (doctoral dissertation). pp 270. Edmonds CA, Lillie AS, Cavalli-Sforza LL. 2004. Mutations arising in the wave front of an expanding population. Proceeedings of the National Academy of Sciences USA 101: 975–9.
Mitochondrial DNA Variation in Northern Island Melanesia
Forster P, Harding R, Torroni A, Bandelt H-J. 1996. Origin and evolution of Native American mtDNA variation: A reappraisal. American Journal of Human Genetics 59: 935–45. Forster P, Torroni A, Renfrew C, Rohl A. 2001. Phylogenetic star contraction applied to Asian and Papuan mtDNA evolution. Molecular Biology and Evolution 18: 1864–81. Friedlaender JS, Gentz F, Friedlaender F, Kaestle F, Schurr TG, Koki G, Mgone C, McDonough J, Smith L, Merriwether DA. 2005a. Mitochondrial genetic diversity and its determinants in Island Melanesia. In: Pawley A, Attenborough R, Golson J, Hyde R, editors. Papuan pasts: Studies in the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 693–716. Friedlaender JS, Schurr TS, Gentz F, Koki G, Friedlaender F, Horvat G, Babb P, Cerchio S, Kaestle F, Schanfield M, Deka R, Yanagihara R, Merriwether DA. 2005b. Expanding Southwest Pacific mitochondrial haplogroups P and Q. Molecular Biology and Evolution 22: 1506–17. Ho SY, Phillips MJ, Cooper A, Drummond AJ. 2005. Time dependency of molecular rate estimates and systematic overestimation of recent divergence times. Molecular Biology and Evolution 22: 1561–8. Howell N, Elson JL, Turnbull DM, Herrnstadt C. 2004. African haplogroup L mtDNA sequences show violations of clock-like evolution. Molecular Biology and Evolution 21:1843–54. Huoponen K, Schurr TG, Chen Y, Wallace DC. 2001. Mitochondrial DNA variation in an aboriginal Australian population: Evidence for genetic isolation and regional differentiation. Human Immunology 62: 954–69. Ingman M, Gyllensten U. 2003. Mitochondrial genome variation and evolutionary history of Australian and New Guinean aborigines. Genome Research 13: 1600–6. Kivisild T, Bamshad MJ, Kaldma K, Metspalu M, Metspalu E, Reidla M, Laos S, Parik J, Watkins WS, Dixon ME, Papiha SS, Mastana SS, Mir MR, Ferak V, Villems R. 1999. Deep common ancestry of Indian and westernEurasian mitochondrial DNA lineages. Current Biology 9: 1331–4. Kivisild T, Tolk HV, Parik J, Wang Y, Papiha SS, Bandelt HJ, Villems R. 2002. The emerging limbs and twigs of the East Asian mtDNA tree. Molecular Biology and Evolution 19: 1737–51. Klopfstein S, Currat M, Excoffier L. 2006. The fate of mutations surfing on the wave of a range expansion. Molecular Biology and Evolution 23: 482–90. Lum JK, Cann RL, Martinson JJ, Jorde LB. 1998. Mitochondrial and nuclear genetic relationships among Pacific Island and Asian populations. American Journal of Human Genetics 63: 613–24. Macaulay V, Hill C, Achilli A, Rengo C, Clarke D, Meehan W, Blackburn J, Semino O, Scozzari R, Cruciani F, Taha A, Shaari NK, Raja JM, Ismail P, Zainuddin Z, Goodwin W, Bulbeck D, Bandelt HJ, Oppenheimer S, Torroni A, Richards M. 2005. Single, rapid coastal
settlement of Asia revealed by analysis of complete mitochondrial genomes. Science 308: 1034–6. Mack SJ, Bugawan TL, Moonsamy PV, Erlich JA, Trachtenberg EA, Paik YK, Begovich AB, Saha N, Beck HP, Stoneking M, Erlich HA. 2000. Evolution of Pacific/Asian populations inferred from HLA class II allele frequency distributions. Tissue Antigens 55: 383–400. Martinson JJ, Boyce AJ, Clegg JB. 1994. VNTR alleles associated with the alpha-globin locus are haplotype and population related. American Journal of Human Genetics 55: 513–25. Melton T, Clifford S, Martinson J, Batzer M, Stoneking M. 1998. Genetic evidence for the proto-Austronesian homeland in Asia: mtDNA and nuclear DNA variation in Taiwanese aboriginal tribes. American Journal of Human Genetics 63: 1807–23. Merriwether DA, Hodgson JA, Friedlaender FR, Allaby R, Cerchio S, Koki G, Friedlaender JS. 2005. Ancient mitochondrial M haplogroups identified in the Southwest Pacific. Proceedings of the National Academy of Sciences, USA 102: 13034–9. Metspalu M, Kivisild T, Metspalu E, Parik J, Hudjashov G, Kaldma K, Serk P, Karmin M, Behar DM, Gilbert MT, Endicott P, Mastana S, Papiha SS, Skorecki K, Torroni A, Villems R. 2005. Most of the extant mtDNA boundaries in South and Southwest Asia were likely shaped during the initial settlement of Eurasia by anatomically modern humans. BMC Genetics 6: 41. Mishmar D, Ruiz-Pesini E, Golik P, Macaulay V, Clark AG, Hosseini S, Brandon M, Easley K, Chen E, Brown MD, Sukernik RI, Olckers A, Wallace DC. 2003. Proceedings of the National Academy of Sciences, USA 100: 171–6. Oppenheimer S, Richards M. 2001. Polynesian origins: Slow boat to Melanesia? Nature 410: 166–7. Palanichamy MG, Sun C, Agrawal S, Bandelt HJ, Kong QP, Khan F, Wang CY, Chaudhuri TK, Palla V, Zhang YP. 2004. Phylogeny of mitochondrial DNA macrohaplogroup N in India, based on complete sequencing: Implications for the peopling of South Asia. American Journal of Human Genetics 75: 966–78. Pietrusewsky M. 2005. The physical anthropology of the Pacific, East Asia, and Southeast Asia: A mutlivariate craniometric analysis. In: Sagart L, Blench R, Sanchez-Mazas A, editors. The peopling of East Asia: Putting together archaeology, linguistics, and genetics. London and New York: Routledge Curzon. pp 201–29. Redd AJ, Stoneking M. 1999. Peopling of Sahul: mtDNA variation in Aboriginal Australian and Papua New Guinean populations. American Journal of Human Genetics 65: 808–28. Redd AJ, Takezaki N, Sherry ST, McGarvey ST, Sofro ASM, Stoneking M. 1995. Evolutionary history of the COII/tRNALys intergenic 9 base pair deletion in human mitochondrial DNAs from the Pacific. Molecular Biology and Evolution 12: 604–15. Richards M, Oppenheimer S, Sykes B. 1998. mtDNA suggests Polynesian origins in Eastern Indonesia. American Journal of Human Genetics 63: 1234–6.
79
core studies in northern island melanesia
Saillard J, Forster P, Lynnerup N, Bandelt H-J, Norby S. 2000. mtDNA variation among Greenland Eskimos: The edge of the Beringian expansion. American Journal of Human Genetics 67: 718–26. Serjeantson S. 1985. Migration and admixture in the Pacific: Insights provided by human leukocyte antigens. In: Kirk RL, Szathmary E, editors. Out of Asia: Peopling the Americas and the Pacific. Canberra, Australia: Australian National University Press. pp 133–54. Serjeantson S. 1989. HLA genes and antigens. In: Hill AVS, Serjeantson SW, editors. The colonization of the Pacific: A genetic trail. New York: Oxford University Press. pp 120–35. Serjeantson SW, Ryan DP, Thompson AR. 1982. The colonization of the Pacific: The story according to human leukocyte antigens. American Journal of Human Genetics 34: 904–18.
80
Sun C, Kong QP, Palanichamy MG, Agrawal S, Bandelt HJ, Yao YG, Khan F, Zhu CL, Chaudhuri TK, Zhang YP. 2006. The dazzling array of basal branches in the mtDNA macrohaplogroup M from India as inferred from complete genomes. Molecular Biology and Evolution 23: 683–90. Tajima A, Hayami M, Tokunaga K, Juji T, Matsuo M, Marzuki S, Omoto K, Horai S. 2004. Genetic origins of the Ainu inferred from combined DNA analyses of maternal and paternal lineages. Journal of Human Genetics 49: 187–93. Thangaraj K, Chaubey G, Kivisild T, Reddy AG, Singh VK, Rasalkar AA, Singh L. 2005. Reconstructing the origin of Andaman Islanders. Science 308: 996. Trejaut JA, Kivisild T, Loo JH, Lee CL, He CL, Hsu CJ, Li ZY, Lin M. 2005. Traces of archaic mitochondrial lineages persist in Austronesian-speaking formosan populations. Public Library of Science. Biology 3: e247.
5 Y Chromosome Variation in Northern Island Melanesia Laura B. Scheinfeldt, Françoise R. Friedlaender, Jonathan S. Friedlaender, Krista Latham, George Koki, Tatiana Karafet, Michael Hammer, Joseph Lorenz
Introduction The Y chromosome provides the paternal counterpoint to the maternally inherited mtDNA in the study of human variation. One of the characteristics unique to the Y chromosome is that only two small portions recombine with the X chromosome (pseudoautosomal regions 1 and 2), both of which together consist of only 3 million base pairs (Mb). The remaining 57 Mb of the Y chromosome is nonrecombinant (NRY) with other chromosomes, and inherited solely from father to son. Therefore, most of the chromosome can only accumulate variation through mutation, unlike autosomal chromosomes where variant combinations are also created through recombination between maternal and paternal chromosomes. Furthermore, in any given parental couple, there exist four copies of each autosomal chromosome, three copies of the X chromosome, and only one Y chromosome copy. Therefore, the effective population size for Y chromosomes is significantly reduced relative to other nuclear chromosomes. Another important feature is that selective pressure on the nonrecombinant portion of the Y chromosome is thought to be low (e.g. Hammer et al., 1997). These combined characteristics make the study of NRY variation essential in the understanding of male-mediated population structure and history. The NRY includes 78 protein-coding genes (27 distinct Y chromosome proteins), over 200 well-characterized single nucleotide polymorphisms (SNPs), and over 30 microsatellite repeat regions (trinucleotide, tetranucleotide, and
pentanucleotide). SNP polymorphisms tend to be unusual and unique events; that is, there is essentially no recurrence of these mutations. On the other hand, the mutation rate at microsatellite or short tandem repeat (STR) loci is relatively high. As a result, SNP mutations have been used to construct the primary NRY lineage tree in global population comparisons (Y Chromosome Consortium, 2002). Microsatellite variation is ordinarily analyzed within SNP haplogroups to compare their relative diversities, with higher microsatellite diversity interpreted to represent older ages. Additionally, the geographic location of the highest haplogroup diversity is generally taken to represent its origin. The analysis of NRY SNP variation in human populations globally has yielded a hierarchical framework or a nested set, just as with mtDNA (e.g. Underhill et al., 2001b). This NRY chromosome phylogeny, like that for the mtDNA, has its deepest root in Africa, and all nonAfrican populations descend from a single shallower branch (delineated by the M168 SNP). The non-African branches (shown in figure 5.1) can be further subdivided into macrohaplogroups, each with their own subbranches. Figure 5.1 shows the biallelic markers analyzed in our study and their hierarchical relationships (Ellis and Hammer, 2002; Kayser et al., 2003; Wilder et al., 2004). We followed the recently standardized Y Chromosome Consortium (YCC) nomenclature (Ellis and Hammer, 2002; Jobling and Tyler-Smith, 2003). In addition, we have adopted the more recent use of the * to denote a paragroup,
81
core studies in northern island melanesia
Figure 5.1 Phylogeny of non-African Y chromosome single nucleotide polymorphisms.
or a haplogroup that is defined by the presence of a defining SNP marker and the absence of any tested sublineages. The * is only employed when discussing a specific paragroup as opposed to the SNP polymorphism itself.
Prior Findings on NRY Variation in Southeast Asia and the Pacific Previous NRY studies in Southeast Asia have identified four NRY haplogroups that are particularly common or informative in Island Southeast Asia and Near Oceania with regard to past population dynamics: O, M, Km230, and C2 (Capelli et al., 2001; Kayser et al., 2000, 2001a, 2003). As shown in table 5.1, the distribution of the O lineage encompasses all of Asia while the M, Km230, and C2 lineages are regionally restricted to Near and Remote Oceania, and have almost never been identified east of the Wallace Line. Therefore, O is presumed to have arisen in Asia while M, Km230, and C2 are thought to have arisen in Near Oceania (Kayser et al., 2001a, 2003; Shi et al., 2005). Prior reports of lineage frequencies in Polynesian populations are not directly comparable because of differences
82
in the batteries of SNPs tested. However, lineage C2 was consistently the most common Polynesian NRY lineage, while M and O have been found in lower but variable frequencies (Capelli et al., 2001; Kayser et al., 2003; Underhill et al., 2001a). This distribution supports a model of significant genetic contribution of New Guinea or Island Melanesian populations on the Polynesian expansion with a lesser degree of SE Asian influence, contrary to the commonly accepted mtDNA account (see chapter 4).
Our Study Findings NRY sampling from the Bismarck Archipelago and Bougainville was poor prior to our intensive survey. Only 16 Tolai-speakers from the Gazelle Peninsula of New Britain had been covered (Kayser et al., 2003). Our survey sample complements the mtDNA series described in chapter 4, so that the populations included the same language groups and dialects from Northern Island Melanesia that were sampled in the field seasons 1998, 2000, and 2003. However, we could not cover the additional populations included in the wider mtDNA survey
Table 5.1
Chromosome Lineage Frequencies of Island Melanesia and Nearby Regions Haplogroup Frequencies (%)
Asia Korea China Southeast Asia Vietnam Taiwan
Highlands Lowlands/Coast Coast Highlands Highlands Coast
M89 F*
RPS4Y M38 C* C2*
M208 M217 390.1d M9 C2b C3 K*
25 36
8 0
0 0
0 0
0 0
12 6
0 0
8 11
0 0
11 26 43 53 50 50 53 40 115 50 53 40 551 55 34 31
0 0 0 0 0 0 0 0 1 6 2 5 0 0 0 0
0 0 0 0 0 0 0 2 4 2 2 0 2 6 9 7
0 0 0
0 0 0
9 4 0
0 0 0
0 0 0
? 0 0 0 3 0 27 15 16
? 0 0 0 0 0 0 0 0
? 1 4 0 3 0.2 0 0 0
0 0
0 4 0 0 0 0 0 0 20 8 2 10 1 20 18 32
94 89 31 31 7 25
0 0 0 0 0 0
0 0 3 0 0 0
0 9 13 3 0 8
25 0 10 0 0 28
0 0 0 0 0 0
0 0 0 0 0 0
0 11 16 7 29 16
0 0 0 0
M230 M4 K5 M*
P22 M175 M119 M95 M2a O* O1a O2a
M122 O3
M74 P
M173 R Other
0 0
0 0
32 0
4 22
0 3
28 58
4 0
4 0
a a
0 0 0
0 12 5 4 0 0 4 0 0 0 2 5
36 0 5 0 24 0 6 28 3 32 42 38 59
0 0
0 0
46 58 12 0 2 2 47 0 39 30 23 18 7 9 12 3
0 0 0 0 0 0 0 0 0 0 2 3 0
? 0 0
9 23 79 96 74 98 43 70 28 8 23 15 18 13 6 23
0 0 0
2 0 0 0 0.4 13 21 10
0 0 0 0 0 0 0 0 0 2 0 0 1 13 21 7
0 0
0 3
a a a c c c c c a,c,f a,f a a e f a a
0 2 16 52 57 32
75 78 29 36 14 12
0 0 0 0 0 4
0 0 0 0 0 0
0 0 0 3 0 0
0 0 0 0 0 0
1 0 10 0 0 0
0 0 3 0 0 0
0 0 0 0
0 0 0 0
2 4 4 3 4
5
8
Source1
a a a a this study this study Continued
83
Y Chromosome Variation in Northern Island Melanesia
Philippines Malaysia Java Southern Borneo Balinese East Indonesians Moluccas Nusa Tenggara New Guinea WNG WNG PNG PNG PNG
Chinese Aborigines Paiwan Bunun Atayal Amis Yami
N
Chromosome Lineage Frequencies of Island Melanesia and Nearby Regions—cont’d Haplogroup Frequencies (%) N
Island Melanesia Trobriand Manus New Britain Mussau New Hanover New Ireland Bougainville Vanuatu Fiji Polynesia Tonga WesternSamoa Cook Islands Atiu FrenchPolynesia Maori Australia Arnhem Desert
Islands East West
North Central
M89 F*
RPS4Y M38 C* C2*
M208 M217 390.1d M9 C2b C3 K*
53 7 145 245 20 43 109 54 18 234 55
0 0 0 0 0 0 0 7 0 0 0
0 0 0 0 0 0 0 0 0 18 3
0 0 3 1 0 0 7 2 0 ? ?
9 14 1 0.4 5 0 1 0 0 ? ?
0 0 0 0 0 0 0 0 0 ? ?
55 16 28 42 87 54
0 0 0 0 0
23 69 0 84 53 0
? ? 0 ? ? 42
? ? 82 ? ? ?
? ? 0 ? ? 0
60 35
2 3
10 0
0 0
0 0
0 0
0 0 0 0 0 0 0 0 0 ? ?
0
53 69
M230 M4 K5 M*
P22 M175 M119 M95 M2a O* O1a O2a
23 29 57 64 75 12 29 48 17 41 41
0 29 8 3 0 2 6 2 0 6 ?
30 0 11 7 5 12 21 2 0 30 15
0 14 15 16 15 70 27 18 83 ? ?
0 0 1 0 0 2 5 4 0 0
1 6 4 1 8 4
? ? 0 ? ? ?
8 0 0 0 0 ?
?
0 0 0 0 0
30 17
0 0
0 0
0 0
0
0 0
28 14 1 3 0 2 4 15 0 0 6
0 0 ? 0 0 ? ? ? 0 0
2 6 0 0 2
0 6 0 0 0
0 0
1 data sources a: Kayser et al. (2003), b: Cox (2003), c: Capelli et al. (2001), d: Underhill et al. (2001), e: Karafet et al. (2005), f: Hammer et al. (2005). ?, untested.
0 0
M122 O3
M74 P
M173 R Other
9 0 3 5 0 0 0 2 0 4 9
0 0 0 0.4 0 0 0 0 0 0
0
58 13 7 3 35 5
0 0 0 3 2
0 3
0 0
?
2 26 8 7 9 33 5 9
16
Source1 a this study this study this study this study this study this study this study this study b c c c a c c d a a
core studies in northern island melanesia
84 Table 5.1
Y Chromosome Variation in Northern Island Melanesia
because these were represented by old plasmas dating back to 1966. We could not recover adequate Y DNA from those samples for analysis. For the analysis, we used a panel of 23 Y chromosome markers. Sixteen were biallelic markers: M89, M9, RPS4Y, 50f2/c, M175, M4, M230, M74, M38, M119, M122, P22, M208, P79, P117, and P87. The last three have only been recently defined (Scheinfeldt et al., 2006), and add a great deal of information on the population structure in this region. Seven microsatellite markers were analyzed because they had already been shown to be regionally informative: DYS19, DYS389I, DYS389II, DYS390, DYS391, DYS392, DYS393 (Forster et al., 1998; Kayser et al., 2001b). Details on the analytic methodology can be found in Scheinfeldt et al., 2006.
SNP Haplogroup Distributions While table 5.1 summarizes the frequencies of the NRY SNPs that have commonly been screened in populations
across Southeast Asia and Oceania; table 5.2 presents the complete battery of SNP results for the populations in our Northern Island Melanesian series, as well as for a set of New Guinea men who had married into the island communities. Figure 5.2 shows the phylogenetic tree of the SNP data in table 5.2, color-coded by island. Figure 5.3 shows the geographical distributions of the SNP haplogroups by population in our series. Even at first inspection, the SNP variation across Northern Island Melanesia was remarkable. Not only did New Guinea, New Britain, New Ireland, and Bougainville have very different SNP signatures, but SNP frequency variation within islands was also particularly clear, most especially for New Britain, the largest and best-sampled island in the survey. Lineage C-RPS4Y. As mentioned, C is considered the oldest lineage in the Southwest Pacific, likely introduced with the first settlers of both the Sunda shelf and the ancient continent of Sahul, with an estimated origin time of perhaps 50,000 YBP (Underhill, 2004).
Figure 5.2 Evolutionary tree for the 14 major NRY haplogroups in Island Melanesia. See color insert.
85
core studies in northern island melanesia
Table 5.2
Chromosome Lineage Frequencies in Island Melanesia Haplogroup Frequencies (Individuals)
Island Region New Guinea PNG Coast
PNG Highlands
PNG Island Manus New Britain West New Britain
East New Britain
Mussau New Hanover New Ireland New Ireland (O)
New Ireland (P) Bougainville North Bougainville
Central Bougainville South Bougainville
Populationa
Languageb
N
North Coast Markham Rigo Eastern Highlands Morobe Highlands Western Highlands Misima
O O,P O P P P O
25 1 2 4 1 2 2 7
Kove* Anêm* Mangseng* Mamusi* Nakanai* Loso (Nakanai)* Mengen* Melamela* Ata* Kol Tolai* Sulka* Mali (Baining)* Kaket (Baining)*
O P O O O O O O P P O P P P
Lavongai*
O
24 34 11 43 36 15 23 14 45 4 49 33 24 39 20 43
Tigak* Nalik* Notsi* Madak* Patpatar Kuot*
O O O O O P
21 17 14 19 6 32
Saposa* Teop* Buka* Aita* Nagovisi Siwai
O O O P P P
26 18 10 18 1 2 685
M89 F*
M38 C2* 2
M208 C2b 7
M9 K*
M230 K5
P79 K6
2
8
2
1
1
1
2
4
2
1
2 2 2 2
1 3 1 1 2 1 8 5 5 10 8 4 6 1 13 3 10 17 1 2 5 5
1 2
1
3 1
2 3 17 8 5 9 5 12
1
2 10
1
8 6 1 16 10
1
2 4 1 4
3 1
2
1
4 4
5
1
6
1
4
14 6
1
2 1 1
3
4
17
14
1 150
41
124
a
Core populations denoted with * were used in the gene diversity, average gene diversity, and AMOVA analysis. Languages: O, Oceanic; P, Papuan.
b
The C3 division is Asian, found as far to the east as Borneo, and a special branch of C (390,1del) occurs exclusively in Aboriginal Australia (Kayser et al., 2003). Lineage C2-M38. Besides the unique Aboriginal Australian lineage shown in table 5.1, there are separate branches of C that are native to Near Oceania. While C2
86
has been identified as far west as Borneo, its major concentration appears to be in eastern Indonesia and coastal New Guinea (Kayser et al., 2003). The 17 C2-M38* samples identified in our series were primarily found in New Ireland (and in the Tolai, who migrated centuries ago from New Ireland to East New Britain).
Y Chromosome Variation in Northern Island Melanesia
Haplogroup Frequencies (Individuals) P117 K7
M4 M*
P87 M2*
2 1
1
P22 M2a
M175 O*
M119 O1a
M122 O3
M74 P
1
1
1 1 16 1
7 5
8
1
1 1 4 5
2 2 1 11 2 5 2
1 1
4
4 1
1
5
2 3
2
1
15
1
9 3 8 2 3 30
1 4 8 9
6 9 1 5 2
1 3 6
2 10 1 2 1
60
24
46
8 7 3 3 1 7 2 4 4 15 1 1 153
5
Standard Gene Diversity (SNP) ˆ H
SD
Average Gene Diversity (STR) ave πn
SD
0.823 -
0.050 -
0.679 -
0.383 -
0.797 0.742 0.764 0.732 0.848 0.752 0.767 0.747 0.747 0.855 0.695 0.743 0.738 0.695 0.496
0.050 0.064 0.107 0.035 0.028 0.056 0.054 0.066 0.029 0.018 0.071 0.043 0.042 0.081 0.085
0.553 0.485 0.681 0.579 0.621 0.652 0.585 0.502 0.619 0.697 0.659 0.528 0.576 0.607 0.543
0.321 0.283 0.406 0.327 0.349 0.380 0.337 0.305 0.346 0.384 0.369 0.308 0.327 0.351 0.310
1
2
1
1
1 2
2 1 1
0.833 0.787 0.857 0.825 0.000 0.815
0.065 0.075 0.056 0.048 0.000 0.035
0.645 0.659 0.752 0.634 0.000 0.671
0.369 0.381 0.434 0.366 0.000 0.375
1 1
3 3 2
1
0.686 0.837 0.844 0.294 -
0.087 0.057 0.103 0.119 -
0.641 0.626 0.523 0.444 -
0.363 0.372 0.339 0.270 -
10
23
18
1 1
One north Bougainville Teop (Austronesian) was also C2-M38*. Lineage C2b-M208. This subdivision of C has been interesting because it was found in high frequencies in the highlands of West Papua and in 23 of the 28 typed Cook Islanders (Kayser et al., 2000, 2003). As a result, it
1
was taken as clear evidence for the heavy Melanesian (New Guinea) origin of Polynesian males. However, only 14 C2b samples were found in our series, seven of which were from the Sepik region of New Guinea. Of the six C2b individuals from our Island Melanesian series, five were from Austronesian-speaking groups. If there is a
87
core studies in northern island melanesia
Figure 5.3 Y chromosome haplogroups and their frequency distribution in Northern Island Melanesia populations. See color insert.
C2b link with Lapita and the colonization of the Remote Pacific, this represents a surprisingly weak signal in the Bismarcks, the apparent home of Lapita. Lineage C4-P55. This branch was found in one sample from the New Guinea highlands in our series (not shown, but included in diversity calculations). The K lineage was by far the most common clade, but we have been able to detect a number of haplogroups within it. Lineage K-M9*. The M9 polymorphism has been found widely across Eurasia, the Americas, and Aboriginal Australia. In our series, K-M9* is the residual haplogroup defined by the presence of M9 and the absence of the other haplogroup-defining polymorphisms within it. In the Southwest Pacific, K-M9* is particularly
88
common in Wallacea, the Trobriand Islands, sections of New Britain and New Ireland, and Fiji. We made a particular effort to identify new polymorphic subdivisions within K-M9*, discussed below as K6 and K7. Lineage K5-M230. This was one of the first polymorphisms found to sub-classify K-M9* in Melanesia (Kayser et al., 2003). It has been found as far west as Bali, where it is rare, in eastern Indonesia (10–20%), but it reaches its greatest frequency in the highlands of Papua New Guinea (52%), before declining in frequency in Island Melanesia. It has not been found in those Polynesian samples that have been tested (Karafet et al., 2005; Kayser et al., 2003). It therefore appears to be a “New Guinea highlands” marker with a likely origin there. A large portion of our New Guinea samples (32%) also fall into this
Y Chromosome Variation in Northern Island Melanesia
lineage (primarily among Sepik men), and it is relatively common among the Tolai and New Ireland populations, but almost absent in Bougainville. Lineage K1-M177. K1-M177 has generally been considered a private polymorphism, only identified in a single Papuan-speaking Nasioi individual collected in a cell line by JSF from the Nasioi of Bougainville (Whitfield et al., 1995). When it was subsequently incorporated in Pacific population surveys, it had not been found (Cox, 2003), until another K1 individual was identified in Malaita, Solomon Islands (Cox and Lahr, 2006). In our current survey, which included sampling in north Bougainville, we were unable to find it, so that it remains rare but now intriguing in its narrow distribution. Finding a tie between Bougainville and Malaita is not unreasonable since trading in shell money and other traditional valuables occurred between the islands over the last two centuries at a minimum. The two cases of K1-M177 are not shown in the tables or figures. Lineage K6-P79, recently identified (Scheinfeldt et al., 2006), has not been screened widely, but was common in our series. It was infrequent along the New Guinea coast, but common among the Mussau and certain inland New Britain populations: two Papuan-speaking (Ata and Kaket Baining), and one inland Oceanic-speaking group (Mamusi). It also had an appreciable frequency in central New Ireland, including the Papuan-speaking Kuot. It was uncommon in Bougainville. Lineage K7-P117 had a more restricted distribution than K6. Because it also was only recently characterized, it has not been screened outside our series. Except for three samples from north Bougainville, it was limited to New Britain populations, and undetected in New Ireland. K7 was particularly common in certain Papuan groups. In the Anêm, where K6 is low, 47% were K7. There was a similar preponderance of K7 to K6 frequencies in the Baining Mali. In contrast, K7 was absent in a number of groups where K6 predominated (the Mamusi, Ata, and Mussau). 50f2/c (data not shown). We discovered that the 50f2/c deletion was associated with three different haplogroups within the K lineage in our series: some (but not all) K*, K6, and K7 samples had this deletion. It occurred almost exclusively in New Britain samples. Because of its peculiar distribution and clear recurrence, we have not incorporated the 50f2/c results in further analyses. The M division of the K lineage is known to be heavily Near Oceanic in its distribution, found only in a very few samples to the west of the Wallace Line (see table 5.1). Again, its subdivisions have not been regularly screened, but from our series its diversity suggests, if not an origin in Near Oceania, then certainly its major expansion and diversification there (table 5.2). Lineage M-M4*. Twenty-four samples were assigned in our series to haplogroup M*. These were scattered
widely, from New Guinea (Sepik, Markham Valley, and Eastern Highlands) to New Ireland. Its highest frequency was in the Papuan-speaking Anêm, along with their Kove neighbors in far western New Britain. Lineage M2-P87*. M2* is another recently identified haplogroup that in this case subdivides the M clade, and was found in 46 individuals in our series. It was almost entirely restricted to New Ireland and east New Britain, and was most common in the Papuan-speaking Kuot and Baining groups there. It was also found in their immediate neighbors in lower frequencies. Lineage M2a-P22. This lineage was the most frequently occurring haplogroup in our population (153 positive samples). It occurred in particularly high frequencies in the isolated Papuan-speaking Aita of Bougainville (>80%), in New Hanover (70%), and in frequencies ~30% in parts of New Ireland and New Britain (the Papuan-speaking Ata in particular). In contrast, it occurred in only one of the samples from New Guinea. The O lineage is ubiquitous in East Asia, as table 5.1 suggests. All branches of O were rare in our sample, but when they occurred, they were almost exclusively found in Oceanic-speaking groups. Because there was no great difference in their distributions in our series, the lineages of O will be discussed together. Three lineages of O were distinguished—lineage O-M175, lineage O1a-M119, and O3-M122—which decline in frequency throughout Southeast Asia, Aboriginal Taiwan, Indonesia, and Melanesia. O2a is very common in Southeast Asia, especially Bali, but it is almost absent in Melanesia and Polynesia (it has been found in one Western Samoan). O1a and O3 have been found in frequent and often equal proportions throughout Island Southeast Asia. However, O3 is the most frequent O haplogroup in Polynesia (over 50% in Tonga, and 35% in French Polynesia), while the other O branches are either rare or lost there. Interestingly, among Taiwan Aboriginals, O3 is common only among the Amis, who have been suggested to be the ancestors of the branches of all Austronesians outside Taiwan (Capelli et al., 2001; Trejaut et al., 2005; also refer to table 5.1). However, both O3 and O1a are rare throughout New Guinea and Northern Island Melanesia. In our series, no O branches were common. Where they did occur, they almost always occurred together in Oceanicspeaking groups. Among the Papuan-speaking series, one Kuot was O, and three Sulka were either O or O1a. The Sulka are known to have been heavily influenced linguistically and otherwise by their Oceanic-speaking neighbors (Reesink, 2005). Taken together, the O lineages had their highest frequencies in Bougainville, but O3 was concentrated in only a few New Britain Austronesian-speaking groups along the north coast, which has been thought to be the specific Lapita homeland.
89
core studies in northern island melanesia
To summarize the NRY SNP distributions, while some haplogroups were scattered through the general region (K, K5, M, and to a lesser extent the C haplogroups), many appeared to be specific to different sections of Northern Island Melanesia, where they likely had their origins (K6, K7, M2*, and M2a). This latter set was rare or absent in our New Guinea series. In contrast, O3 (and other O lineages) were closely identified with certain Oceanic-speaking groups, but the O lineages were surprisingly scarce. The measure of SNP heterogeneity calculated for the populations in our series, where sample sizes were 10 or more, is presented as the gene diversity (SNP) and average gene diversity (STR) for all haplogroups/haplotypes within each population sample in table 5.2. These values and their standard deviations were calculated using Arlequin 3.0 (Excoffier, 2005). The most homogeneous group (with the lowest diversity values) was the Papuanspeaking Aita of Bougainville, who were until recently very isolated and subject to considerable genetic drift, while the most heterogeneous sample (the highest diversity values) was from the Notsi of New Ireland. Otherwise, there was no apparent pattern of heterogeneity to the distribution of diversities within the sample series, especially given their relatively large standard deviations. Comparing values of diversity across studies, or simply values of segregating sites, is difficult since SNP panels differ so widely, and in our case, we had added three regionally polymorphic SNPs. The primary finding here is that, because we have been able to subdivide the K and M macrohaplogroups with newly defined SNPs, our estimation of haplogroup diversities is higher than earlier studies had suggested.
Microsatellite (STR) Measures of Diversity and Age As mentioned, we chose the seven STRs for typing because they had been shown to be informative in earlier Pacific studies. Median joining networks for the STRs were calculated on each SNP-defined haplogroup background using NETWORK 4.1.1.2 (Bandelt et al., 1999). Not all networks showed expansions, since some haplotypes were poorly populated, and others apparently incorporated more than one undetected lineage. Those that were indigenous to Near Oceania, on the other hand, gave clear signals of expansions and their median networks are shown in figure 5.4. These are haplogroups K5, K6, K7, M2 (including M2a). They all show typical starlike networks associated with population expansions and most have smooth pairwise difference distributions (not shown). The exception is possibly K7, which is more ragged in its mismatch distribution, and may contain more than one haplogroup. By comparison, O and O3
90
have very ragged mismatch distributions (not shown), suggesting they arose elsewhere. Age estimates from microsatellite diversities are necessarily problematic, given the difficulties in estimating an appropriate mutation rate and the large standard errors inherent in any dating method. We can, however, get a clear sense of the relative ages of those haplogroups that are autochthonous to the region. The rank of increasing diversity (and therefore estimated age) of the native New Guinea and Northern Island Melanesian haplogroups is C2, C2b, M, K6, K7, K5, M2a, and M2, as shown in table 5.3. The associated age estimates for these Melanesian branches are within the timeframe 32,000–50,000 YBP. This conforms to our calculations for a number of ancient mitochondrial DNA expansions specific to the region (Friedlaender et al., 2005). It is also consistent with the currently accepted earliest settlement dates for the region, of ~42,000 YBP (Groube, 1986; Leavesley et al., 2002). The SE Asian haplogroups O and O3 are the youngest in the current analysis. The small set of O3 microsatellite samples yielded an estimated expansion date of ≈20,000 YBP. This can be dismissed as either unreliable because of the small sample, or as mentioned, it could suggest that the O3 immigrants from Southeast Asia had diversified well before they arrived in Northern Island Melanesia by 3,200 YBP. Also note that M2 ranks younger than M2a, contrary to the expected relative age of a sublineage. This can most likely be explained by a number of things including: these haplogroups share the large majority of individuals, the standard errors are large, and both mutations probably arose around the same time.
Partitioning the NRY Variance by Island and Language We used Arlequin 3.0 (Excoffier, 2005) for the Analysis of Molecular Variance (AMOVA) to analyze the structure of NRY variation across the three major island samples— New Britain, New Ireland, and Bougainville. We wanted to know if molecular variation among islands was a significant component of the structuring. If so, was there an effect of island size? Also, did language classification (Papuan or Oceanic) remain a significant factor as well? Table 5.4 gives the AMOVA results. Over 80% of the variance was within the different populations in the series (except for Bougainville, which will be discussed below). The proportion of the variance among groups within islands was also significant at the 0.01 level, and the remaining variance component among the three islands was just short of significance at that level (P = 0.013). Clustering populations by language affiliation (Austronesian vs. Papuan or otherwise) produced non-significant variance components between language groups. As for the effect of island landmass on internal variation, which would be predicted with
Y Chromosome Variation in Northern Island Melanesia
Figure 5.4 Median-joining microsatellite networks for Island Melanesian haplogroups. See color insert. Table 5.3 Y Chromosome Haplogroup Age Estimates from STRs Lower Limitsc (Years)
Upper Limitsc (Years)
Haplogroup
n
ASDa
TMRCAb (Years)
C2 C2b
31 14
1.369 1.276
49,600 46,200
42,000 39,000
61,000 57,000
K5 K6 K7
41 124 60
0.925 1.025 0.978
33,500 37,100 35,400
28,000 31,000 30,000
41,000 46,000 44,000
M M2 M2a
223 199 150
1.052 0.904 0.925
38,100 32,700 33,500
32,000 28,000 28,000
47,000 40,000 41,000
51 18
0.831 0.532
30,100 19,300
25,000 16,000
37,000 24,000
O O3 a
ASD (Average Square Distance) was calculated using the program Ytime (Behar et al., 2003). b TMRCA (Time of the Most Recent Ancestor) estimates were calculated using ASD. with the mutation rate of [2.8 ± 0.5] x 10-5 per locus per year (Zhivotosky, 2004). c Limits were calculated based on the Standard Error from the above mutation rate.
a simplistic biogeographic model, New Ireland, which is long and narrow, did have the least significant withinisland, among-group, variation. However, New Britain, the largest island, only showed a modest increase in withinisland, among-group, variation. The four Bougainville populations produced a far larger within-island, among-group haplogroup variance component than either of the other two islands, indicating the importance of other less simplistic effects such as genetic drift, especially when considering the Aita, which is the most homogeneous group sampled (the Aita males belong to only two haplogroups). This uniformity in combination with the relatively smaller sample analyzed from Bougainville explains the unusually high within-group variance.
NRY Population Relationships Using Multidimensional Scaling As a corollary to the AMOVA analysis, SPSS (version 13) was used for non-parametric multidimensional scaling (MDS) to provide a two-dimensional representation of
91
core studies in northern island melanesia
Table 5.4a
AMOVA Based on Y-chromosomal Haplogroups (Fst) Variance Components
Group
No. of Populations
n
No. of Groups
Between Groups
Within Groups
Within Populations
No grouping (23 populations) Geography (3 Islands)a, b Language (2 groups)c
608 608 608
23 23 23
1 3 2
… 8.4 −0.6
15.9 10.7 16.3
84.1 80.9 84.3
New Ireland New Britain Bougainville
146 390 72
6 13 4
1 1 1
… … …
9.4 9.7 27.5
90.6 90.3 72.5
Papuans Oceanians
225 383
7 16
1 1
… …
18.6 14.8
81.4 85.2
Non-significant values are shown in italic. a Groups for geography: New Ireland, New Britain and Bougainville. b For this AMOVA analysis Lavongai was included with New Ireland. c Linguistic groups: Papuans, Oceanians.
the pairwise population Φst values (figure 5.5). The plot has a very low stress level (0.01), indicating it is a reliable representation of the pairwise distances. The main features of the MDS display are: New Britain populations generally fall to one side of the plot, while New Ireland populations fall to the other; the four Bougainville populations are scattered throughout (consistent with the AMOVA results); the Papuan Aita are the most removed group, followed by the Lavongai; none of the Papuanspeaking groups are particularly closely paired, even those considered linguistically closely related (e.g. the Ata and Anêm, and the two Baining languages, Kaket
Table 5.4b
AMOVA Based on Y-chromosomal STRs (Rst) Variance Components (%)
Group
Between Groups
Within Groups
Within Populations
No grouping (23 populations) Geography (3 Islands)a, b Language (2 groups)c
… 9.7 −1.1
12.9 7.1 13.5
87.1 83.2 87.6
New Ireland New Britain Bougainville
… … …
4.3 9.3 5.3
95.7 90.7 94.7
Papuans Oceanians
… …
14.8 12.6
85.2 87.4
Non-significant values are shown in italic. a Groups for geography: New Ireland, New Britain, and Bougainville. b For this AMOVA analysis Lavongai was included with New Ireland. c Linguistic groups: Papuans, Oceanians.
92
and Mali); and geographic proximity, while important overall, does not seem to determine the closest pairwise relationships.
Discussion and Conclusion In concert with our intensive sampling regimen in Northern Island Melanesia, newly identified regionally specific SNPs have greatly enhanced our ability to identify highly structured NRY variation. Earlier surveys concluded that populations there were comparatively low in NRY SNP heterozygosity, but that now appears to have been caused by an ascertainment bias. The considerable NRY structure is consistent with the extremely curtailed marital migration distances (male and female) that have been observed in the region, particularly in the island interiors (Friedlaender, 1975; Friedlaender et al., 2006). To judge by their microsatellite diversity, all of the indigenous lineages are of considerable age, with divergence estimates in the 32,000–50,000 YBP range for those haplogroups that developed in this region. These estimates are compatible with those from studies of mtDNA and X variation in the region (Friedlaender et al., 2005; Merriwether et al., 2005; also see chapter 13), and are accommodated by the established earliest settlement dates for the region. Even the earliest archeological dates for Buka/Bougainville, which are the youngest for these islands (29,000 YBP), falls very close to the estimated date for the origin of M2 and M2a. Almost all populations covered had considerable internal SNP and STR diversity. However, there remained a large among-group variance, which had both significant
Y Chromosome Variation in Northern Island Melanesia
Figure 5.5 Two-dimensional multidimensional scaling plot, generated from pairwise Φst values for Northern Island Melanesian populations. See color insert.
within-island and among-island components. The size of the islands did not predict their within-island variation, and language distinctions did not account for a significant partitioning of the among-group variance. It appears that each of the haplogroups that is indigenous to this region has its highest concentration in a Papuan-speaking population, but has diffused from that center across neighboring languages so that any original specific haplogroup–language association has been diluted. We can still identify likely expansion centers for certain regional variants from their distributions and microsatellite diversities. New Guinea is the likely origin for C2b, K5, and M. Different populations of New Britain were the origins of both K6 and K7, and also M2a. While K6 has also dispersed to New Ireland and Mussau, microsatellite diversity helps to locate the expansion of K6 more specifically to New Britain. M2 likely developed in New Ireland, where the Papuan-speaking Kuot appear to be the epicenter. The Aita of north Bougainville are at the current focus for M2a, the most common haplogroup in our survey. It is intriguing that the Lavongai populations just to the north of New Ireland also have very high M2a frequencies.
We identify haplogroup O3 most clearly with the Southeast Asian component in the development of the Lapita Peoples, and O3 has a remarkably low frequency in this sample series (18 of 685 samples, or 2.6%). Even if all O haplogroups might have been introduced just with the immediate Southeast Asian ancestral component of the Lapita peoples, their frequency in our series totals just 7% (table 5.2), almost entirely restricted to Oceanicspeaking populations. We may have missed higher concentrations of the O haplogroups in our sampling strategy. For example, many Lapita sites in the Bismarck Archipelago are located on smaller or off-shore islands, and we did not sample those, except for Mussau (which has important Lapita sites, but where no O samples were found). It is also possible that most of the “Southeast Asian” male component either moved on to other islands or was lost/obliterated in this region, especially when considering the effects of genetic drift. The clear deduction from the NRY data is that the Southeast Asian male contribution to contemporary Oceanic-speaking populations in Northern Island Melanesia is unexpectedly small. This contrasts with the accepted interpretation of the mtDNA haplogroup frequency data in this region, because of the
93
core studies in northern island melanesia
extremely high frequency of the so-called Polynesian Motif (B4a1a1) there, even in many Papuan-speaking groups (Merriwether et al., 1999). Our suggested resolution of these conflicting interpretations is forthcoming (Friedlaender et al., 2006). If they are incorporated into screens in Remote Oceania (i.e. Southern Island Melanesia, Polynesia, and Micronesia), the newly described NRY haplogroup markers should help in identifying ancestral male contributions from particular sections of Northern Island Melanesia and New Guinea. We know that haplogroup M occurs in Vanuatu in high frequency, as well as Fiji and Tonga (table 5.1). Haplogroup K is common in Vanuatu and Fiji, as well as occurring in low frequencies in all Polynesian series to date. Therefore, identifying K6, K7, M2, and M2a in these series will make a major difference in understanding their prehistory. As a final conclusion, the NRY variation we have identified belies the generally accepted notion of clinal loss of variation from Southeast Asia through New Guinea, Northern Island Melanesia, and out into Remote Oceania. This distribution of all of the regionally specific haplogroups (C2, C2b, K5, K6, K7, M, M2, and M2a) suggests that this region east of the Wallace Line was an important center of population diversity in the Pacific.
Acknowledgments We would first like to express our gratitude to the participating people of the Bismarck Archipelago and Bougainville. We thank William Beggs and Patrick Bender for their advice, and Charles Mgone, Heather Norton, Daniel Hrdy, and Andrew Merriwether for their participation in sample collection. The laboratory aspect of this research was supported by a Wenner-Gren Foundation pre-doctoral grant, while the fieldwork was supported by post-doctoral grants from the National Science Foundation, the Wenner-Gren Foundation for Anthropological Research, and the National Geographic Society Research Fund.
References Bandelt HJ, Forster P, Rohl A. 1999. Median-joining networks for inferring intraspecific phylogenies. Molecular Biology and Evolution 16: 37–48. Capelli C, Wilson JF, Richards M, Stumpf MP, Gratrix F, Oppenheimer S, Underhill P, Pascali VL, Ko TM, Goldstein DB. 2001. A predominantly indigenous paternal heritage for the Austronesian-speaking peoples of insular Southeast Asia and Oceania. American Journal of Human Genetics 68: 432–43. Cox MP. 2003. Genetic patterning at Austronesian contact zones. In: Anthropology. Dunedin, New Zealand (doctoral dissertation): University of Otago. pp 270.
94
Cox MP, Mirazon Lahr M. 2006. Y-chromosome diversity is inversely associated with language affiliation in paired Austronesian- and Papuan-speaking communities from the Solomon Islands. American Journal of Human Biology. 18:35–50. Ellis N, Hammer M. 2002. Y chromosome consortium: A nomenclature system for the tree of human Y-chromosomal binary haplogroups. Genome Research 12: 339–48. Excoffier L, Schneider S. 2005. Arlequin version 3.0: An integrated software package for Population Genetics Data Analysis. Evolutionary Bioinformatics Online. 1:47-50. Forster P, Kayser M, Meyer E, Roewer L, Pfeiffer H, Benkmann H, Brinkmann B. 1998. Phylogenetic resolution of complex mutational features at Y-STR DYS390 in aboriginal Australians and Papuans. Molecular Biology and Evolution. 15: 1108–14. Friedlaender JS. 1975. Patterns of human variation: The demography, genetics, and phenetics of Bougainville Islanders. Cambridge, MA: Harvard University Press. Friedlaender J, Schurr T, Gentz F, Koki G, Friedlaender F, Horvat G, Babb P, Cerchio S, Kaestle F, Schanfield M, Deka R, Yanagihara R, Merriwether DA. 2005. Expanding Southwest Pacific mitochondrial haplogroups P and Q. Molecular Biology and Evolution 22: 1506–17. Friedlaender JS, Gentz F, Thompson F, Kaestle F, Schurr TG, Koki G, Mgone C, McDonough J, Smith L, Merriwether DA. 2006. Mitochondrial genetic diversity and its determinants in Island Melanesia. In: Pawley A, Attenborough R, Golson J, Hyde R, editors. Papuan pasts: Studies in the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 693–716. Groube L. 1986. Waisted axes of Asia, Melanesia, and Australia. In: Ward GK, editor. Archaeology at ANZAAS. Canberra, Australia: Canberra Archaeological Society. pp 168–77. Hammer MF, Spurdle AB, Karafet T, Bonner MR, Wood ET, Novelletto A, Malaspina P, Mitchell RJ, Horai S, Jenkins T, Zegura SL. 1997. The geographic distribution of human Y chromosome variation. Genetics 145: 787–805. Jobling MA, Tyler-Smith C. 2003. The human Y chromosome: An evolutionary marker comes of age. Nature Reviews Genetics 4: 598–612. Karafet TM, Lansing JS, Redd AJ, Reznikova S, Watkins JC, Surata SP, Arthawiguna WA, Mayer L, Bamshad M, Jorde LB, Hammer MF. 2005. Balinese Y-chromosome perspective on the peopling of Indonesia: Genetic contributions from pre-neolithic hunter-gatherers, Austronesian farmers, and Indian traders. Human Biology 77: 93–114. Kayser M, Brauer S, Weiss G, Underhill PA, Roewer L, Schiefenhovel W, Stoneking M. 2000. Melanesian origin of Polynesian Y chromosomes. Current Biology 10: 1237–46. Kayser M, Brauer S, Weiss G, Schiefenhovel W, Underhill PA, Stoneking M. 2001a. Independent histories of
Y Chromosome Variation in Northern Island Melanesia
human Y chromosomes from Melanesia and Australia. American Journal of Human Genetics 68: 173–90. Kayser M, Krawczak M, Excoffier L, Dieltjes P, Corach D, Pascali V, Gehrig C, Bernini LF, Jespersen J, Bakker E, Roewer L, de Knijff P. 2001b. An extensive analysis of Y-chromosomal microsatellite haplotypes in globally dispersed human populations. American Journal of Human Genetics 68: 990–1018. Kayser M, Brauer S, Weiss G, Schiefenhovel W, Underhill P, Shen P, Oefner P, Tommaseo-Ponzetta M, Stoneking M. 2003. Reduced Y-chromosome, but not mitochondrial DNA, diversity in human populations from West New Guinea. American Journal of Human Genetics 72: 281–302. Leavesley MG, Bird MI, Fifield LK, Hausladen PA, Santos GM, di Tada ML. 2002. Buang Merabak: Early evidence for human occupation in the Bismarck Archipelago, Papua New Guinea. Australian Archaeology 54: 55–7. Merriwether DA, Friedlaender JS, Mediavilla J, Mgone C, Gentz F, Ferrell RE. 1999. Mitochondrial DNA variation is an indicator of Austronesian influence in Island Melanesia. American Journal of Physical Anthropology 110: 243–70. Merriwether DA, Hodgson JA, Friedlaender FR, Allaby R, Cerchio S, Koki G, Friedlaender JS. 2005. Ancient mitochondrial M haplogroups identified in the Southwest Pacific. Proceedings of the National Academy of Sciences USA 102: 13034–9. Reesink G. 2005. Sulka of East New Britain: A mixture of Oceanic and Papuan traits. Oceanic Linguistics 44: 145–93. Scheinfeldt L, Friedlaender F, Friedlaender J, Latham K, Koki G, Karafet T, Hammer M, Lorenz J. 2006. Unexpected NRY chromosome variation in Northern Island Melanesia. Molecular Biology and Evolution. 23:1628–41.
Shi H, Dong YL, Wen B, Xiao CJ, Underhill PA, Shen PD, Chakraborty R, Jin L, Su B. 2005. Y-chromosome evidence of southern origin of the East Asian-specific haplogroup O3–M122. American Journal of Human Genetics 77: 408–19. Trejaut JA, Kivisild T, Loo JH, Lee CL, He CL, Hsu CJ, Li ZY, Lin M. 2005. Traces of archaic mitochondrial lineages persist in Austronesian-speaking formosan populations. Public Library of Sceinces. Biology 3: e247. Underhill PA. 2004. A synopsis of extant Y chromosome diversity in East Asia and Oceania. In: Sagart L, Blench R, Sanchez-Mazas A, editors. The peopling of East Asia: Putting together archaeology, linguistics and genetics. London: Routledge Curzon. pp 301–19. Underhill PA, Passarino G, Lin AA, Marzuki S, Oefner PJ, Cavalli-Sforza LL, Chambers GK. 2001a. Maori origins, Y-chromosome haplotypes and implications for human history in the Pacific. Human Mutation 17: 271–80. Underhill PA, Passarino G, Lin AA, Shen P, Mirazon Lahr M, Foley RA, Oefner PJ, Cavalli-Sforza LL. 2001b. The phylogeography of Y chromosome binary haplotypes and the origins of modern human populations. Annals of Human Genetics 65: 43–62. Whitfield LS, Sulston JE, Goodfellow PN. 1995. Sequence variation of the human Y chromosome. Nature 378: 379–80. Wilder JA, Kingan SB, Mobasher Z, Pilkington MM, Hammer MF. 2004. Global patterns of human mitochondrial DNA and Y-chromosome structure are not influenced by higher migration rates of females versus males. Nature Genetics 36: 1122–5. Y chromosome consortium (Ellis N and Hammer M, organizers). 2002. A nomenclature system for the tree of human Y–chromosomal binary haplogroups. Genome Research 12: 339–48.
95
6 Pigmentation and Candidate Gene Variation in Northern Island Melanesia Heather L. Norton, George Koki, and Jonathan S. Friedlaender
Introduction Skin pigmentation shows remarkable variation across the human species. It generally follows a pattern in which populations living in regions of intense ultraviolet radiation (UVR) tend to have darker skin pigmentation than regions living in lower UVR regions (Walter, 1971; Jablonski and Chaplin, 2000; Chaplin, 2004). This obvious connection between pigmentation and UVR has led to the development of a number of hypotheses that attempt to explain normal variation in human skin pigmentation in terms of differential natural selection (Murray, 1934; Cowles, 1959; Blum, 1961; Wasserman, 1965; Loomis, 1967; Walter, 1971; Pathak and Fitzpatrick, 1974; Post et al., 1975; Branda and Eaton, 1978; Mackintosh, 2001; Chaplin, 2004). Others have suggested that sexual selection may have also shaped skin pigmentation variation (Darwin, 1871; Van den Berghe and Frost, 1986; Frost, 1988; Diamond, 1992). These hypotheses attempt to explain average skin pigmentation differences across broad geographic regions, but the often-considerable variation in pigmentation within smaller geographic regions is frequently overlooked. Most natural-selection-based hypotheses predict that darker skin is an adaptation to life in high UVR environments because melanin, the primary pigment of the skin, can provide some form of protection from UVR-related damage (Pathak and Fitzpatrick, 1974; Kollias et al., 1991; Sheehan et al., 2002). The photo-protection hypothesis argues dark skin provides important protection from
96
UV-induced damage such as sunburn and melanoma in regions of intense UV radiation (Blum, 1961; Walter, 1971; Pathak and Fitzpatrick, 1974; Roberts and Kahlon, 1976). The folic acid hypothesis (Branda and Eaton, 1978) suggests that a highly melanized skin will prevent UVmediated folic acid degradation, which may lead to neural tube birth defects (Bower and Stanley, 1989; MVSR, 1991; Jablonski, 1992; Flemming and Copp, 1998). Alternatively, the vitamin D hypothesis proposes that lighter skin in regions of lower UVR is an adaptation to maximize UV-mediated cutaneous vitamin D synthesis (Murray, 1934; Loomis, 1967). Since Northern Island Melanesians live close to the equator (where UVR is typically quite high), one might well expect all groups there to be heavily pigmented. However, this does not appear to be true from casual observations. In this chapter, we quantitatively assess pigmentation across the region as well as test just how homogeneous populations are with respect to pigmentation and associated candidate gene variation. To accomplish this, we surveyed skin pigmentation in 1135 Northern Island Melanesians who belonged to a variety of populations, language groups, and islands. We also sought to identify gene variants that might underlie pigmentation variation in the region by typing ten single nucleotide polymorphisms (SNPs) from six pigmentation candidate genes in a sub-sample of 647 individuals, and also compared these results with samples from five other regions on different continents (an average of 60 individuals per sample) for comparison.
Pigmentation and Candidate Gene Variation in Northern Island Melanesia
Phenotypic Variation Subject Classification We classified individuals into both linguistic and geographic categories (phylum and island), as well as a third category (neighborhood) that incorporated aspects of both geography and language. Phylum classification (Austronesian vs. Papuan) was based on the linguistic affiliation of the study subject and both of his or her parents. To be placed into the Austronesian (AN) or Papuan (P) category, an individual and both parents had to speak languages belonging to the same linguistic phylum. We followed a similar rule for island and neighborhood classification. We adopted this stringent classification to try to minimize the effects of migration and admixture during the last generation. For further details about classification schemes, see Norton et al. (2006). We excluded individuals who could not be assigned to a particular category.
Phenotypic Measurements We measured skin and hair pigmentation with a very sensitive instrument, the Dermaspectrometer (Cortex Technologies, Hadsund, Denmark), which is a narrowband reflectance spectrophotometer. It is able to differentiate between the light absorbance properties of melanin and hemoglobin, the primary pigments of the skin. As a result, we could estimate the amount of melanin in the skin, discounting the effects of hemoglobin (Diffey et al., 1984). This measurement is the Melanin (M) Index. Dermaspectrometer measurements are not comparable to those made by less-sensitive instruments used in previous skin reflectance studies, which is a limitation. Three measurements were taken on the right and left upper inner arms and averaged together to provide an average skin M Index value for each individual. In the same way, the M Index was also obtained from three measurements of hair at the crown and averaged together.
Variation by Sex Since a number of previous reports showed that males and females may differ in skin pigmentation (Barnicot, 1958; Tobias, 1961; Conway and Baker, 1972; Byard and Lees, 1982; Harvey, 1985; Van den Berghe and Frost, 1986; Frost, 1988; Jablonski and Chaplin, 2000) we wanted to verify that males and females in our sample were also significantly different for skin or hair pigmentation. A standard two-sample t-test demonstrated that males in our sample were significantly darker for skin pigmentation (male M Index = 74.0, female M Index = 71.2, p < 0.0001). Males were also darker in hair pigmentation, although the difference reaches only suggestive significance levels
(male M Index = 155.4, female M Index = 151.2; p < 0.0537). Because of these differences, all subsequent comparisons were carried out using M Index values standardized for sex. Table 6.1 gives mean skin and hair M Index values (in their raw forms) for each island and neighborhood.
Variation by Language Phylum Northern Island Melanesians speak languages conventionally grouped into two linguistic phyla: Austronesian (AN) and Papuan (P), and we will use this dichotomy to test the general importance of historical language affiliation in determining skin pigmentation. However, the dichotomy is not straightforward. As described in chapters 3 and 8, the Austronesian languages in this region are spoken by groups that are, probably to quite a variable degree, the descendants of migrants originating from a Southeast Asian homeland that first began arriving in the region ~3,300 years ago, who were associated with the appearance of the Lapita cultural complex in Northern Island Melanesia. The Papuan languages are far more diverse and do not form an actual language family in the same way that Austronesian languages do (more specifically, the Oceanic subdivision). Some modern Papuanspeaking groups are clearly the descendants of some of the earliest arrivals to the region, while others have intermixed with incoming populations, as shown in chapter 8. Previous work has suggested that genetic distinctions between AN- and P-speaking groups do still exist in some instances, as in New Guinea’s Markham Valley (Giles et al., 1965), but other early studies suggested that the distinctions may not be so clear in other instances in Near Oceania (Serjeantson and Gao, 1995; Merriwether et al., 1999). We thought that if there were a remaining simple distinction between groups speaking the two different language clusters, then the Austronesian-speaking groups would be more lightly pigmented and similar to Southeast Asians than the (likely more variable) Papuanspeaking groups, whose ancestors were some of the original migrants to the region. Although no significant differences between the two groups (pooled across the region) were observed for skin pigmentation, the two groups were significantly different for hair M Index, with Papuan speakers having lighter hair (t = 3.81, p < 0.01). We also compared pigmentation within each island between AN and P speakers. AN and P speakers were significantly different for skin pigmentation in New Ireland (t = 2.93, p < 0.01), meaning the Kuot (the lone Papuan-speaking group) were darker than everyone else there. AN and P neighborhoods differed significantly from each other in hair pigmentation on both Bougainville (t = 6.40, p < 0.0001) and New Ireland (t = 2.39, p < 0.05). The difference in hair pigmentation
97
core studies in northern island melanesia
Table 6.1
Mean Skin and Hair M Index Values for Islandsa and Neighborhoodsb
Island
Neighborhood
PNG
N
Skin Mean M
S.D.
N
Hair Mean M
S.D.
21 11 491 45 25 29 34 22 11 13 17 44 20 96 36 18
67.9 69.4 67.9 65.0 72.2 67.1 69.7 68.6 67.3 67.3 70.3 67.4 67.6 65.7 69.3 68.3
9.4 10.6 6.8 5.3 7.5 7.2 6.6 6.9 5.2 6.5 5.5 5.4 9.7 6.2 6.6 6.3
19 10 476 41 25 25 32 22 11 13 17 44 19 93 36 18
160.1 169.7 155.2 157.5 150.4 147.5 155.3 160.0 157.5 160.6 160.2 150.2 173.3 144.4 165.0 151.9
15.3 11.7 17.7 11 8.6 13.2 15.9 16.0 13.2 16.9 29.7 14.2 18.2 12.2 20.2 20.4
35 34
65.5 65.0
7.9 7.4
33 32
162.5 162.4
10.0 10.1
15. North Lavongai 16. West Lavongai 17. South Lavongai
102 73 10 13
77.1 76.9 75.5 77.3
7.5 6.9 7.6 9.9
98 69 10 13
154.4 154.9 157.3 150.3
16.1 15.9 18.4 14.2
18. Tigak 19. Nailik 20. KABIL (KUOT) 21. LAMALAUA (KUOT) 22. Notsi 23. Madak
242 27 26 41 11 21 26
74.2 72.6 71.9 76.2 75.0 74.8 78.0
8.1 8.4 6.4 7.8 6.2 9.0 8.9
222 26 20 38 9 20 26
151.6 156.4 148.9 147.4 141.3 160.8 148.0
23.1 15.9 29.8 19.1 29.5 18.4 27.5
24. Saposa Island (Saposa) 25. Inivus (Teop) 26. Sunahoara (Teop) 27. KUKUAVO (AITA)
153 41 10 10 32
89.8 86.1 93.1 94.6 91.9
9.5 9.9 9.0 6.5 8.5
127 29 9 9 30
150.7 154.3 164.5 158.5 138.6
14.6 11.4 11.6 12.0 8.9
SEPIK New Britain 1. Arimegi Island (Kove) 2. KARIAI (ANÊM) 3. PURELING (ANÊM) 4. Kisiluvi (Mamusi) 5. Lingite (Mamusi) 6. Welu (Mamusi) 7. Other Mamusi 8. Loso (Nakanai) 9. UASILAU (ATA) 10. LUGEI (ATA) 11. Bileki (Nakanai) 12. Ubili (Melamela) 13. Kuanua (Tolai) Mussau 14. Kapugu New Hanover
New Ireland
Bougainville
a
Includes individuals belonging to the neighborhoods listed below each island heading as well as individuals who could not be assigned to a particular neighborhood within the island. b Neighborhoods shown in capital letters are Papuan-speaking.
between the two language groups on Bougainville is due to the very light-haired Aita of Kukuavo. While differences between AN and P speakers on New Britain are not significant, there is a suggestion of a trend towards more darkly pigmented people inland (primarily P-speaking groups) and more lightly pigmented people (primarily AN-speaking groups) on the coast. As there is some evidence for the northern coast of New Britain being a potential Lapita homeland (chapter 8), these differences in pigmentation between coastal and inland groups may reflect original patterns of pigmentation differences between AN and P speakers in the region. The absence of significant 98
pigmentation differences between AN and P speakers across the region suggests two reasonable alternative scenarios: admixture between the two groups has erased any pre-existing differences; or both groups have inhabited the region long enough to adapt to the intense Near Oceanic UVR levels. While we were interested in seeing if simple differences in pigmentation existed between the two linguistic phyla, we also wanted to test their homogeneity for skin and hair pigmentation. We used the Analysis of Variance approach (ANOVA) to compare the pigmentation of neighborhoods within each language phylum on
Pigmentation and Candidate Gene Variation in Northern Island Melanesia
an island-by-island basis (table 6.2). As the table shows, there is generally a great deal of variation, no matter how the comparisons are structured. There is significant variation for both hair and skin pigmentation among New Britain’s Austronesian- (p < 0.0001) and Papuan-speaking (p < 0.01) neighborhoods. Significant skin pigmentation variation also exists among Austronesian-speaking neighborhoods on Bougainville (p < 0.05). These results suggest that heterogeneity exists within both AN- and P-speaking groups in the large islands of New Britain and Bougainville, but not for the smaller (and much narrower) island of New Ireland. This shows that both island size and topographic complexity impact genetic diversity, an inference that is consistent with mtDNA and NRY findings in this same region (see chapters 4, 5, and 7).
Island and Neighborhood Variation Figures 6.1 and 6.2 give the mean skin and hair M Index values for each neighborhood and island (the Sepik of PNG are excluded from the figures for clarity). It was something of a surprise to see that the islands are significantly different from each other for hair pigmentation (F = 3.88, df = 5, p < 0.01). The skin pigmentation differences among islands are much more obvious, as we expected (F = 222.23, df = 5, p < 0.0001), with New Guinea, Mussau, and New Britain people at the lighter end of the spectrum,
Table 6.2 ANOVA Results Comparing Islands, Neighborhoods, Neighborhoods within Phyla and Neighborhoods within Islands Skin Model Island Neighborhoods (entire sample) Neighborhoods (within phyla) AN P Neighborhoods (within islands) New Britain New Hanover New Ireland Bougainville Neighborhoods (within phyla within islands) New Britain AN New Britain P New Hanover AN New Ireland AN New Ireland P Bougainville AN New Hanover *p < 0.05; **p < 0.01; ***p < 0.0001.
Hair
F
df
F
df
222.23*** 34.96***
5 27
3.88** 6.06***
5 27
37.30*** 35.50***
18 8
4.40*** 8.44***
8 8
3.92*** 0.19 2.53 5.25**
12 2 5 3
8.74*** 1.00 1.79 18.81***
12 2 5 3
4.28*** 4.06** 0.19 1.53 1.18 4.87* 0.19
8 3 2 2 2 2 2
7.64*** 16.12*** 1.00 1.70 0.40 2.46 1.00
8 3 2 2 2 2 2
New Hanover and New Ireland people intermediate, and Bougainvilleans clearly being the darkest—a clear westto-east gradient. Thinking that the very darkly pigmented Bougainvilleans might be causing these results by themselves, we re-ran the analysis excluding Bougainvilleans; the results are still highly significant (F = 68.47, df = 4, p < 0.0001). Table 6.2 and figures 6.1 and 6.2 also show there is also significant heterogeneity among neighborhoods within each island. For New Hanover, New Britain, New Ireland, and Bougainville (the four islands that had multiple neighborhoods) we observed significant variation in hair pigmentation among neighborhoods on the islands of Bougainville (F = 18.81, df = 3, p < 0.0001) and New Britain (F = 8.74, df = 12, p < 0.0001). Significant variation in skin pigmentation was present on New Ireland (F = 2.53, df = 5, p < 0.05), New Britain (F = 3.92, df = 12, p < 0.0001), and Bougainville (F = 5.25, df = 3, p < 0.01). Language phylum differences cannot account for this heterogeneity (see above).
Correlation with Latitude or UVR We tested for correlations between skin M Index and latitude (using Global Gazetteer Version 2.1—http://www. fallingrain.com/world/PP/), since the natural-selectionbased hypotheses predict skin pigmentation should increase towards the equator (where UVR tends to be higher). There was no significant correlation of skin M Index with latitude (R2 = 0.0000, p < 0.9805). This was not surprising since our sampling covered only 9.3° of latitude. Because latitude is only a surrogate for UVR levels, and since UVR has been shown to be a better predictor of skin pigmentation than latitude (Jablonski and Chaplin 2000), we also examined NASA’s reported levels of UVMED (a measure of the minimum amount of UV exposure required to produce a reddening in the skin of a lightly pigmented individual) across Island Melanesia over seven years. Table 6.3 gives UVMED levels, and they do not correspond with the variation in skin pigmentation. In particular, UVMED levels over Bougainville are more comparable to those observed over the much lighter islands of New Britain and PNG. This suggests that other forces, most likely the same historical demographic ones determining the mtDNA and NRY variation across the region, are responsible for the significant localized variation in pigmentation variation. Although differential natural selection cannot explain the extensive pigmentation variation within Near Oceania, we do believe that natural selection has influenced the overall pigmentation levels there, since first colonization. This is simply because they are relatively dark when compared to other groups, such as European and even African Americans (figure 6.3). We suggest a melaninselective threshold model may have been in effect, where 99
core studies in northern island melanesia
Figure 6.1 Hair M index for 27 of the 28 neighborhoods reported in this study (the Sepik of New Guinea are not included). Intensity of the circle marking each neighborhood on the map corresponds to mean pigmentation as measured by the M index. Neighborhoods are listed as follows: 1Kapugu, 2Kove, 3Kariai (Anˆem), 4 Pureling (Anˆem), 5Bileki (Nakanai), 6Uasilau (Ata), 7Lugei (Ata), 8Loso (Nakanai), 9 Kisiluvi (Mamusi), 10Lingite (Mamusi), 11Welu (Mamusi), 12”other Mamusi,” 13Ubili (Melamela), 14Kuanua, 15North Lavongai, 16West Lavongai, 17South Lavongai, 18Tigak, 19 Nailik, 20Kabil (Kuot), 21Lamalaua (Kuot), 22Notsi, 23Madak, 24Saposa Island (Saposa), 25 Invivus (Teop), 26Sunahoara (Teop), 27Kukuavo (Aita).
a certain level of pigmentation that is “dark enough” is maintained to provide adequate protection from UVR. Above this threshold, pigmentation would be free to vary due to historical factors or even sexual selection. Chaplin (2004) proposed a similar idea, although his concept may be better described as a melanin maximum. Under this melanin maximum model there could be a point past which it becomes ever more difficult to increase pigmentation production—for example, Bougainvilleans might represent a group that may have achieved this theoretical maximum.
candidate gene frequencies as well. If the same genes cause dark skin in Near Oceanic and West African populations, then functional single nucleotide polymorphism (SNP) allele frequencies at pigmentation candidate genes in these two groups should not be different, while there should be significant frequency differences between populations in Near Oceania and lighter-skinned Europeans and East Asians. We therefore analyzed SNPs at six candidate genes to test their variation and association with pigmentation in the sample. First, the potential role of these genes in regulating pigmentation variation will be reviewed.
Pigmentation Candidate Genes
Melanin Biosynthesis and Pigmentation Candidate Genes
Because of the great variation in pigmentation across the sample, we expected to see comparable variation in associated
Melanin is the primary pigment of the skin (and hair), although other molecules such as hemoglobin also affect
100
Pigmentation and Candidate Gene Variation in Northern Island Melanesia
Figure 6.2 Skin M index for 27 of the 28 populations reported in this study (the Sepik of New Guinea are not included). Intensity of the circle marking each neighborhood on the map corresponds to mean pigmentation as measured by the M index. Numbers refer to neighborhoods as specified in the legend for figure 6.1.
skin coloration and appearance. Melanin production takes place inside specialized cells found in the basal layer of the skin known as melanocytes. There are two forms of melanin in mammalian skin and hair: brown/ black eumelanin, and red/yellow pheomelanin. The initial Table 6.3
Mean Skin M Index and UVMED by Island
Island PNG New Britain Mussau New Hanover New Ireland Bougainville
Skin M Index 67.9 67.9 65.5 77.1 74.2 89.8
UVMED 275 274 292 289 288 274
steps in the production of both eumelanin and pheomelanin require the hydroxylation of tyrosine to 3,4-dihydroxyphenylalanine (DOPA) and the oxidation of DOPA to DOPAquinone by tyrosinase. At this point, the melanin synthesis pathway diverges: pheomelanin is produced from the metabolites of 5-S-cysteinylDOPA, and eumelanin is produced from the metabolites of DOPAchrome. After it is synthesized in the melanocytes, melanin is packaged into membrane-bound organelles known as melanosomes. Eumelanin is stored in larger ellipsoidal melanosomes with a highly complex glycoprotein matrix and pheomelanin is stored in smaller, spherical melanosomes with a less-complex glycoprotein matrix (Sturm et al., 2001). Nordlund et al. reported that individuals of
101
core studies in northern island melanesia
Figure 6.3 Distributions of skin M index values for the Island Melanesians sampled in this stidy and comparison populations of African Americans and African Caribbeans measured by Shriver et al. (2003).
“diverse ethnic background” demonstrate quantifiable differences in both melanosome distribution as well as the degree of melanization (Nordlund et al., 1998). Darker skin has greater numbers of larger melanosomes, while lighter skin has sparsely distributed aggregates of smaller melanosomes. Following maturation, melanosomes are transported out of the melanocyte cell and into keratinocytes. The exact transfer mechanism is unknown, but two primary candidates are phagocytosis or endocytosis (Westbroek et al., 2001). Once they have been transmitted to the keratinocytes the melanosome membranes will degrade, allowing the melanin to diffuse throughout the cell. In many cases the melanin may form a protective “cap” over the nucleus of the keratinocyte, providing protection from UVR (Montagna and Carlisle, 1991).
102
Exposure to UVR can stimulate melanogenesis as part of the tanning response (Friedmann and Gilchrest, 1987), most likely due to the resulting DNA damage or the DNA repair intermediates that are produced following UVR exposure (Agar and Young, 2005). Some studies have reported that in the presence of pheomelanin and its precursor, 5-S-cysteinyldopa, UVR is highly mutagenic (Harsanyi et al., 1980; Koch and Chedekel, 1986). This suggests that the type of melanin produced may have an effect on the risk of UVR-induced damage such as melanoma (Schmitz et al., 1995; Sturm, 1998). Because of this entire process, genes involved in either the production of melanin itself, the construction of the melanosomes, the switch between the production of eumelanin and pheomelanin, or the transport of melanosomes to the keratinocyte are all good candidates
Pigmentation and Candidate Gene Variation in Northern Island Melanesia
for affecting normal variation in skin and hair pigmentation. Many of the candidate genes were originally identified for the role that they play in pigmentary disorders such as albinism or through studies of mouse coat color genetics (Bennett and Lamoreux, 2003). The six candidate genes in this study are TYR, TYRP1, OCA2, MATP, ASIP, and MC1R. TYR, TYRP1, OCA2, and MATP are associated with various forms of albinism (Giebel et al., 1991; Inagaki et al., 2004; King et al., 1991; Lee et al., 1994; Manga et al., 1997; Spritz et al., 1995; Stevens et al., 1995; Suzuki et al., 2003; Tomita et al., 1989; Yi et al., 2003) and there is reason to believe that they may have more subtle effects on normal phenotypic variation. For example, admixture mapping techniques have been used to demonstrate this for TYR and OCA2 among African Americans and African Caribbeans (Shriver et al., 2003). MC1R is well-known for its association with red hair and fair skin among European populations (Box et al., 1997; Flanagan et al., 2000; Smith et al., 1998; Valverde et al., 1995). Relative to African populations, it also shows high levels of polymorphism among East Asians (Harding et al., 2000; Peng et al., 2001; Rana et al., 1999; Yao et al., 2000). MC1R’s antagonist, ASIP, is associated with darker pigmentation among Europeans (Kanetsky et al., 2002) and with darker skin among African Americans (Bonilla et al., 2005). Three SNPs from TYR were included in this study: A192C, C308G, and A402G. Of these, only A192C had been found to be associated with normal skin pigmentation variation, even after controlling for differences in individual admixture proportions in a combined sample of African-Caribbeans and African-Americans (Shriver et al., 2003). However, given the important role that TYR plays in melanin production, we felt that the two additional non-synonymous SNPs might also show associations with pigmentation variation. The SNP typed in TYRP1, A209T, is a synonymous substitution but was selected because it may be located or linked to a functional SNP in the gene and so might show an association. The OCA2 SNP used in the Shriver et al. paper, A355G, was typed in this study even though it does not result in an amino acid change. We typed two MC1R polymorphisms, neither of which shows strong association with red hair color (RHC). Although it is classified as a “weak” RHC allele (Duffy et al., 2004) G92A also occurs at variable frequencies (10–32%) in East and South Asian samples (Rana et al., 1999; Yao et al., 2000; Peng et al., 2001). Since Near Oceanic Austronesian speakers originated at least in part from a Southeast Asian homeland, we felt that G92A might show appreciable variation in our sample. The second MC1R SNP we typed was a synonymous substitution, G314A. Its allele frequency differs significantly among Africans, Native Americans, and Europeans, and
consequently has been used as an ancestry informative allele (AIM) by Shriver et al. (2003) and others. It is also polymorphic in East Asians (Yao et al., 2000; Peng et al., 2001). As no coding polymorphisms associated with pigmentation in ASIP have been discovered to date, we typed the promoter variant described in Kanetsky et al. (2002) and Bonilla et al. (2005) in our samples. The sixth pigmentation candidate gene we typed, the membrane associated transport protein (MATP, formerly known as AIM1), encodes a melanoctye differentiation antigen. MATP mutations have been implicated in OCA4 (Newton et al., 2001; Inagaki et al., 2004; Rundshagen et al., 2004). In 2002, Nakayama et al. described two polymorphisms (A272G and C374G) that had clear allele frequency differences between populations known to differ in skin pigmentation phenotype. As this work was being completed, Graf et al. (2005) reported significant associations between two MATP SNPs (A272G and C374G) and skin, hair, and eye pigmentation in Europeans. We typed both of these SNPs.
Phenotypic Characteristics of the Genotyped Sample A total of 647 individuals, primarily from the islands of New Britain, New Ireland, and Bougainville were genotyped for the ten pigmentation candidate SNPs at the six genes. To ensure that this sub-sample captured a similar range of variation as the original larger phenotype sample set, we re-ran the same analyses of phenotypic variation within and between islands and phyla just on this sub-sample. The statistical findings on the subset were in accord with those from the larger sample except that significant variation was observed in hair M Index between AN and P neighborhoods on New Britain (F = 5.00, df =1, p < 0.05) and that no significant variation was observed in skin pigmentation between AN and P neighborhoods on New Ireland.
Allele Frequencies Table 6.4 gives the allele frequencies for each of the ten SNPs typed in this sub-sample by each neighborhood, island, and phylum. Allele frequencies in five populations of non-Melanesians are included for comparison. Six of the SNPs (TYR A192C, TYR C308G, TYR A402G, TYRP1 A209T, MATP A272G, and MATP C374G) were either monomorphic or showed very low levels of polymorphism (heterozygosity < 0.05) among Near Oceanic populations. Two of these, TYR A192C and MATP C374G, showed sharp allele frequency differences between Europeans and all other groups. The allele frequencies of the four polymorphic SNPs are plotted separately by island and for the outside comparative
103
core studies in northern island melanesia
Table 6.4
Allele Frequencies of SNPsa in Each Island, Neighborhoodb, Language Phylum, and Comparison Population
Island
Neighborhood
New Britain
TYR 192*A
ASIP 8818*A
OCA2 355*A
MATP 374*C
MC1R 92*G
MC1R 314*G
Arimegi Island (Kove) KARIAI (ANEˆ M) PURELING (ANEˆ M) Kisiluvi (Mamusi) Lingite (Mamusi) Welu (Mamusi) Loso (Nakanai) UASILAU (ATA) LUGEI (ATA) Bileki (Nakanai) Ubili (Melamela) Kuanua (Tolai)
288 45 24 27 24 12 10 11 39 12 19 36 15
0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.01 0.00 0.00 0.00 0.04
0.86 0.85 0.69 0.68 0.87 0.83 1.00 0.95 0.97 1.00 0.97 0.84 0.83
0.64 0.48 0.59 0.45 0.78 0.96 0.95 0.55 0.72 0.83 0.56 0.70 0.40
0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
0.86 0.63 0.93 0.92 0.97 1.00 1.00 0.95 0.91 0.92 0.83 0.82 0.82
0.31 0.43 0.28 0.17 0.22 0.09 0.11 0.30 0.32 0.14 0.41 0.36 0.37
Tigak Nailik KABIL (KUOT) LAMALAUA (KUOT) Notsi Madak
158 25 21 33 10 17 26
0.00 0.00 0.00 0.00 0.00 0.00 0.00
0.83 0.89 0.82 0.77 0.90 0.81 0.88
0.45 0.70 0.53 0.27 0.43 0.41 0.50
0.01 0.02 0.05 0.00 0.00 0.00 0.00
0.79 0.63 0.75 0.83 1.00 0.89 0.91
0.39 0.50 0.44 0.33 0.17 0.22 0.43
135 36 10 10 32 386 219 61 63 65 44 66
0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.04 0.01 0.52 0.01
0.48 0.50 0.43 0.69 0.32 0.79 0.76 0.72 0.77 0.98 0.86 0.15
0.25 0.22 0.63 0.22 0.14 0.52 0.46 0.37 0.29 0.38 0.44 0.04
0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.02 0.07 0.06 0.86 0.05
0.82 0.79 0.79 0.94 0.90 0.80 0.90 0.72 0.97 0.95 0.85 1.00
0.34 0.34 0.43 0.35 0.24 0.37 0.28 0.09 0.19 0.11 0.10 0.47
New Ireland
Bougainville Saposa Island (Saposa) Inivus (Teop) Sunahoara (Teop) KUKUAVO (AITA) Austronesian PAPUAN East Asian South Asian Native American European West African
N
Only those SNPs with an allele frequency > 0.05 in at least one group are shown. Papuan-speaking groups are shown in capital letters.
a
b
populations in figure 6.4. When pooled together, Northern Island Melanesians are significantly different (p < 0.05) from West Africans and Native Americans at all four loci, from South Asians at OCA2 A355G, MC1R G92A, and MC1R G314, from East Asians at ASIP A8818G OCA2 A355G and MC1R G92A, and from Europeans at both MC1R loci. Within Northern Island Melanesia, Bougainville is significantly different from both New Britain and New Ireland at ASIP A8818G and OCA2 A355G (p < 0.05). Although Bougainvilleans are the closest to West Africans at these loci, the two groups are still significantly different (p < 0.05). The three islands are significantly different from each other at OCA2 A355G (p < 0.05), and New Ireland and New Britain show
104
significant differences in allele frequencies at both MC1R loci (p < 0.05). The frequencies of the OCA2 A allele among New Britain islanders and among Europeans are not significantly different, which is suggestive, since skin pigmentation in our sample was lightest on New Britain. Figures 6.5 and 6.6 plot the allele frequencies of OCA2 A355G and ASIP A8818G across the different neighborhoods and islands in this study. For OCA2 the A allele (the “light” allele) occurs at its highest frequencies in New Britain, at intermediate frequencies in New Ireland, and at its lowest frequencies on Bougainville. For ASIP, the G allele (the “dark” allele) occurs at elevated frequencies only on Bougainville.
Pigmentation and Candidate Gene Variation in Northern Island Melanesia
Figure 6.4 Allele frequencies for the four SNPs found to be polymorphic in our Island Melanesian sample for each island and in the five comparison populations.
Six of the 10 SNPs that we typed were essentially invariant in our sample. Two of those SNPs that showed little to no variation were TYR A192C and MATP C374G alleles, which are only common in European populations. TYR A192C is associated with lighter skin pigmentation (Shriver et al., 2003), while an association between the derived variant at MATP C374G has recently been demonstrated (Norton, 2005). The remaining four SNPs with little to no variation in our series were included because the literature suggested they were polymorphic. However, these loci were also essentially invariant in the comparison populations. This shows the effects of ascertainment bias in SNP discovery that are usually carried out in populations of European origin (Akey et al., 2003). While these invariant genes might have shared some effect on overall pigmentation levels in Northern Island Melanesia, they cannot be determining the extensive pigmentation variation in Northern Island Melanesia. One method to overcome this ascertainment bias would be to sequence a small Northern Island Melanesian sample for this same set of genes (or selected coding regions within each gene) to identify any polymorphisms that might be common in Northern Island Melanesia.
Hierarchical Locus-specific F Statistics We compared variation within and between neighborhoods and islands by calculating hierarchical F-statistics for the polymorphic pigmentation SNPs in two different ways. In the first, variation in neighborhoods was nested within islands (FNI), islands within the total region (FIT), and neighborhoods within the total region (FNT). If FNI values were consistently greater than those for FIT, it would indicate within-island heterogeneity in pigmentation alleles was more important than between-island heterogeneity. The second method considered language phylum relationships. The samples were nested by variation between neighborhoods within phyla on a single island (FNP), variation between phyla on a single island (FPI), and neighborhoods within the island as a whole (FNI), for each of the three major islands. If FPI was greater than FNI at a pigmentation locus, it suggests the most important within-island population structuring occurred at the language phylum level rather than at the neighborhood level. Table 6.5 gives these FST values. Variation between islands at the ASIP A8818G locus was greater than the variation found within islands. The sharp differences between Bougainville and the other
105
core studies in northern island melanesia
Figure 6.5 Distribution of allele frequencies for the OCA2 A355G SNP. Darker shading is associated with the frequency of the G, or “dark allele.” Numbers refer to neighborhoods as specified in the legend for figure 6.1.
islands are the primary contributing components. At OCA2 A355G, the difference in within- vs. between-island variation was small. At both MC1R loci, variation within islands was greater than variation observed between islands, although in both cases the locus-specific F-statistics were low, indicating relatively little importance. Comparisons of variation among neighborhoods within language phyla on each island show that on New Britain only OCA2 A355G has higher variation between phyla than within. On Bougainville, it was the reverse: variation between language phyla was greater at all loci except for OCA2 A355G. New Ireland showed relatively low levels of variation across all loci, both within and between phyla. Hard conclusions on the meaning of these hierarchical locus-specific F-statistics will depend on comparisons with the pattern(s) of average genetic variation at neutral loci. For example, is the FIT value at ASIP A8818G of 0.13 particularly elevated (indicating strong inter-island
106
divergence at this locus), or is it just consistent with the general pattern from neutral variants? Chapter 9 covers some of these issues.
Genotype/Phenotype Associations We tested for single-locus effects on skin and hair pigmentation using standard analysis of variance (ANOVA) (see table 6.6). For the pooled sample, there were significant associations between the ASIP A8818G and OCA2 A355G genotypes and skin pigmentation, and between the MC1R G314A genotype and hair pigmentation. Table 6.7 gives mean skin and hair M Index values across the three genotype classes at these loci for the entire sample. For both ASIP A8818G and OCA2 A355G, skin pigmentation increases with the number of copies of the G allele, which is the allele that is most common in West Africans. In the same way, the G allele at MC1R G314A may increase
Pigmentation and Candidate Gene Variation in Northern Island Melanesia
Figure 6.6 Distribution of allele frequencies for the ASIP A8818G SNP. Darker shading is associated with the frequency of the G, or “dark allele.” Numbers refer to neighborhoods as specified in the legend for figure 6.1.
hair pigmentation. Although mean hair M Index is slightly darker for heterozygotes than for GG homozygotes, this may be due to the much smaller sample size of the GG homozygotes. However, all of these apparently clear-cut results are complicated by the pervasive population stratification across the region. Are the apparently significant genotype–phenotype associations simply the chance result of there being so much variation among
Table 6.5
islands? To correct for this stratification bias, we tested for genotype–phenotype associations one island at a time. The resulting mean M Index values at loci showing significant associations within an island are in table 6.8. After this correction, we only found significant associations between OCA2 A355G and skin pigmentation within New Britain (F = 3.03, df = 2, p < 0.0498), and MC1R G314A on hair pigmentation within New Ireland (F = 5.35,
Hierarchical FST Statistics for the Region at Large and within Each Island Total Region
ASIP A8818G OCA2 A355G MC1R G92A MC1R G314A
New Britain
New Ireland
Bougainville
FNI
FIT
FNT
FNP
FPI
FNI
FNP
FPI
FNI
FNP
FPI
FNI
0.04 0.10 0.06 0.02
0.13 0.08 0.01 0.01
0.17 0.17 0.07 0.03
0.15 0.04 0.08 0.03
0.03 0.09 0.00 0.00
0.18 0.12 0.08 0.04
0.00 0.02 0.04 0.03
0.00 0.02 0.05 0.01
0.00 0.04 0.08 0.03
0.00 0.10 0.00 0.00
0.04 0.04 0.01 0.01
0.03 0.13 0.00 0.00
107
core studies in northern island melanesia
Table 6.6 Genotype–Phenotype Associations within the Pooled Sample and on Each Island Skin Source
df
Total region ASIP A8818G OCA2 A355G MC1R G92A MC1R G314A Bougainville ASIP A8818G OCA2 A355G MC1R G92A MC1R G314A New Britain ASIP A8818G OCA2 A355G MC1R G92A MC1R G314A New Ireland ASIP A8818G OCA2 A355G MC1R G92A MC1R G314A
2 2 2 2
Table 6.8 Mean Skin and Hair M Index Values for Each Island Where a Significant Genotype–Phenotype Association Was Observed
Hair F 37.16*** 40.97*** 0.56 0.62
df
F
2 2 2 2
2.99 2.03 2.38 3.35*
2 2 2 2
0.57 2.90 0.14 0.05
2 2 2 2
1.51 0.92 1.13 0.42
2 2 2 2
1.76 3.03* 1.83 0.59
2 2 2 2
0.05 0.04 0.49 0.83
2 2 2 2
1.22 1.76 0.42 0.09
2 2 2 2
0.94 0.84 1.59 5.35**
*p < 0.05; **p < 0.01; *** p < 0.0001.
df = 2, p < 0.0059). The effect of OCA2 on skin pigmentation on Bougainville is suggestive, but not significant (F = 2.9, df = 2, p < 0.0590). Of course, by testing for effects within each island our sample size was sharply reduced, decreasing its power. As there remains heterogeneity within island samples, it is also possible that even these significant associations could be caused by within-island stratification.
Table 6.7 Mean Skin and Hair M Index Values for Each of the Three Genotype Classes at the Four Polymorphic Loci in Island Melanesia Genotype ASIP 8818*AA ASIP 8818*AG ASIP 88A8*GG OCA2 355*AA OCA2 355*AG OCA2 355*GG MC1R 92*GG MC1R 92*AG MC1R 92*AA MC1R 314*GG MC1R 314*AG MC1R 314*AA
108
N
Skin M
Hair M
368 158 52 175 237 186 390 118 32 85 217 267
71.3 76.7 83.0 69.4 73.4 79.3 73.7 74.4 71.6 73.3 74.6 73.8
154.6 154.5 147.5 154.9 155.5 152.0 153.9 155.7 160.9 155.3 156.4 152.1
Island
Genotype
New Britain
OCA2 355*AA OCA2 355*AG OCA2 355*GG MC1R 314*GG MC1R 314*AG MC1R 314*AA
New Ireland
N
Skin M
Hair M
118 111 43 25 46 55
66.7 68.7 68.0 -
155.7 158.3 145.2
Summary and Conclusions The remarkable pigmentation variation in Northern Island Melanesia is structured by both inter-island and within-island distinctions, which also relate to distinctions among neighborhoods and language phyla. Since these differences are not associated with UVR variation across the region, differential natural selection is not a cause. Rather, it is the complex population history of the region, involving a number of different migrations and founder effects that shaped this variation. Natural selection may have constrained overall variation, keeping pigmentation levels above a protective threshold, but that was untestable. We were partially successful in linking this extensive pigmentation variation to variation in pigmentation candidate SNPs. Four of the ten pigmentation candidate SNPs that we typed showed significant variation across the region, but our attempt to identify strong genotype– phenotype correlations is open to question because of population stratification issues. Our findings may have been limited by our choice of SNPs in our test battery. Allele frequency information available in the literature is biased towards studies of European, African, Native American, and East Asian populations. While the SNPs that we selected show variation in other regions of the world, many are apparently notably less polymorphic in Island Melanesia. Direct sequencing of a sub-sample of Northern Island Melanesians might identify SNPs that are polymorphic and do have an effect on pigmentation phenotype in Island Melanesia. Also, we only sampled ten SNPs—there are many more that have been recently identified or genotyped in public databases that may show different patterns of variation (e.g. HapMap at www.hapmap.org). While we felt that many of the SNPs might be associated with normal variation in pigmentation in Island Melanesia, some of the SNPs we chose might be primarily responsible for only major effects on pigmentation (as in albinism), and so might not be appropriate to explain normal variation observed among different Island Melanesian populations. Two loci that are not polymorphic
Pigmentation and Candidate Gene Variation in Northern Island Melanesia
among Island Melanesians but that may still be having a strong effect on pigmentation in certain other populations are TYR A192C and MATP C374G. The TYR 192*A allele and the 374*G variant in MATP are only found at very high frequencies in Europeans. These data, as well as results of an admixture mapping test suggest that TYR and MATP may contribute to overall lighter pigmentation in European and European-derived populations. Island Melanesians are at or near fixation for the non-European allele. Two SNPs that have shown admixture mapping signals affecting normal pigmentation variation between Europeans and West Africans and that also show variation in our series of Island Melanesians are ASIP A8818G and OCA2 A355G. There is some evidence for an association with normal phenotypic variation for OCA2 on New Britain, although the strength of this association is questionable because of the underlying genetic heterogeneity in the region. In spite of these problems, the similarities between Bougainvilleans and West Africans at OCA2 A355G and ASIP A8818G are strongly suggestive in light of their similar pigmentation levels, and also because of recent evidence suggesting that Bougainvilleans bear some genetic affinity to Africans and Australian Aborigines (Culotta, 2005). Such relationships are consistent with the idea of an early southern migration out of Africa ~ 65,000 years ago. This also suggests that darkly pigmented skin is actually quite old, as suggested by Rogers et al. (2004). However, while Bougainvillean allele frequencies approximate West African ones, the allele frequencies of the two groups are still significantly different. This highlights the complex nature of pigmentation and the likely role of multiple yet unidentified genes affecting pigmentation phenotype. It may be that while both OCA2 and ASIP contribute to the darker skin of Bougainvilleans and West Africans in similar ways, other genetic loci might contribute to the phenotype differently in each population. For example, when a small sample (n = 21) of Nasioi (a Papuan-speaking population from southern Bougainville) was typed for the ASIP A8818G SNP the results were quite different from those observed in the larger Island Melanesian sample here. Specifically, the Nasioi in this sample had the non-African, or “light” allele at a frequency of 92%. While some Northern Island Melanesians (descendants of the earliest migrants?) may share the same pigmentation polymorphisms with Africans, it could be that their dark pigmentation levels are caused by different genetic mutations (perhaps common in some South Asian populations). Our knowledge of settlement and some migration patterns into and through Island Melanesia has been greatly improved by advances in archeology (reviewed in chapter 2), linguistics (reviewed in chapters 3 and 8), and
molecular anthropology (reviewed in chapter 4). However, we know less about more localized migration and mating patterns within the region and the effects that these have had on population stratification over the millennia. Interisland distances and geographic features within islands obviously presented barriers to completely random mating in the region, although archeological evidence does indicate some trade between regions as early as 20,000 years BP (Marshall and Allen, 1991; Summerhayes and Allen, 1993; Allen, 1996; Leavesley and Allen, 1998; chapter 2). This partial isolation could have led to the divergence of populations on different islands at both pigmentation genes and other loci. Studies of recent (within the last 50 years) marital migration distances suggest that gene flow among populations, particularly inland groups, has been limited (Friedlaender, 1975, and this volume). Such restricted levels of gene flow would also increase intra-island heterogeneity.
Acknowledgments Thanks to Ken Weiss and Mark Shriver for helpful comments and discussion during the course of this work and to the people of Papua New Guinea who so enthusiastically participated in this research. This work was funded in part by Dissertation Improvement Grant #7138 from the Wenner Gren Foundation.
References Agar N, Young A. 2005. Melanogenesis: A photoprotective response to DNA damage? Mutation Research 571: 121–32. Akey J, Zhang K, Xiong M, Jin L. 2003. The effect of single nucleotide polymorphism identification strategies on estimates of linkage disequilibrium. Molecular Biology and Evolution 20: 232–42. Allen J. 1996. The Pre-Austronesian settlement of Island Melanesia: Implications for Lapita archaeology. In: Goodenough W, editor. Prehistoric settlement of the Pacific. Philadelphia, PA: Transactions of the American Philosophical Society 86. pp 11–27. Barnicot N. 1958. Reflectometry of the skin in southern Nigerians and in some mulattoes. Human Biology 30: 150–60. Bennett D, Lamoreux M. 2003. The color loci of mice—a genetic century. Pigment Cell Research 16: 334–44. Blum H. 1961. Does the melanin pigment of human skin have adaptive value? Quarterly Review of Biology 36: 50–63. Bonilla C, Boxill L, Donald S, Williams T, Sylvester N, Parra E, Dios S, Norton H, Shriver M, Kittles R. 2005. The 8818G allele of the agouti signaling protein (ASIP) gene is ancestral and is associated with darker skin color in African Americans. Human Genetics 116: 402–6.
109
core studies in northern island melanesia
Bower C, Stanley F. 1989. Dietary folate as a risk factor for neural tube defects: Evidence from a case-control study in Western Australia. Medical Journal of Australia 150: 613–19. Box N, Wyeth J, O’Gorman L, Martin N, Sturm R. 1997. Characterization of melanocyte stimulating hormone receptor variant alleles in twins with red hair. Human Molecular Genetics 6: 1891–7. Branda R, Eaton J. 1978. Skin color and nutrient photolysis: An evolutionary hypothesis. Science 201: 625–6. Byard P, Lees F. 1982. Skin colorimetry in Belize. II. Interand Intra-population variation. American Journal of Physical Anthropology 58: 215–19. Chaplin G. 2004. Geographic distribution of environmental factors influencing human skin coloration. American Journal of Physical Anthropology 125: 292–302. Conway D, Baker P. 1972. Skin reflectance of Quechua Indians: The effects of genetic admixture, sex, and age. American Journal of Physical Anthropology 36: 267–82. Cowles R. 1959. Some ecological factors bearing on the origin and evolution of pigment in the human skin. American Naturalist 93: 283–93. Culotta E. 2005. Physical anthropology/paleoanthropology meetings. Snapshots from the meeting. Science 308: 491. Darwin C. 1871. The descent of man, and selection in relation to sex. London: J. Murray. Diamond J. 1992. The third chimpanzee. New York: Harper Collins Publishers. Diffey B, Oliver R, Farr P. 1984. A portable instrument for quantifying erythema induced by ultraviolet radiation. British Journal of Dermatology 111: 663–72. Duffy D, Box N, Chen W, Palmer J, Montgomery G, James M, Hayward N, Martin N, Sturm R. 2004. Interactive effects of MC1R and OCA2 on melanoma risk phenotypes. Human Molecular Genetics 13: 447–61. Flanagan N, Healy E, Ray A, Philips S, Todd C, Jackson I, Birch-Machin M, Rees J. 2000. Pleiotropic effects of the melanocortin 1 receptor (MC1R) gene on human pigmentation. Human Molecular Genetics 9: 2531-7. Flemming A, Copp A. 1998. Embryonic folate metabolism and mouse neural tube defects. Science 280: 2107–9. Friedmann P, Gilchrest B. 1987. Ultraviolet radiation directly induced pigment production by cultured human melanocytes. Journal of Cellular Physiology 133: 88–94. Friedlaender JS. 1975. Patterns of human variation: The demography, genetics, and phenetics of Bougainville Islanders. Cambridge, MA: Harvard University Press. Frost P. 1988. Human skin color: A possible relationship between its sexual dimorphism and its social perception. Perspectives in Biology and Medicine 32: 38–58. Giebel L, Musarella M, Spritz R. 1991. A nonsense mutation in the tyrosinase gene of Afghan patients with tyrosinase negative (type IA) oculocutaneous albinism. Journal of Medical Genetics 28: 461–7. Giles E, Ogan E, Steinberg A. 1965. Gamma-globulin factors (Gm and Inv) in New Guinea: Anthropological significance. Science 150: 1158–60.
110
Graf J, Hodgson R, van Daal A. 2005. Single nucleotide polymorphisms in the MATP gene are associated with normal human pigmentation variation. Human Mutation 25: 278–84. Harding R, Healy E, Ray A, Ellis N, Flanagan N, Todd C, Dixon C, Sajantila A, Jackson I, Birch-Machin M, Rees J. 2000. Evidence for variable selective pressures at MC1R. American Journal of Human Genetics 66: 1351–61. Harsanyi Z, Post P, Brinkmann J, Chedekel M, Diebel R. 1980. Mutagenicity of melanin from human red hair. Experientia 36: 291–2. Harvey R. 1985. Ecological factors in skin color variation among Papua New Guineans. American Journal of Physical Anthropology 66: 407–16. Inagaki K, Suzuki T, Shimizu H, Ishii N, Umezawa Y, Tada J, Kikuchi N, Takata M, Takamori K, Kishibe M, Tanaka M, Miyamura Y, Ito S, Tomita Y. 2004. Oculocutaneous albinism type 4 is one of the most common types of albinism in Japan. American Journal of Human Genetics 74: 466–71. Jablonski N. 1992. Sun, skin colour and spina bifida: An exploration of the relationship between ultraviolet light and neural tube defects. Proceedings of the Australasian Society of Human Biology 5: 455–62. Jablonski N, Chaplin G. 2000. The evolution of human skin coloration. Human Evolution 39: 57–106. Kanetsky P, Swoyer J, Panossian S, Holmes R, Guerry D, Rebbeck T. 2002. A polymorphism in the agouti signaling protein gene is associated with human pigmentation. American Journal of Human Genetics 70: 770-5. King R, Mentink M, Oetting W. 1991. Non-random distribution of missense mutations within the human tyrosinase gene in type I (tyrosinase-related) oculocutaneous albinism. Molecular Biology and Medicine 8: 19–29. Koch W, Chedekel M. 1986. Photoinitiated DNA damage by melanogenic intermediates in vitro. Photochemistry and Photobiology 44: 703–10. Kollias N, Sayre R, Zeise L, Chedekel M. 1991. Photoprotection by melanin. Journal of Photochemistry and Photobiology B 9: 135–60. Leavesley M, Allen J. 1998. Dates, disturbance and artefact distributions: Another analysis of Buang Merabek, a Pleistocene site on New Ireland, Papua New Guinea. Archaeology in Oceania 33: 63–82. Lee S, Nicholls R, Schnur R, Guida L, Lu-Kuo J, Spinner N, Zackai E, Spritz R. 1994. Diverse mutations of the P gene among African-Americans with type II (tyrosinase positive) oculocutaneous albinism (OCA2). Human Molecular Genetics 3: 2047–51. Loomis W. 1967. Skin pigment regulation of vitamin-D biosynthesis in man. Science 157: 501–6. Mackintosh J. 2001. The antimicrobial properties of melanocytes, melanosomes, and melanin and the evolution of black skin. Journal of Theoretical Biology 211: 101–13. Manga P, Kromberg J, Box N, Sturm R, Jenkins T, Ramsay M. 1997. Rufous oculocutaneous albinism in
Pigmentation and Candidate Gene Variation in Northern Island Melanesia
southern African Blacks is caused by mutations in the TYRP1 gene. American Journal of Human Genetics 61: 1095–101. Marshall B, Allen J. 1991. Excavations at Panakiwuk Cave, New Ireland. In: Allen J, Gosden C, editors. Report of the Lapita homeland project. Canberra, Australia: Research School of Pacific Studies. pp 59–91. Merriwether D, Friedlaender JS, Mediavilla J, Mgone C, Gentz F, Ferrell R. 1999. Mitochondrial DNA variation is an indicator of Austronesian influence in Island Melanesia. American Journal of Physical Anthropology 110: 243–70. Montagna W, Carlisle K. 1991. The architecture of black and white facial skin. Journal of the American Academy of Dermatologists 24: 929–37. Murray F. 1934. Pigmentation, sunlight and nutritional disease. American Anthropologist 36: 438–45. MVSR. 1991. Prevention of neural tube defects: Results of the Medical Research Council Vitamin Study. Lancet 338: 131–7. Nakayama K, Fuakamachi S, Kimura H, Koda Y, Soemantri A Ishida T. 2002. Distinctive distribution of AIM1 polymorphism among major human populations with different skin color. Journal of Human Genetics 47: 92–4. Newton J, Cohen-Barak O, Hagiwara N, Gardner J, Davisson M, King R, Brilliant M. 2001. Mutations in the human orthologue of the mouse underwhite gene (uw) underlie a new form of oculocutaneous albinism, OCA4. American Journal of Human Genetics 69: 981–8. Nordlund J, Boissy R, Hearing V, King R, Ortonne J. 1998. The pigmentary system—physiology and pathophysiology. Oxford: Oxford University Press. Norton H. 2005. Human skin pigmentation variation: A phenotypic, genotypic, and evolutionary perspective. Department of Anthropology, The Pennsylania State University, University Park, PA. Norton H, Friedlaender JS, Merriwether D, Koki G, Mgone C, Shriver M. 2006. Skin and hair pigmentation variation in Island Melanesia. American Journal of Physical Anthropology 130: 254–68. Pathak M, Fitzpatrick T. 1974. The role of natural photoprotective agents in human skin. In: Pathak M, editor. Sunlight and man. Tokyo: University of Tokyo Press. pp 725–50. Peng S, Lu X, Luo H, Xiang-Yu J, Zhang Y. 2001. Melanocortin-1 receptor gene variants in four Chinese ethnic populations. Cell Research 11: 81–4. Post P, Daniels F, Binford R. 1975. Cold injury and the evolution of ‘white’ skin. Human Biology 39: 131–43. Rana B, Hewett-Emmett D, Jin L, Chang B, Sambuughin N, Lin M, Watkins S, Bamshad M, Jorde L, Ramsay M, Jenkins T, Li W. 1999. High polymorphism at the human melanocortin 1 receptor locus. Genetics 151: 1547–57. Roberts D, Kahlon D. 1976. Environmental correlations of skin colour. Annals of Human Biology 3: 11–22. Rogers A, Iltis D, Wooding S. 2004. Genetic variation at the MC1R locus and the time since loss of human body hair. Current Anthropology 45: 105–8.
Rundshagen U, Zuhlke C, Opitz S, Schwinger E, KasmannKellner B. 2004. Mutations in the MATP gene in five German patients affected by oculocutaneous albinism type 4. Human Mutation 23: 106–10. Schmitz S, Thomas P, Allen T, Poznansky M, Jimbow K. 1995. Dual role of melanins and melanin precursors as photoprotective and phototoxic agents: Inhibition of ultraviolet radiation-induced lipid peroxidation. Photochemistry and Photobiology 61: 650–5. Serjeantson S, Gao X. 1995. Homo sapiens is an evolving species: Origins of the Austronesians. In: Bellwood P, Fox J, Tryon D, editors. The Austronesians: Historical and comparative perspectives. Canberra, Australia: Australian National University. Sheehan J, Cragg N, Chadwick C, Potten C, Young A. 2002. Repeated ultraviolet exposure affords the same protection against DNA photodamage and erythema in human skin types II and IV but is associated with faster DNA repair in skin type IV. Journal of Investigative Dermatology 118: 825–9. Shriver M, Parra E, Dios S, Bonilla C, Norton H, Jovel C, Pfaff C, Jones C, Massac A, Cameron N, Baron A, Jackson T, Argyropoulos G, Jin L, Hoggart C, McKeigue P, Kittles R. 2003. Skin pigmentation, biogeographical ancestry and admixture mapping. Human Genetics 112: 387–99. Smith R, Healy E, Siddiqui S, Flanagan N, Steijlen P, Rosdahl I, Jacques J. 1998. Melanocortin 1 receptor variants in Irish populations. Journal of Investigative Dermatology 111: 119–22. Spritz R, Fukai K, Holmes S, Luande J. 1995. Frequent intragenic deletion of the P gene in Tanzanian patients with type II oculocutaneous albinism (OCA2). American Journal of Human Genetics 56: 1320–3. Stevens G, van Beukering J, Jenkins T, Ramsay M. 1995. An intragenic deletion of the P gene is the common mutation causing tyrosinase-positive oculocutaneous albinism in Southern African Negroids. American Journal of Human Genetics 56: 586–91. Sturm R. 1998. Human pigmentation genes and their response to solar UV radiation. Mutation Research 422: 69–76. Sturm R, Teasdale R, Box N. 2001. Human pigmentation genes: Identification, structure and consequences of polymorphic variation. Genes and Genetic Systems 277: 49–62. Summerhayes G, Allen J. 1993. The transport of Mopir obsidian to Late Pleistocene New Ireland. Archaeology in Oceania 28: 145–9. Suzuki T, Miyamura Y, Matsunaga J, Shimizu H, Kawachi Y, Ohyama N, Ishikawa O, Ishikawa T, Terao H, Tomita Y. 2003. Six novel P gene mutations and oculocutaneous albinism type 2 frequency in Japanese albino patients. Journal of Investigative Dermatology 120: 781–3. Tobias P. 1961. Studies on skin reflectance in BushmenEuropean hybrids. Proceedings, Second International Congress of Human Genetics, Rome: 461–71. Tomita Y, Takeda A, Okinaga S, Tagami H, Shibahara S. 1989. Human oculocutaneous albinism caused by single base insertion in the tyrosinase gene. Biochemical and Biophysical Research Communications 164: 990–6.
111
core studies in northern island melanesia
Valverde P, Healy E, Jackson I, Rees J, Thody A. 1995. Variants of the melanocyte stimulating hormone receptor gene are associated with red hair and fair skin. Nature Genetics 11: 328–30. Van den Berghe P, Frost P. 1986. Skin color preference, sexual dimorphism and sexual selection: A case of gene culture co-evolution? Ethnic and Racial Studies 9: 87–113. Walter H. 1971. Remarks on the environmental adaptation of man. Humangenetik 13: 85–97. Wasserman H. 1965. Human pigmentation and environmental adaptation. Archives of Environmental Health 11: 691–4.
112
Westbroek W, Lambert J, Naeyaert J. 2001. The dilute locus and Griscelli syndrome: Gateways towards a better understanding of melanosome transport. Pigment Cell Research 14: 320–7. Yao Y, Lu X, Luo H, Li W, Zhang Y. 2000. Gene admixture in the silk road region of China: Evidence from mtDNA and melanocortin 1 receptor polymorphism. Genes and Genetic Systems 75: 173–8. Yi Z, Garrison N, Cohen-Barak O, Karafet T, King R, Erickson R, Hammer M, Brilliant M. 2003. A 122.5-kilobase deletion of the P gene underlies the high prevalence of oculocutaneous albinism type 2 in the Navajo population. American Journal of Human Genetics 72: 62–72.
7 The Distribution of an Insertion/Deletion Polymorphism on Chromosome 22 Renato Robledo
Introduction Mitochondrial DNA (mtDNA) is inherited exclusively from the maternal side: therefore, analysis of mtDNA polymorphisms will describe human variation as the result of establishment and expansion of female-derived lineages. Likewise, the Y chromosome is inherited solely from the paternal side and analysis of Y-linked chromosomal markers will describe human variation in terms of establishment and expansion of male-derived lineages. On the other hand, autosomal markers are bi-parental since they are inherited from both maternal and paternal sides in equal proportion. The picture of human variation described by the analysis of autosomal markers is therefore not restricted to male-only or female-only lineages, and the analysis of autosomal markers will reflect the major migration patterns by both sexes. The most abundant source of genetic variation is provided by the single nucleotide polymorphisms (SNPs); however, a major problem with the SNPs is the possible independent re-occurrence of the same mutation, which may confound the interpretation of the data. An alternative source of human genomic variation is provided by a distinct class of polymorphic markers due to insertions or deletions (indel). In view of the erratic mechanisms that generate insertion and deletion events, the possibility of a second occurrence of an insertion or a deletion at the very same site can be confidently ruled out. Therefore, individuals who share an indel polymorphism (i.e. an Alu insertion) at a specific locus are considered identical by
descent (IBD). For this reason, indel polymorphisms are particularly useful markers in describing human genomic variation in the fields of population genetics, evolution or anthropology. Populations sharing an indel polymorphism can be said to have some relationship. In the early 1990s, Wigler and colleagues reported the development of a new technique, called Representational Difference Analysis (RDA), which detects differences between two complex genomes (Lisitsyn et al., 1993; Lisitsyn and Wigler, 1995). Applications of RDA led to the identification of a DNA sequence, named R271, possibly involved in a kidney carcinoma (Lisitsyn et al., 1995). A pair of primers was developed for a fast PCR screening that detects the presence of a 179-bp amplicon, included in the original R271 sequence: the entire sequence of the RDA-derived R271 clone with the forward and reverse primers is fully reported (Robledo et al., 2002). Unexpectedly, a first screening indicated that approximately 20% of normal individuals did not show any PCR product, suggesting the occurrence of a deletion. One possible interpretation was that the R271 deletion was a rather common polymorphism. A more extensive screening conducted in different genetic isolates living in Sardinia confirmed that the R271 is an example of indel polymorphism. Moreover, we conducted a segregation analysis performed in several Sardinian pedigrees, as well as in five large, multigenerational, CEPH pedigrees. The results (Robledo, unpublished results) are consistent with a stable polymorphic marker that is inherited in a simple Mendelian fashion.
113
core studies in northern island melanesia
Sequence comparison with the human reference sequence allowed us to identify the exact endpoints of the polymorphism: we were able to conclude that the R271 is an indel polymorphism of 9.1 kb. The knowledge of the precise endpoints of the polymorphism allowed us to develop an accurate and easy assay, described below, which unambiguously identifies the three possible genotypes: 9.1-kb (+/+), 9.1-kb (+/−), 9.1-kb (−/−). We extended our study to include a worldwide population screening to estimate the world distribution of the indel polymorphism (Siniscalco et al., 2000). The results showed that the polymorphism is present, with markedly different frequencies, in almost all the population tested: the genotypic and allele frequencies were calculated (figure 7.1). The results from the population analysis suggest that the polymorphism has an ancient origin, possibly pre-dating the out-of-Africa expansion of early humans. Interestingly, we noted that the only population that showed no polymorphism was a small sampling (22 individuals) of the Nasioi-speaking population of Bougainville, in Island Melanesia, where all 44 chromosomes tested showed the 9.1-kb– allele.
A subsequent fine mapping localized the polymorphism to band 22q11.2 (Robledo et al., 2002), within the cluster of the immunoglobulin variable lambda lightchain genes (Ig V λ). The localization offers some speculation that the indel might be under selective constraints. However, sequence analysis has showed that the 9.1-kb fragment does not contain coding sequences: we found one IgλV pseudo-gene, as well as one IgλJ pseudo-gene (Roe, personal communication). The lack of any expressed sequences or active genes suggests that the indel polymorphism may be selectively neutral. Finally, comparative genomic analysis of chromosomes in nonhuman primates suggests that the ancestral allele is the 9.1-kb+, and that the 9.1-kb– allele is the result of a deletion that occurred in the human lineage (Robledo et al., 2004).
Results The methodology has been described in detail elsewhere (Robledo et al., 2003). Briefly, the genotyping was performed
Figure 7.1 World-wide frequency distribution of the 9.1-kb+ indel allele on chromosome 22.
114
The Distribution of an Insertion/Deletion Polymorphism on Chromosome 22
by using a multiplex PCR with two primer pairs: R271 and Del9. The sequences of the primers are: R271-fwd: 5’-CTCAGCTAAGAATCCTCAGAGGATTG-3’ R271-rev: 5’-GCCATCTTCCATTTTGGTATCAGTGC-3’ Del9-fwd: 5’-AGGAGGTTGTAAGCAAGGAG-3’ Del9-rev: 5’-CCTAAGGACCTGTAAGGACAC-3’
R271 primers amplify a 179-bp fragment that is present only on the 9.1-kb+ chromosomes; therefore the presence of the 179-bp band will identify specifically the 9.1-kb+ allele. Del9 primers amplify a 285-bp fragment that is present only on the 9.1-kb– chromosomes; therefore the presence of the 285-bp band will identify specifically the 9.1-kb– allele. By using a multiplex PCR with both primer pairs, individuals who are homozygous for the 9.1-kb+ chromosomes will show exclusively the 179-bp band; individuals who are homozygous for the 9.1kb– chromosomes will show exclusively the 285-bp band; heterozygous individuals will show both bands (figure 7.2). The earlier worldwide population study included a small sampling (n = 22) of Nasioi, a Papuan-speaking group that lives in the island of Bougainville. Remarkably, we found that all 44 chromosomes carried the 9.1-kb– allele: therefore, the Nasioi was the first example of a population that was monomorphic for the 9.1-kb marker. In contrast, results from Asiatic populations showed frequencies of the Figure 7.2 A, The difference in the form of the two 9.1 kb alleles. B, Multiplex PCR showing the two bands used in identifying homozygotes and heterozygotes for the 9.1 kb indel.
9.1-kb– allele ranging from 0.45 to 0.70 (Siniscalco et al., 2000; Robledo et al., 2002). These findings motivated us to expand our observations in Island Melanesia. The present study reports the allelic and genotypic frequencies in 21 populations that live in the islands of Mussau, New Hanover, New Ireland, and New Britain of the Bismarck Archipelago, as well as small samplings from New Guinea (figure 7.3). The isolates tested include 13 Austronesianspeaking (AN) and 8 non-Austronesian (NAN) or Papuanspeaking populations (table 7.1). The results, summarized in figure 7.3 and table 7.1, clearly show that the 9.1-kb– is the most common allele throughout the region. A frequency of the 9.1-kb– allele greater than 50% has been observed in different islands and in every population investigated, whether AN-speaking or Papuan-speaking. The distribution of the 9.1-kb polymorphism is rather uniform in the island of New Ireland: the frequency of the 9.1-kb– allele ranges from 0.69, as observed in the AN-speaking Notsi, to 0.83, as observed in the AN-speaking Tigak. Similar values were recorded also for the Papuan-speaking Kuot (0.78). We note that the island of New Ireland is rather narrow and all populations live along the coastal areas of the island. A similar distribution of the 9.1-kb– allele is present in the AN-speaking Lavongai (0.74) and the AN-speaking Kapugu (0.81) who live along the coastal areas of the islands of New Hanover and Mussau, respectively. On the other hand, the distribution of the 9.1-kb polymorphism in New Britain shows high variability across the different regions of the island. The frequency of the 9.1-kb– allele ranges from 0.68, as observed in the ANspeaking Nakanai (Bileki dialect), to 0.99, as observed in the Papuan-speaking Ata, as well as in the AN-speaking Mamusi. The variation among villages living in different parts of the island is highly significant (p<0.001). We note that the polymorphism is distributed according to a distinct pattern: the highest frequencies, approaching fixation, of the 9.1-kb– allele are found among the populations, whether AN-speaking or Papuan-speaking, who live in the interior, more remote and less accessible, regions of the island. Indeed, in addition to the 0.99 value for the 9.1-kb– allele found in the above-mentioned Ata and Mamusi, we found a frequency of 0.96 among the Papuan-speaking Baining (Kaket dialect), and 0.93 among the AN-speaking Mengen. Instead, the lowest frequency of the 9.1-kb– allele is found among the shore-dwelling populations of the AN-speaking Nakanai (Bileki dialect) and the Papuan-speaking Baining (Mali dialect), with values of 0.68 and 0.76, respectively. A distribution pattern similar to that observed in New Britain appears to hold true also in New Guinea: we found that the samples of New Guineans from the Trans New Guinea Phylum area that live in the interior
115
core studies in northern island melanesia
Figure 7.3 Frequency distribution of the 9.1 kb+ indel allele in Island Melanesia.
were all homozygous for the 9.1-kb– allele, like the Nasioi of Bougaiville. In contrast, the even smaller sampling of a coastal Sepik population was found to be polymorphic, with a frequency of the 9.1-kb– allele of 0.86 (figure 7.3).
Conclusion Our data support the idea that language relationships, together with geographical barrier and distance, can play a major role in determining patterns of human genome diversity. We have described a repeating pattern of genetic diversity which is commonly shared by different islands of significant size, like Bougainville, New Britain and,
116
possibly, New Guinea. Papuan-speaking populations that live in the more remote and inaccessible areas of these islands, like the Nasioi in Bougainville and the Ata in New Britain, tend to share the near fixation for the 9.1-kb– allele. However, we saw a near fixation for the 9.1-kb– allele also in the AN-speaking Mamusi that live in the interior of New Britain. On the other hand, we saw a different profile in the populations that live on the coastal areas: these populations, whether AN-speaking or Papuan-speaking, tend to show higher frequencies of the 9.1-kb+ allele. The described pattern of genetic diversity is consistent with independent sets of data derived by mtDNA analysis, Y chromosome markers, as well as other autosomal markers, like the ones involved in skin pigmentation, which are described in different chapters of this section.
The Distribution of an Insertion/Deletion Polymorphism on Chromosome 22
Table 7.1
Distribution of the 9.1 kb− indel Allele in the Bismarck Archipelago and New Guinea
Island
Population
Phylum
2N
New Guinea
Sepik Non-Sepik Anêm Kove Nakanai (Bileki) Nakanai (Loso) Ata Mamusi Mangseng Melamela Kol Mengen Tolai Sulka Baining (Mali) Baining (Kaket) Kapugu Lavongai Tigak Nalik Notsi Kuot Madak Aita Teop Saposa Nasioi
Papuan Papuan Papuan AN AN AN Papuan AN AN AN Papuan AN AN Papuan Papuan Papuan AN AN AN AN AN Papuan AN Papuan AN AN Papuan
22 28 36 40 72 22 120 104 14 32 8 42 96 68 50 82 36 84 54 30 32 82 40 34 44 56 44 1372
New Britain
Mussau New Hanover New Ireland
Bougainville
Total chromosomes
References Lisitsyn N, Wigler M. 1995 Representational difference analysis in detection of genetic lesions in cancer. Methods in Enzymology 254: 291–304. Lisitsyn N, Lisitsyn N, Wigler M. 1993. Cloning the differences between two complex genomes. Science 259: 946–51. Lisitsyn N, Lisitsina N, Dalbagni G, Barker P, Sanchez CA, Gnarra J, Linehan WM, Reid B, Wigler M. 1995. Comparative genomic analysis of tumors: Detection of DNA losses and amplification. Proceedings of the National Academy of Sciences USA 92: 151–5. Robledo R, Orru S, Sidoti A, Muresu R, Esposito D, Grimaldi MC, Carcassi C, Rinaldi A, Bernini L, Contu L, Romani M, Roe B, Siniscalco M. 2002. A 9.1-kb gap
p(−) 19 28 23 37 49 20 119 103 11 26 7 39 82 55 38 79 29 62 45 23 22 64 31 33 41 41 44
(−)%
SD
86 100 64 93 68 91 99 99 79 81 88 93 85 81 76 96 81 74 83 77 69 78 78 97 93 73 100
0.07 0.00 0.08 0.04 0.05 0.06 0.01 0.01 0.11 0.07 0.12 0.04 0.04 0.05 0.06 0.02 0.07 0.05 0.05 0.07 0.08 0.05 0.07 0.03 0.04 0.06 0.00
in the genome reference map is shown to be a stable deletion/insertion polymorphism of ancestral origin. Genomics 80: 585–92. Robledo R, Scheinfeldt L, Merriwether DA, Thompson F, Friedlaender J. 2003. A 9.1 kb insertion/deletion polymorphism suggests a common pattern of genetic diversity in Island Melanesia. Human Biology 75: 941–9. Robledo R, Bender P, Leonard J, Zhu B, Osoegawa K, de Jong P, Xu X, Yao Z, Roe B. 2004. The immunoglobulin λ variable light-chain region in Primates has been shaped by multiple, independent, small-scale and large-scale insertion/deletion events. Genomics 84: 678–85. Siniscalco M, Robledo R, Orru S, Contu L, Yadav P, Ren Q, Lai H, Roe B. 2000. A plea to search for deletion polymorphism through genome scans in populations. Trends in Genetics 16: 435–7.
117
8 The Languages of Island Melanesia Eva Lindström, Angela Terrill, Ger Reesink, and Michael Dunn
Introduction This chapter follows on from the description of the main groups of Papuan languages given by Pawley in chapter 3, by focusing on the linguistic landscape of the bigger islands to the east and southeast of New Guinea, i.e., Island Melanesia, and describing both Austronesian and Papuan languages in that region. The Papuan languages, descending from the languages of the region prior to the Lapita expansion (see chapter 2), have proved intractable to the best established method in historical linguistics, the Comparative Method (CM), presumably due to the great time depth and fragmentary nature of the evidence. We will show both how that method is applied to the Austronesian languages and what the results can tell us about prehistory; as well as why it cannot be applied to the older, Papuan, stratum of languages. The history and relations of the Papuan languages are explored with the aid of a database of abstract structural linguistic features. First a cladistic tree structure is described (Dunn et al., 2005), then contact-induced similarity is investigated by comparison of the structural and geographic distances of pairs of Papuan and Oceanic languages, per group and between groups. While these methods are no replacement for the CM, they complement it, both in that they can be applied in situations of greater time depth, and in that they are a useful way of exploring the structural relations among languages separated at any time depth. The focus here will be on the two major types of languages in Island Melanesia; the Oceanic languages, closely related members of subgroups of a single family; and the Papuan languages, a disparate and dispersed group whose relations are still unclear. Below, we describe
118
the distribution of languages in the area under study, and then outline the main methods that have been used to investigate their linguistic relations to each other. The rest of the chapter has four sections. The first discusses the Oceanic languages, beginning with a typological sketch of their linguistic characteristics, followed by a description of historical reconstruction as it has been applied to these languages, and the sorts of insights that the CM can yield about prehistory. The second main section similarly starts with a typological sketch of the Papuan languages, in as far as that is possible for such disparate languages, then outlines various attempts that have been made to uncover the historical connections between them. The ever-present issue of linguistic contact is the topic of the third section, and known effects of contactinduced change on the languages of the region are described, and the effect of contact-induced language change in uncovering Papuan language histories is discussed. The last section forms the conclusions of this chapter and discusses various hypotheses that can be advanced regarding the linguistic prehistory of Island Melanesia.
The Linguistic Scene in Island Melanesia The region of Northern Island Melanesia is home to more than 150 languages belonging to various subgroups of Oceanic, itself a subgroup of the widespread Austronesian language family (Lynch et al., 2002; see chapter 3). The scientific consensus is that Austronesian-speaking groups migrated from Taiwan roughly 6,000 years ago and that speakers of (a precursor of) Proto-Oceanic (POc) traveled from the Cenderawasih Bay in western New Guinea in present-day Indonesia to its homeland somewhere near the Willaumez peninsula of New Britain 3,300 years
The Languages of Island Melanesia
ago (chapter 2; Lynch et al., 2002: 37). Interspersed in Northern Island Melanesia are just over 20 Papuan languages that are not clearly genealogically related. The term ‘Papuan’ does not imply any linguistic unity, but simply denotes languages of the greater New Guinea area that are non-Austronesian. The exact number of Papuan languages in Island Melanesia depends on how many languages are distinguished in the Baining family of East New Britain. Stebbins (2002) listed Asimbali, Kaket, Kairak, Ura (or Uramet), and Mali, in addition to Taulil and extinct Butam, which are either unrelated or remotely related to the Baining languages. With respect to the languages of the Santa Cruz Islands, Northern Santa Cruz, Äyiwo/Reefs, and Nanggu, their status as Austronesian or Papuan is still a matter of debate. These Papuan languages are most likely remote descendants of the languages spoken by people who had colonized the islands long before the Austronesians arrived (Dunn et al., 2002: 29). Their geographic distribution is given below (see figure 8.1). We use a restricted definition of ‘Island Melanesia’ in this chapter, including only the areas where both Papuan
and Oceanic languages are found, that is, the Bismarck Archipelago (New Ireland and New Britain), Bougainville, and the Solomon Islands (to the general exclusion of the Admiralties, Vanuatu, New Caledonia, and Fiji). Further, although “Solomon Islands” is a geographic term for a larger area, it will be used here to denote the political unit (also “Solomons”), while “Bougainville” will be used for the portion of that geographic area which is in Papua New Guinea (the North Solomons Province).
New Britain Aneˆm is the sole Papuan language in the midst of the Siassi group of the North New Guinea linkage (NNG) of Oceanic languages (Thurston, 1987) in West New Britain. It appears to be distantly related to Ata, the Papuan language spoken 200 km to the east. These languages are geographically separated by languages of the Willaumez linkage, a subgroup of the Meso-Melanesian linkage of Oceanic, including Nakanai. Along the south coast of New Britain the heterogeneous Southwest New Britain linkage of the North New Guinea linkage of Oceanic includes languages such as Mangseng and Mamusi,
Figure 8.1 Language map showing the locations of all Papuan languages in Island Melanesia and those Oceanic languages that are mentioned in figures 8.2–8.4. Butam and Kazukuru are extinct. Mali and Kaket are part of a small cluster known as the Baining languages (the other members are Asimbali, Kairak and Ura (Uramet)).
119
core studies in northern island melanesia
southern neighbors of Papuan Ata. Across an uninhabited area east of Oceanic Mamusi and Papuan Ata, a few dialects of Mengen are spoken, the eastern-most member of the NNG linkage of Oceanic. The Papuan isolate Kol is spoken in the interior of East New Britain, located between a number of Oceanic languages: Melamela, a close relative of Nakanai, in the north, Mengen to the south, and Tomoip to the southeast. Another Papuan isolate, Sulka, borders on the two Oceanic languages Mengen (Poeng dialect) and Tomoip. It is spoken mainly along the Wide Bay and, for about the last 100 years, in three villages in the Gazelle peninsula. Reesink (2005a) provided some evidence of linguistic borrowings between the Oceanic languages Mengen and Tomoip and the Papuan languages Kol and Sulka. The Gazelle Peninsula is for the most part inhabited by the Papuan Baining language family and Tolai (Kuanua), a member of the St. George linkage of the MesoMelanesian linkage of Oceanic languages.
New Ireland Kuot is the only Papuan language spoken on New Ireland, surrounded by three subgroups of the Oceanic MesoMelanesian linkage.
Bougainville Bougainville is home to the most clearly related group of Papuan languages, divided into two recognized unrelated families: North Bougainville with the members Eivo, Rotokas, Konua (or Raiposi), and Keriaka; and South Bougainville: Nasioi, Nagovisi, Motuna (or Siwai), and Buin. Buka Island, the north of Bougainville and some pockets on the east and west coast are home to various Oceanic languages, all members of the Northwest Solomonic linkage.
Solomon Islands The Northwest Solomonic linkage is spread out over the central Solomon Islands, with only four Papuan languages interspersed: Bilua on Vella Lavella, Touo (or Baniata) on Rendova, Lavukaleve on the Russell Islands, and Savosavo on Savo Island. Possibly, some of the languages on the Santa Cruz and Reef Islands are Papuan as well (Lincoln, 1978; Wurm, 1978).
Different Language Histories, Different Linguistic Methodologies The Comparative Method Genealogical relationships between languages are normally established by the only generally accepted linguistic
120
method, named the Comparative Method (CM), which presupposes a shared ancestor for the languages under comparison, and works by identifying and comparing cognate sets of lexical items and morphological material. Words or morphemes are cognate when they are related in meaning and derived through regular sound changes from a common ancestor. Comparing sets of cognates allows reconstruction of possible ancestral forms, i.e., a proto-language (chapter 3; Campbell, 1998: 112–132; Ross and Durie, 1996). The different steps involve (i) identifying candidate corresponding words, (ii) restricting the semantic denotations, (iii) working out the sound correspondences, and (iv) reconstruction of the protophonology, followed by (v) the reconstruction of the proto-morphemes, and (vi) the establishment of subgroups of languages based on shared innovations. Language change is often unidirectional, which greatly facilitates linguistic reconstruction. For example, s frequently develops into h, which may later disappear altogether, while the reverse does not occur. There is a large body of linguistic work on directionality in language change, which often enables two different states to be dated relative to each other. It is also known what types of changes are so common that they occur independently and lack diagnostic value. However, the CM cannot operate in all linguistic settings and across all types of linguistic material (Harrison, 2003; Rankin, 2003). A major limitation with the Papuan languages in Island Melanesia is that all linguistic material gradually erodes over the millennia through natural processes of sound change, which after a long enough period of time renders any cognates no longer recognizable, even without lexical replacement. With no cognates to identify and reconstruct, there is no material for the CM. This limitation, essentially one of time depth, means that ancestral linguistic states going back beyond around 8,000 ± 2,000 years (Nichols, 1992) can no longer be detected. As human settlement in the region dates back well into the Pleistocene (chapter 2; Kirch, 1997; Spriggs, 1997), it is almost a certainty that the ancestral forms of at least some of the Papuan languages have been in situ beyond the 8,000 ± 2,000 year ceiling. The regional diversity and the absence of clear family-like relationships in many cases, also suggest very great time depths of development.
Other Methods Attempts to uncover distant genetic relationships not using the CM have mostly been unsuccessful (Campbell, 2003). Perhaps the best-known method outside of linguistics is the technique used by Greenberg (e.g., 1987) known as Mass Comparison. This method ignores the requirement that other causes of linguistic correspondence, such as borrowing and chance similarities, should not
The Languages of Island Melanesia
contaminate the evidence for genealogical relationships (Campbell, 1998: 315). It also widens the acceptable semantic domain correspondingly. The method simply looks at a few words with in some cases loosely related meanings across as many languages as possible and counts any resemblance in form as evidence for genealogical relationship. By amassing a mixed bag of true cognates and accidental look-alikes from many languages belonging to a great number of families, Greenberg (1971, not using mass comparison) advanced his hypothesis of an Indo-Pacific phylum, which contained all languages of Oceania from the Andaman Islands to the barely attested languages of Tasmania, with the exception of the Austronesian and mainland Australian languages. Partly following Greenberg, but with reservations regarding the Austronesian loans in a number of languages, Wurm (1975a: 783) claimed to have established an East Papuan phylum, comprising all the Papuan languages of Island Melanesia: Yélıˆ Dnye of Rossel Island off the southeastern tip of Papua New Guinea, and all the non-Austronesian languages of the Bismarck and Solomon archipelagos. His evidence was based on low to moderate numbers of putative cognates and some shared pronominal forms and typological features. However, Ross (2001a) and Dunn et al. (2002) have challenged Wurm’s classification, because the adduced evidence does not pass the scrutiny of the more stringent requirements of the CM. Clearly, if the Papuan languages of Island Melanesia did have a common ancestor, it must have predated the time barrier inherent to the Comparative Method. Does that mean that linguistics has nothing to add to our knowledge about the earliest colonization of the region, as Harrison (2003) and others seem to imply? For reasons explained below, our answer to this question is ‘no’.
Computational Methods Computational linguistic approaches to linguistic relationships are becoming more common (e.g., Gray and Jordan, 2000; Ringe et al., 2002; Forster and Toth, 2003; Gray and Atkinson, 2003; McMahon and McMahon, 2003; Rexová et al., 2003). All of these have worked on known language families and relied on comparison of words and forms of other morphemes. They have also relied on the work of linguists to identify the related forms (cognates) to be compared. In a recent paper Dunn et al. (2005) outlined a computational approach to uncovering linguistic prehistory among the Papuan languages of Island Melanesia. This paper presented a way of dealing with languages so distantly connected that no formal cognates can be discovered. Unlike the CM, which relies on comparing similar forms, Dunn et al. (2005) proposed a technique, structural phylogenetics, which uses as its raw
material the structural features of languages, that is, typological characteristics. Using the same maximum parsimony method that is commonly employed in evolutionary biology, similarity trees were produced, from which a partially phylogenetic history could be read. We will return to the different methods and their results below.
Oceanic Languages Among the Oceanic subgroup of Austronesian languages, certain patterns and structures tend to recur over large geographic distances and in different genealogical subgroupings. In this section, we will first give a brief outline of some linguistic features of Proto-Oceanic (based on Lynch et al., 2002: 63–91 unless otherwise noted) and then show how these are reflected in the daughter languages (based on Lynch et al., 2002: 34–53 unless otherwise noted). The next subsection introduces some of the procedures and results of the CM as applied to Oceanic languages, in terms of subgrouping, migration history, and cultural reconstruction.
A Typological Sketch of Oceanic Languages POc had more consonants than most of its daughter languages, and indeed the mergers and losses of consonants are an important factor in subgrouping. POc did have word-final consonants, which have been dropped in most modern Oceanic languages, and most likely had stress on the antepenultimate syllable. The system of pronouns had three persons with an inclusive–exclusive distinction in the first person non-singular (see below), and the numbers for ‘two’ to ‘four’ were cliticized to give dual, trial, and paucal number. The exact forms of subject and object clitics remain unclear, while possessor suffixes are well attested and quite different in form from independent pronouns. POc appears to have had two articles, distinguishing personal nouns and common non-human nouns as in many of the present-day languages, but the function and distribution of the surviving article forms a, na, and ta in the proto-language are not yet resolved. Articles appeared before the noun (although common nouns denoting humans and locative nouns were used bare). There was a small class of adjectival nouns, and a large open class of adjectival verbs. POc had direct and indirect possession, as do the modern languages (see below). Verb roots typically had more than one syllable, and lacked an inherent transitivity distinction. There were two classes: A-verbs and O-verbs, that is, verbs whose subject remains the subject whether it is used transitively or not (as in English ‘I am drinking’; ‘I am drinking water’), and those where the intransitive subject corresponds to the transitive object
121
core studies in northern island melanesia
(as in ‘it broke’; ‘I broke it’). Affixes for transitivization, reciprocal action, and causation were present. In the verb phrase there were proclitics for mood and aspect, and for singular subjects, and enclitics for objects and direction. The following sections briefly outline a number of these features as they appear in some of the daughter languages spoken today, mainly for the Western Oceanic languages of Island Melanesia.
Generally, modern Oceanic languages have a simple syllable structure consisting of a consonant and a vowel (CV), which means that words normally end in a vowel, and that consonant clusters are absent. Stress is fully predictable, falling on the penultimate syllable of the word.
Pronouns Most Oceanic languages have four kinds of pronoun paradigms: (i) independent or free pronouns, which function like noun phrases in the clause; (ii) possessor suffixes on bound nouns for inalienable possession, or on possessive classifiers marking alienable possession (see below); (iii) subject pronouns which are often clitics, fused with expression of tense, aspect, and mood distinctions; and (iv) postverbal object clitics. All Oceanic languages distinguish first, second, and third person for singular and plural, and in many languages also a third number category, dual (indicating two participants). In the first person non-singular there is an inclusive–exclusive distinction, that is, the form for ‘we’ is different depending on whether the listener is included in the reference. Apart from just a few Oceanic languages of southwest New Britain, Oceanic languages do not distinguish gender in their pronoun systems.
Nouns and Noun Phrases Most Oceanic languages have articles that precede the noun in a noun phrase. The forms of these articles often distinguish singular and plural number for common nouns. A separate article marks proper nouns, i.e., the names of persons and places. Frequently, a distinction between specific (or definite) and non-specific (or indefinite) reference is also made. For example, Bali-Vitu (New Britain) distinguishes proper noun phrase (e), common definite (a), and common indefinite (ta):
tuni tarangini ti person chief REAL.PERF.3
zahata bad ‘The chief’s canoe is broken’ (3) Bali-Vitu (Ross 2002: 367)
Lingei hanga Lingei only
na-to REAL.HAB-CONT
viri-a PREPV-1SG ‘Only Lingei lives with me’
‘Bring us some betelnut’
The languages of the Admiralties and southwest New Britain only have a remnant of the article system in elements that are morphologically fused with the noun root. Attributive modifiers of nouns mostly do not belong to a separate adjective class, but are nominalized forms of stative intransitive verbs, or take the form of a relative clause, both following the noun, for example in Mengen: (4) Mengen (Rath 1986: 34) Ng-o-kel-e gie ae iau IRR-2SG-see-OBJ pig of 1SG
monge sleep
[e-bollau] 3SG-big
[e-pe]. 3SG-good
‘Look at my nice big pig.’
Possessive Constructions The Oceanic languages of Island Melanesia all make a distinction between inalienable possession, typically used with nouns denoting body parts and(/or) kin relations, and alienable possession which is used in all other cases. Inalienable possession is expressed in a construction known as direct possession, where a suffix marking the person and number of the possessor is attached directly to the noun denoting the possessed item. For alienable possession, the indirect construction, which makes use of different classifiers to which the possessor suffix is attached, is used. In most Island Melanesian Oceanic languages such classifiers make a distinction between items for consumption (or associated concepts) and other items (the default category). These distinctions can be illustrated by Mangseng: Mangseng (Milligan 1992: 7) (5) Meni-k. arm-1SG ‘my arm (body-part).’ (6) A-k FOOD-1SG
(1) Bali-Vitu (Ross 2002: 378)
122
A vaga kina ART canoe POSS
Vahi ta pazi mai viri hami get ART.INDEF betelnut come PREPV 1PL.EXCL
Phonology
E ART
(2) Bali-Vitu (Ross 2002: 369)
meni. arm
‘my foreleg (of pig) that I am eating.’ (7) Le-k GENERAL-1SG
meni. arm
‘my arm (that I am carving).’
The Languages of Island Melanesia
Another indirect construction for alienable possession makes use of a preposition between the possessed item and the possessor, as in Nakanai and Mengen: (8) Mengen (Rath 1986: 73) Malo loincloth
ae POSS.PREP
Leo. Leo
‘Leo’s loincloth.’
Verbal Derivation and Inflection The main derivational processes in Oceanic languages involve a causative prefix, a reflex of Proto-Oceanic *pa(ka)-, and a reciprocal prefix, reconstructed as *paRi-.1 Here are some examples from Meramera: (9) Meramera (Ohtsuba 1996: 27) Eau na ma-ite oo 1SG(FREE) IRR CAUS-see 2SG buo big
a ART
in-ade NOM-say
sanii. some
‘I will show you some big words (laws).’ (10) Meramera (Ohtsuba 1996: 26) Su mai-sinoa. 3DU RECIP-be.angry ‘They two quarreled with each other.’
A number of languages can add more arguments (valency) to intransitive verbs by forms that reflect POc *-i when the object is patient, or POc *-aki(ni) when the object is a location, goal or instrument, illustrated by Nakanai and Tolai. (11) Tolai (Mosel 1984: 147) Dia they
momo-e drink-TRANS
ra ART
tava. water
‘They drank the water.’ (12) Nakanai (Johnston 1980: 39) E ART
Baba Baba
gilo-a swear-TRANS
e ART
Bubu. Bubu
‘Baba swore at Bubu.’
Verb Phrase and Clause Structure The basic verb phrase found in Oceanic languages, also reconstructed for Proto-Oceanic (Lynch et al., 2002: 83), consists of a subject proclitic indicating aspect and/or mood and subject person and number, followed by one or more verbs and (pro)nominal object (if appropriate) and optional oblique constituents. Most languages make a distinction between realis and irrealis by means of a preverbal morpheme, often intricately fused with the subject proclitic. In some cases a number of other tense, aspect, and mood categories, such as perfective, habitual, and
sequential, are marked this way, as for example in BaliVitu (Ross, 2002: 375). The subject–verb–object (SVO) word order, normally concomitant with prepositions, is geographically the most widely distributed pattern, while SOV with postpositions, attested in Oceanic languages of the Papuan Tip Linkage and along the north coast of New Guinea, is due to contact with local Papuan languages (Lynch et al., 2002: 49–50). The original word order in Proto-Oceanic was most likely verb-initial (Ross, 1988: 385; Thurston, 1994: 587; Lynch et al., 2002: 86), still attested in the languages descended from that of early migrants out of the Oceanic homeland in the Southeast Solomonic group and some languages of remote Oceania (but not in Western Oceanic; see below on contact phenomena in syntax). The position of the negative morpheme is not uniform. Most Oceanic languages have an initial or preverbal negative (ad)verb that appears to reflect *ta or *tikai, also present in Austronesian languages in the western extremity of New Guinea. There are also a number of Western Oceanic languages with clause-final negative morphemes, some of which reflect a proto-form *bwa(li), which also traces back to the area around the Cenderawasih Bay (Reesink, 2002).
Reconstructing Oceanic Language History Using the Comparative Method As set out above, the CM proceeds by establishing cognates and reconstructing sound changes and protoforms of words and morphological material. A tree structure emerges, where the branchings are defined by shared innovation (as in biology); typically a bundle of innovations. Proto-Oceanic (POc) is thus defined by a number of innovations relative to its precursor ProtoMalayo-Polynesian (PMP). For example, in intermediate stages of the proto-language, PMP *p merged with *b, and *k with *g, so that *p and *k remained. They then split again, through other linguistic processes, with the result that POc also had all of *k, *g, *p, and *g but with a different distribution in the lexicon from the original PMP stops. Similar processes affected numerous other consonants, and the PMP diphthongs and vowels also underwent various changes (Lynch et al., 2002: 63–67). The same line of reasoning applies to subgroupings at all levels in the family tree. The relevance of the linguistic tree to population history is of course that the branchings are taken to mirror splits in populations, at least to some degree.
Population Spread and Linguistic Tree Structures—Families As the speakers of Proto-Oceanic spread across Island Melanesia and on into the Pacific, populations split off
123
core studies in northern island melanesia
and their contact with other speakers of Oceanic languages decreased. With time, the languages in different communities evolved differently, and eventually became separate languages, in turn separating into dialects that have sometimes become new languages. Daughter languages by definition retain inherited innovations and in normal circumstances lack innovations that have happened in other branches of the family, after their parent language split off. The geographic area that forms the center of a dispersal is identified by genealogical diversity, that is, the area with the largest number of linguistic subgroups is thought to be the homeland of the family (Pawley and Ross, 1995: 57). This logic can be applied recursively within a language family, as subgroups split into further subgroups and so forth. Based on this reasoning, and of course the relative age of the subgroups, the homeland of the Austronesian family as a whole is placed in Taiwan. The dispersal centers directly leading to the Oceanic subgroup are first at or near Cenderawasih Bay in western New Guinea (Ross et al., 2003: 21); then probably on or near the north coast of New Guinea; then north New Britain, where the ancestral language reconstructed as Proto-Oceanic formed. From there, the Admiralties branch split off (and possibly a St Matthias/Mussau branch; Ross et al., 2003: 26), and there were migrations further southeast into the Solomons and on into Remote Oceania, north into New Ireland, and west along the north and south coasts of New Guinea. Linguistic homelands are identifiable when a population stays in a place long enough for their language to accumulate innovations. As some members of this linguistic population then move on to new areas, the innovations are passed on to the daughter languages. This process of internal development followed by fissioning allows for a tree structure representation of the resulting linguistic relationships.
Population Contact and Linguistic Network Structures—Linkages Not all languages can be assigned to single nodes in a tree. That is, not all language evolution proceeds by neat splits with subsequent lack of contact between populations. Pawley and Ross (1995: 50) spoke of two types of subgroups: innovation-defined versus innovation-linked. Innovation-defined subgroups are those subgroups that can be defined by a unique set of innovations, thus identifying this subgroup as distinct from all others, as was just described. These are the clear cases of language families, which can be described by tree representations. They contrast with what is known as innovation-linked subgroups, where not all innovations are shared by all members of the group. The innovations do not identify further
124
subgroups but are distributed in various combinations across the languages of the subgroup. This is because, in language, innovations do not just spread vertically, by inheritance, but also laterally, by contact, and the result is that the related languages 1, 2, and 3 may share innovations A, B, and C in such a way that 1 has AB, 2 has AC and 3 has BC. This situation is taken to reflect a shared history of those languages as dialects in a chain or network, each accumulating innovations in the normal fashion, but also interacting and acquiring some innovations that originated with their neighbors. This sort of distribution of features leads to the positing of proto-linkages, which are more network-like in structure, rather than proto-languages. Innovation-linked subgroups cause problems for the CM since these kinds of relations cannot be accurately represented in a family tree, and a single proto-language for the linkage cannot be reconstructed (Lynch et al., 2002: 92–93). Three primary subgroups of Oceanic languages were proposed by Lynch et al. (2002: 94–120): Admiralties, Western Oceanic, and Central-Eastern Oceanic. Only the first is innovation-defined. Western Oceanic is innovation-linked, while the status for Central-Eastern Oceanic is as yet unclear. Admiralties is a small family with little internal structure. Within Western Oceanic, the higher-order groups are all linkages: Meso-Melanesian, Papuan Tip, and North New Guinea. Western Oceanic therefore does not present a neat phylogenetic tree structure in which nodes are defined by unique innovations. Within CentralEastern Oceanic, Southeast Solomonic and Micronesian are families, while Southern Oceanic and Central Pacific are linkages. Utupua and Vanikoro are a small but very diverse group of languages bearing unclear relations to the other families and linkages of the region.
Multiple Migrations The distribution of Oceanic languages belonging to different subgroups across the Melanesian islands suggests two major spread events (Lynch et al., 2002: 97–98). Some early populations of POc speakers moved away from their New Britain homeland, moving to the Admiralties, forming the Admiralties subgroup; to other areas in New Britain and the north coast of New Guinea, northeast into New Ireland, and southwest into Bougainville, through the Solomon Islands and on into Remote Oceania, leaving in their wake languages which became the Central-Eastern Oceanic group. This migration wave is correlated with the salient Lapita imprint in the archeological record (see chapter 2). Back in New Britain, the speech varieties of the people who did not leave diversified and changed in the usual fashion, and the same area became the dispersal
The Languages of Island Melanesia
Cultural Reconstruction The CM can also provide a window on the culture, modes of subsistence, social structure, and habitat of speakers long gone. In the process of reconstructing the protoforms of lexical items, linguists also attempt to reconstruct the original meanings. By thematic reconstruction of various areas of the POc lexicon relating to material culture, Ross et al. (1998, 2003) have been able to show continuities and discontinuities in culture as speakers move across the territory. For instance, house terminology supports archeological observations of rectangular houses on piles in the western part of the region, but houses on the ground in the Solomons and Remote Oceania, as well as the change from pile houses to houses on the ground in the Bismarcks. So far, terms have been reconstructible for shell armlets and earrings, combs; grating and graters, axes and knives, fishing and hunting implements; weaving,
plaiting, sewing, making string, baskets and mats; for a variety of cultivation-related activities and crops; for many aspects of food preparation; for canoes, canoe parts, seafaring and navigation, and for many acts of cutting, hitting, piercing and so on. Linguistic reconstruction can access items and aspects of culturally transmitted behavior that are not preserved archeologically, such as birdlime to catch birds, and sharpened sticks and snares for hunting. The absence of reconstructible terms is harder to interpret; for example, the reconstruction effort suggests discontinuity of some practices attested in gardening, such as mulching and digging-in of decomposed vegetable matter and irrigation (Osmond, 1998: 142). That is, linguistics cannot show that such practices were handed down in an unbroken line of transmission. They may have been, but if so the terminology was not preserved with the practices. Within some language families, reconstruction has been able to shed light on the issue of the homeland. Thus, in Indo-European, the reconstruction of items such as ‘salmon’ (*laks-) and ‘birch’ (*bher g-) and ‘beech tree’ (*bha-go) in the proto-language have been taken to indicate an original location containing these species (see, e.g., Trask, 1996: 355; Watkins, 2000). As for the Oceanic languages, however, not much is clear beyond the fact that speakers of POc lived near the coast, had canoes and did a lot of fishing. The location of that coast is not apparent. e
center of another wave, overlaying the first and reaching as far north as New Ireland and the north New Guinea coast (but not the Admiralties and St Matthias/Mussau), west along New Britain, Huon, Milne Bay and part of the south coast of New Guinea, and southeast halfway through the Solomon Islands. This second wave resulted in the Western Oceanic group that we see today. In the Solomons, the Western Oceanic languages form rather a sharp boundary with the Central-Eastern Oceanic languages, which were in situ as a result of the first wave. The boundary can be seen in various linguistic differences between Northwest Solomonic (part of Western Oceanic) and Southeast Solomonic (part of Central-Eastern Oceanic). According to Ross (1988: 382–385), the Southeast Solomonic languages are more conservative in lexicon and phonology, and one hypothesis is that there were no (or few) speakers of Papuan languages in that territory. They share no morphological innovations with Northwest Solomonic and no phonological innovations (except the merger of *dr and *d which is common throughout the wider area and has no diagnostic value). Intriguingly, however, there are two morpho-syntactic features that both subgroups share: VSO word order in the clause, and the use of former possessive noun phrase structures in the verb phrase. Ross’s interpretation is that it is likely that the first wave left settlers in the northwest of the Solomons, whom the next wave encountered, and that these shared phenomena are a result of that contact. While word order is relatively easily diffused, the morpho-syntactic reassignment of particular types of noun phrases would indicate more intense contact (see below) and it is curious that other linguistic features should not have been affected.
Papuan Languages Compared to the Oceanic languages that share these islands, the Papuan languages are extremely diverse. With the exception of the four languages of southern Bougainville (Nasioi, Motuna, Nagovisi, and Buin), possibly the Reefs–Santa Cruz languages, and possibly some of the languages of New Britain (Aneˆm and Ata, and the Baining languages) these languages do not form family groups. Indeed, there appears to be little besides geography to link all or most of the Island Melanesian Papuan languages into a group. As outliers, they tend to be excluded from descriptions of the main body of New Guinea Papuan languages (e.g., Foley, 1986, 2000). Yet they afford interesting insights into questions of typological similarities between presumably unrelated or only distantly related languages. In the following two sections we will see first a description of some of the linguistic characteristics of the Island Melanesian Papuan languages, followed by a summary of proposals that have been made as to their relations to one another, and a presentation of a computational method for investigating such relations.
125
core studies in northern island melanesia
A Typological Sketch of Papuan Languages Dunn et al. (2002) provided a typological survey of Island Melanesian Papuan languages, showing a series of recurrent typological similarities between these languages, which for the most part did not correspond to formal similarities. The next sections give a brief linguistic comparison of the Island Melanesian Papuan languages, both with each other and with the mainland New Guinea Papuan languages, according to the categories of Foley (2000).
Phonology With the famous exception of Yélıˆ Dnye (Henderson, 1995; Levinson, in prep.; Maddieson and Levinson, in prep.), most New Guinea languages have simple phonemic inventories (Foley, 2000: 367), as do most of the Island Melanesian Papuan languages. Rotokas in northern Bougainville was thought to have only 11 phonemes (Firchow and Firchow, 1969), and forms the opposite end of the extreme to Yélıˆ Dnye. However, later work has shown Rotokas to have more phonemes, given the phonemic status of length in the five vowels (Robinson, 2006). Syllable patterns vary widely in New Guinea as in Island Melanesia; but most of the Papuan languages of Island Melanesia permit syllables ending in a consonant, in contrast to the common Oceanic syllable pattern of consonant–vowel (CV). Many mainland New Guinea languages, especially along the north coast, have tone, which does not occur in the islands. However, while most Oceanic languages have fixed word stress (Lynch et al., 2002: 35), in most of the Papuan languages the position of stress has to be learnt for each word.
Morphology Foley (2000) discussed a variety of morphological types in Papuan languages of New Guinea. The Papuan languages of Island Melanesia are also hard to characterize in terms of morphological type. Some of the Bougainville languages verge on the polysynthetic, Yélıˆ Dnye on Rossel Island is very complex with much irregularity, while the Solomon Islands languages have less elaborate morphological structures, and the New Britain languages less still. Lavukaleve is an exception within the Solomon Islands in this respect, having, like Kuot of New Ireland, Kol of New Britain and Yélıˆ Dnye, many forms combining several grammatical functions in a single morpheme, and many grammatically conditioned stem-changes. Kuot for instance has rather complex nominal morphology including eleven declensions, recognizable by their singular forms, which are relevant for number marking. The singular forms identifying each of the declensions end in -ma, -na, -bun, -bu, uom, -bam, -n m, -nim, -n, -m, and ‘other’, respectively (Lindström, 2002: 149). Bougainville languages are e
126
completely suffixing; the other areas have prefixes, in some instances in addition to suffixes. Nouns in mainland New Guinea are morphologically simpler than verbs (Foley, 2000: 371), as in the Papuan languages of Island Melanesia. Gender is a category of most of the islands Papuan languages (Terrill, 2002), as in the north coast non-Trans New Guinea languages, and some parts of the Trans New Guinea family (Foley, 2000: 371). Nominal morphology is largely taken up with marking gender and number distinctions. In many families, gender is part of a more complex noun classification system, and this holds true for about half of the island languages (Terrill, 2002). An example of gender agreement across a clause in Lavukaleve is the following, where the feminine category on the noun is reflected on all other constituents in the clause (‘F’ in the gloss): (13) Lavukaleve (Terrill field notes) airaol two.women
le'laol two.F
ruia-ol old.woman-DU.F
feol 3DU.F.FOC
lei-aol exist-DU.F ‘two old women live there’
Syntax Syntactically, Foley characterized most of the Papuan languages of New Guinea as having OV rather than VO word order. He mentioned Kuot (New Ireland) and Sulka (New Britain) as exceptions to the general trend (Foley, 2000: 383). Indeed, among the Papuan languages of Island Melanesia, the languages of the Bismarcks and Bilua of Vella Lavella form the exceptions, while the languages of Bougainville and the three remaining Solomons languages are all OV. Foley mentions various clause-combining and discourse-combining strategies, including serial verb constructions, light verb constructions, and clause chaining with or without switch reference. Many of the Papuan languages of Island Melanesia have recourse to some of these strategies as well (see examples below). Some of the Bougainville languages have clause chaining, a feature compatible with verb-final clause order and hence absent in the Bismarcks languages. However, at least some of the Bismarcks languages do have serial verb constructions and light verb constructions. One of the Solomons languages, Lavukaleve, uses all of these clause-combining strategies (except switch reference, which is only found in Nasioi of the South Bougainville family). Touo (or Baniata), spoken in the Solomon Islands, has productive serial verb constructions of two types: nuclear juncture and core juncture. In both, tense, mood, and negation are marked only once and have scope over the whole serial verb construction, whereas aspect can be
The Languages of Island Melanesia
selected by each individual verb. Nuclear serial verb constructions share all arguments; core serial verb constructions can have subject–subject, object–subject, or ambient argument sharing. Peripheral arguments belong to the whole serial verb construction in both types (Frahm, 1998):
attention of linguists interested in higher-level linguistic groupings. The geographic distribution of the Papuan languages across island clusters and their geographic separation from the mainland invites speculation about their genetic unity and internal subgrouping.
(14) Touo (Baniata) (Frahm 1998: 65) Zo 3SG.M
horu-a play-REAL
tu-a. stay-REAL
‘He is/was playing.’ (15) Touo (Baniata) (Frahm 1998: 74) Zo ba høre bavi-v-a voiz-e 3SG.M FUT canoe paddle-3SG.F-REAL return-IRR vea. FUT ‘He will return paddling his canoe.’
In Kuot (New Ireland), the serial verb constructions are somewhat less productive, occurring only with verbs for ‘come’ and ‘go’, for example: (16) Kuot (Lindström field notes) u-la 3SG.M.SUBJ-go
a-ko-oŋ 3SG.M.OBJ-throw-3SG.M.SUBJ
ubian ma fishnet(M) e
‘he went (and) threw the fishnet’
The following example shows a clause chain in Lavukaleve. It consists of the medial verb foari “send down” with the non-finite suffix -re, in construction with a final predicate hi lome “she did it,” which carries the subject and tense marking for the whole chain. (17) Lavukaleve (Terrill field notes) Hovir cough
ga SG.N.ART
e-foa-ri-re 3SG.N.OBJ-go.down-CAUS-NF
hi do/say
lo-me. 3SG.SUBJ-HAB
‘She spits her cough in. [lit. sending down her cough she did it]’
The main predicate hi lome is itself complex, consisting of a light verb construction with verb hi “do” and personmarking and aspectual material expressed on the auxiliary me “habitual.”
The History of Papuan Languages Despite the considerable internal diversity of the Papuan languages of Island Melanesia, there have been repeated attempts to group them together (see also chapter 3). Ever since the genealogical status of the Austronesian family was established, the existence of a remnant set of languages not belonging to this group has attracted the
Previous Proposals For most of the Papuan languages of Island Melanesia the CM has never been applied, since no sets of adequate cognate candidates can be identified. As noted, in the 1970s there were two bold proposals linking the Papuan languages of Island Melanesia together within much larger groupings. These are Greenberg’s “Indo-Pacific Hypothesis” and Wurm’s “East Papuan Phylum”. Greenberg’s proposed macrofamily (Greenberg 1971) linked most of the non-Austronesian languages from the Andaman Islands to Tasmania (excluding mainland Australia). He posited three groups of languages within Island Melanesia: a New Britain group, a Bougainville group, and a Central Solomons–Santa Cruz group. Greenberg’s methodology in this paper involved comparing all the available lexical data and coming up with a rather subjective assessment of similarity first of pairs of lexical items, and then of pairs of languages and ultimately groups of languages. He attempted to identify and remove Austronesian loans in some places, although this was not consistently done (see, e.g., item 23 in the Central Solomons list, which was identified as appearing in Austronesian languages, but which was nonetheless included as a cognate set both for the Central Solomons group and for the much larger Indo-Pacific group). While the paper did not use his better-known Multilateral (or Mass) Comparison method (e.g., Greenberg, 1987), which has been comprehensively discredited by linguists (e.g., Campbell, 2003), it nonetheless had serious methodological flaws. Besides failing to deal adequately with loans, he performed single-instance pairwise comparisons on a relatively small number of lexemes, and thereby ignored the critical constraint of the CM that regular sound correspondences be demonstrated for related languages. That is, Greenberg’s “cognate sets” are more properly unproven cognate candidate sets. Moreover, no attention was given to the fact that furtherflung languages are likely to be more distantly related to each other than nearby languages, so that the putative cognates of far-flung languages would be expected to be less similar than the putative cognates of nearby languages. In other words, many or most of the putative cognates set up by Greenberg are simply too similar to each other to be descended from a distant ancestor. Wurm’s (1975a, 1982) East Papuan Phylum was based on structural and lexical similarities. It encompassed all the Papuan languages of Island Melanesia in a single group (including the geographically distant
127
core studies in northern island melanesia
Yélıˆ Dnye on Rossel Island in the Louisiade archipelago), divided into four stocks: Yele–Solomons, New Britain (including Kuot on New Ireland), East Bougainville, West Bougainville, and the Reefs–Santa Cruz sub-phylum level family. The lexical and structural similarities underlying the groupings were identified and evaluated subjectively. While Wurm based his grouping on more data than were available to Greenberg, his methodology and results have not found general acceptance among specialists. However, the structural similarities (e.g., pervasive gender systems, elaborate verb morphology, medial verbs, etc.) form a sufficiently unusual typological cluster to suggest that the pursuit of genealogical relationships among these languages is worthwhile. Allen and Hurd (1965) produced a lexicostatistical classification (c.f. chapter 3) of the languages of Bougainville, dividing the Papuan languages into two groups, Konua–Keriaka–Eivo–Rotokas (the latter including the dialect Aita) and Nasioi–Nagovisi–Buin–Motuna (or Siwai). Both Greenberg (1971) and Wurm (1975a, 1982) referred to this classification. Ross (2001b) used pronoun paradigms as a heuristic to propose subgroupings of the Papuan languages of Island Melanesia. A pronoun paradigm is a morphologically structured set of grammatical function words indicating the person and usually number of speech-act participants. There is a considerable literature on the phylogenetic stability of pronoun paradigms (see for example the debate in Nichols and Peterson, 1996, 1998; Campbell, 1997). There is a marked statistical tendency for pronoun paradigms to be stable, but notable exceptions do exist, both of the introduction of exogenous lexical forms into a paradigm, and of the replacement of entire paradigms. With knowledge of the phonological systems of the languages being compared, it is possible to assess the probability that the observed similarities are the products of chance (Nichols, 1996). Pronouns are particularly useful, because as members of a closed and internally structured domain, entire paradigms can be compared across languages with much lower probabilities of chance similarity than for individual words. Ross (2001b) demonstrated some likely phylogenetic patterns in the forms of the pronoun paradigms in the Papuan languages of Island Melanesia, finding eight possible groups; Yele–West New Britain, Kol, Sulka, East New Britain, Kuot, North Bougainville, South Bougainville, and Central Solomons. Reefs–Santa Cruz was not considered due to Austronesian admixture in the pronoun paradigms. It is likely that the CM can be applied to some small groups in this area, including Aneˆm–Ata, the Baining languages, North Bougainville, and South Bougainville, but this remains to be demonstrated (beyond Ross’s pronoun work). What is apparent is that, apart from these
128
lower-level groups, the Papuan languages of Island Melanesia must have been separated for a very long time, since the CM has so far been unable to demonstrate any genealogical links. In this context, Ross’s (2001b: 311) suggestion that any genealogical relationships among Island Melanesian Papuan languages must be considerably more ancient than those among the Trans New Guinea Papuan languages seems very reasonable.
Structural Phylogeny The puzzle presented by these languages is their recurrent structural similarities together with the absence of formal correspondences, as noted by Capell (1962), Wurm (1975a, 1982), and all subsequent scholars. That is, a large proportion of the languages may have a gender distinction, but the actual forms that express the categories (such as masculine and feminine) cannot be related to each other. The structural similarities suggest some sort of relationship, but the lack of correspondences in form means that the linguist’s main tool in establishing historical relationships between languages, the CM, cannot be applied. This leaves a methodological gap that the studies above have all attempted to fill, with varying degrees of success. Dunn et al. (2005) addressed the recurrent structural features among the Island Melanesian Papuan languages by applying a computational cladistic analysis to a sample of features of linguistic structure. A typological database was constructed, comprising a set of 125 binary values, each encoding the presence or absence of a structural feature of language (i.e., independent of actual surface forms; see Dunn et al., 2005 and supporting online material for details; a list of languages with their affiliations is provided in Appendix II). Using parsimony analysis, a phylogenetic tree was produced, showing the most likely genealogical relationships between the languages. The method returned a tree with a coherent, geographically organized, structure, showing three well-supported main groupings corresponding to the Bougainville, Bismarcks, and Central Solomons archipelagos (figure 8.2; Dunn et al., 2005: 2074, Figure 4). Some branchings have weak support, and the attachment of Lavukaleve (Solomons) is particularly uncertain. The interpretation of these results is not entirely straightforward. That the method does have the power to show phylogenetic relations in the absence of lexical data was demonstrated in the same paper on a set of Oceanic languages in the same region. The Oceanic tree was a very close reproduction of the independently known phylogeny produced by the CM. Therefore, it is likely that the lower-level clades in the Papuan tree too reflect common descent, and quite possible that the higher-level clades with a better degree of support do so as well. If so, the
The Languages of Island Melanesia
Figure 8.2 A cladistic tree of Papuan languages of Island Melanesia generated using the parsimony method on a matrix of 125 abstract structural features (after Dunn et al 2005, Figure 4).
clustering per archipelago is due to in situ diversification from an ancestral language for each of the well-supported clades. Another possibility is contact (see further below), whereby prolonged contact between populations sharing a geographic territory causes a degree of structural affinity of the descendent languages. A curious feature of the Papuan tree in figure 8.2 is that although the languages appear in archipelago groups, the groups themselves do not appear in clear geographic sequence. Instead, the Central Solomons languages are found in between the Bismarcks and Bougainville, a fact to which we shall return in the Discussion. The phylogenetic tree method requires all languages in the sample to attach to the tree, but let it be made quite clear that we are not in a position to suggest a single origin for all the Papuan languages of Island Melanesia.
Contact between Papuan and Oceanic Languages in Island Melanesia Before the arrival of the Oceanic languages, Papuan speakers had the ground to themselves, although population densities must have been quite low for most of that time (e.g., Leavesley, 2006). Judging by the diversity of present-day Papuan languages of Island Melanesia, it may be assumed that there was quite a lot of linguistic diversity also in pre-Oceanic days. In the millennia following the arrival of speakers of Oceanic languages, many Papuan languages will have disappeared, and with them presumably a number of structures which may have influenced Oceanic languages. The (descendants of the) remaining languages cannot be presumed to be fully representative of the variation that the speakers of ProtoOceanic encountered.
As we have seen, there are structural similarities in many areas of the grammars of Island Melanesian Papuan languages, presumably resulting in part from shared origins of some groups of languages, and in part from contact between the small and scattered populations over the many millennia of human occupation prior to the Oceanic era in Island Melanesia. After speakers of Proto-Oceanic arrived in Island Melanesia, it is clear that there was contact with linguistic consequences also between the newcomers and the original inhabitants. There has been exchange of linguistic items in virtually all aspects of grammar, and in some cases linguists have even debated whether a language was basically Oceanic or Papuan (Lynch et al., 2002: 15–16), although for the vast majority of languages their Oceanic or Papuan identity is evident. Much linguistic work has focused on the interchange between Papuan and Oceanic languages spoken in close vicinity (Lincoln, 1978; Wurm, 1978; Ross, 1996, 1999, 2001b; Thurston, 1994). Indeed, this is an area famous for linguistic contact phenomena. Similarities in languages, whether they involve constructions, words, morphemes, or phonemes, result from four sources: (i) shared inheritance, (ii) contact, (iii) natural tendencies given the design-space of human language, and (iv) chance. All Oceanic languages trace their descent to a common ancestor, presenting a clear case of similarity by shared inheritance. As for similarities between unrelated languages, bilingual speakers may have re-arranged the morpho-syntactic constructions in their first language on the model of a neighboring language. Ross (1996, 2001b) coined the term “metatypy” for this phenomenon. His work has shown that Takia and Maisin, which belong to different Oceanic subgroups and are located some 400 miles apart in New Guinea, have superficially similar innovations, but that these were acquired independently through contact with different Papuan languages of the Trans New Guinea family. Thomason and Kaufman (1988) and Thomason (2001) have shown that almost any structural or grammatical feature can be transferred from one language to another. Case studies show that the intensity and type of contact typically determines the scale and kind of borrowing, roughly along the following lines (after Thomason, 2001: 70): 1. Casual contact without necessary bilingualism— non-basic lexical borrowing only, especially nouns; 2. Less casual contact, with a core of bilinguals— continued non-basic lexical borrowing, some functions words, especially conjunctions; 3. More intense contact, with more bilinguals and a high status of the source language—much lexical
129
core studies in northern island melanesia
borrowing of all categories, including basic vocabulary and pronouns, even derivational affixes, and structural borrowing may now include wordorder innovations, syntax of subordination, and inflectional affixes, and the phoneme inventory may be affected; 4. Very intense contact with extensive bilingualism and strong social forces—continued lexical borrowing in all categories, with sweeping changes in syntax, resulting in major typological changes, changes in type of morphology (e.g., flexional to agglutinative), and changes in phonemes even in native words. In the case of language shift, when a speech community adopts a different language, the new language will normally first undergo phonological change. Ross (1994) ascribed this process to two Oceanic languages of New Ireland, Lamasong and Madak, which share certain phonological features with neighboring Papuan Kuot. Language shift can also involve transference of some morpho-syntactic constructions from the first language into the adopted language. Evidence for such substrate phenomena is available in the Oceanic languages of West New Britain. Thurston (1994: 594) and Goulden (1996: 131) recorded a locative postposition for the Bariai languages. Since postpositions are highly unusual in languages with verb-initial word order, this appears likely to be a remnant of an earlier Papuan language spoken by a population, who then shifted to an Oceanic language. The only Papuan language left in that region, Aneˆm, has been heavily influenced by surrounding Oceanic languages. Indeed, Thurston (1994: 603) concluded that the structural simplification observable in Oceanic languages of northwestern New Britain, normally associated with creoles, looks “rather more like linguistic renovation resulting from language shift than piecemeal innovation.” In this section we first examine cases of linguistic borrowing or suspected borrowing in various subsystems of language, and then present a computational approach for investigating the impact of language contact.
Cases of Contact-Induced Language Change in Island Melanesia Below are reviewed a number of suspected contact phenomena in Oceanic and Papuan languages that involve various types of borrowing associated with more intense contact situations, with or without rampant lexical borrowing.
Phonology There are some documented cases of contact-induced phonological change. For example, the sole Papuan
130
language of New Ireland, Kuot, acquired f and s from its Oceanic neighbors (Lindström, 2002: 97–99), by borrowing words containing them. Lavukaleve (Solomon Islands) too has a phoneme s apparently of relatively recent origin, but it is not clear whether its presence in Lavukaleve is due to Oceanic borrowing, or to processes internal to the language (Terrill, 2003: 377–378). In the Solomon Islands, the phoneme z has an areal distribution, being found in many of the languages of the Western Solomons, both Papuan and Oceanic. The phoneme z in Oceanic languages derives from *j and/or *y (Ross, 1988: 221). Allophony and distribution of z in the two Papuan languages of the Western Solomons, Bilua and Touo (Baniata) is very similar to that in the Oceanic languages and it is clear that contact has been involved, although the direction of influence cannot be conclusively determined.
Lexical Borrowing In general, borrowing of lexical items between the two strata in Island Melanesia has more often occurred from Oceanic languages into Papuan than vice versa. Such borrowing affected not just newer cultural items but also the so-called core vocabulary: body-part terms, geographical features, and kinship terms. There are a few attestations of old Oceanic items in Sulka of East New Britain (Reesink, 2005a: 154) and Lavukaleve of the central Solomon Islands (Terrill, 2003). According to Levinson (2005), Papuan Yélıˆ Dnye on Rossel Island preserves borrowed Proto-Oceanic numerals while its nearest Oceanic neighbors do not. In a number of cases Papuan languages borrowed seafaring or more general maritime vocabulary from their Oceanic neighbors (Thurston, 1982: 57–64; Terrill, 2003: 381), who were the experts on waterborne transport. Oceanic languages have in turn sometimes borrowed vocabulary pertaining to inland flora and fauna, as attested for Oceanic Mangap-Mbula on Umboi Island, borrowing from Papuan Kovai, with which it shares the island (Bugenhagen, 1994: 69). An intriguing possible instance of Papuan–Oceanic borrowing comes from Lavukaleve, which has two words fofo “basin” and koko “drum.” These words both also appear in Oceanic languages of the Southeast Solomonic subgroup, but apparently in no other Oceanic languages. They are reconstructible back to Proto-Southeast Solomonic as *popo “wooden bowl” and *γoγo “slitgong” (Osmond and Ross, 1998: 73, 110) but no further. The question then arises of how these two words came into Proto-SES. A speculative hypothesis could suggest that their origin lies in Lavukaleve (Terrill, 2003: 383), or now extinct relatives of Lavukaleve. It may be that two other words, mon “dugout canoe,” which appears as mola in the Papuan languages of the Solomon Islands, and (gu)rat “basket,” found in both Papuan languages of East New Britain and languages of
The Languages of Island Melanesia
the St. George linkage of Meso-Melanesian, also originate in Papuan languages of this region (Reesink, 2005a: 155), since no reconstruction of these is available for POc.
Pronominal Systems The inclusive–exclusive distinction in non-singular firstperson pronouns is a common feature of Oceanic languages, but is also found in about half of the Papuan languages in Island Melanesia (Dunn et al., 2002: 48). On New Britain, Aneˆm (Thurston, 1994: 594) and Ata (Hashimoto, n.d.) only make the distinction in object suffixes, which also function as possessives, but not in subject prefixes. While Sulka and the Baining languages completely lack this distinction, Papuan Kol appears to have borrowed the free inclusive pronoun mang wholesale from neighboring Mengen (Reesink, 2005a: 169). In Kuot (New Ireland) the distinction is completely integrated in all pronominal paradigms in both plural and dual number (c.f. Lindström, 2002: 213). Many of the Papuan languages of Island Melanesia have a dual number distinction in their pronouns. For some of the languages, the source of the dual form is clear, but this is not always the case. Many Oceanic languages have the dual distinction too, but it is not known whether it was a part of Proto-Oceanic, or whether it was separately innovated in a number of Oceanic subgroups after the split-up of POc (Ross, 1988: 97–98, 100–101); see discussion in Dunn et al. (2002: 45–48).
Noun Class and Gender Systems of nominal classification or gender are quite typical of the Papuan languages throughout Island Melanesia (Terrill, 2002), and along the north coast of New Guinea, mainly in non-Trans New Guinea groupings. In noun class and gender systems, each noun belongs to one of the available categories, and the categories require agreement markers on words associated with the noun, within the noun phrase and/or in the clause (c.f. the example from Lavukaleve above). Sulka and Yélıˆ Dnye are the only Papuan languages in Island Melanesia that do not have this feature (Reesink, 2005a: 168). Oceanic languages prototypically do not have gender or noun classes, but it has been adopted by some Oceanic speech communities on New Britain, mainly the eastern Arawe languages Bebeli, Akolet, Avau, and Atui, which all distinguish the genders masculine, feminine, and neutral in third-person singular pronouns (Ross, 1988: 183). The use of different articles, basically a for individual common nouns and o~u for mass nouns, suggests some nominal classification in nine languages of New Ireland and another nine North Bougainville languages (Ross, 1988: 251, 293–305). These articles are also found in
Papuan Sulka of New Britain, a indicating singular, and o plural, with an additional e for proper nouns—also present in north Bougainville Halia (Allen, 1987: 14)— suggesting contact between Sulka and the New Ireland– North Solomonic linkage, and constituting one of the arguments to suggest that Sulka originated on New Ireland.
Morpho-Syntactic Constructions Possessive Constructions. As we have seen, POc had a distinction between direct and indirect possession, with at least three classifiers (for “food,” “drink,” and “general”). A much more complex system of possessive classifiers exists in the New Britain Papuan languages Aneˆm, Ata, and Kol. In Kol, the possessive “classifiers” are actually agreement markers with the class of the possessed noun, of which there are nine. Both Aneˆm and Ata, which are almost certainly related, have possessive constructions requiring many more semantic distinctions. Typically, the possessive classifiers in Aneˆm and Ata permit the same noun to be used in different possessive constructions. Compare the following Aneˆm examples with those given above for Mangseng: Aneˆm (Thurston, 1987:51) (18) Dêk- g-a. thing-(BODY PART)-1SG ‘my (body part) thing.’ (19) Dêk-at. thing-EDIBLE.1SG ‘my (edible) thing.’ (20) Dêk-n-ai. thing-NEUTRAL-1SG’ ‘my (neutral) thing.’
Therefore, a reasonable hypothesis is that the possessive classification system posited for POc (which in the NNG linkage is restricted to two, *k(a)- for consumables and *na~le ‘general’) originates from contact with Papuan languages of northern New Britain, with Aneˆm, Ata, and Kol as present-day survivors. Nominalized Attributive Adjectives. For POc, two modifier classes are reconstructed, a small class of adjectival nouns and a large, open class of verbs. The noun-like adjectives occur as attributives—modifiers that express some property of the referent of the head noun within the same noun phrase—in a possessive-like construction, still present in an Oceanic-wide distribution. A similar construction with nominalized verbs is used for members of the large open class only in the Admiralties, and within Western Oceanic only in the North New Guinea and Papuan Tip clusters (Ross, 1998a: 108), while the Northwest Solomonic
131
core studies in northern island melanesia
languages Nehan and Halia are analyzed as “strict adjectival-noun languages, that is, languages in which only adjectival nouns occur as property predicates” (Ross, 1998a: 96). The innovation of nominalized attributive adjectives (by means of a pattern extension) for the large adjective class is restricted to Western Oceanic, Admiralties, North New Guinea, and Papuan Tip. It is accounted for by contact between these Oceanic-speaking groups after the break-up of POc, and possibly facilitated by contact with Papuan languages (Ross, 1998b: 271–272). For example, Papuan Sulka in New Britain has a robust attributivization mechanism whereby every verb can be nominalized to form an attributive modifier, yielding structures identical to those in the strict adjectival noun languages of Northwest Solomonic (Reesink, 2005a): (21) Halia (Allen, 1987: 18; Ross, 1998a: 96) a ART
sirö soup
a ART
hiski hot
Computational Approaches to Investigating Contact-Induced Language Change
‘hot soup’ (a soup, a hot one) (22) Sulka (Reesink, 2005a: 175) A ART
mhël man
a ART
køn-ör short-NOM
‘A short man.’ (a man, a short one)
Syntax As mentioned in the section on Oceanic languages, the reconstructed verb phrase of Proto-Oceanic (Lynch et al., 2002: 83) consists of a subject proclitic indicating aspect and/or mood, and subject person and number, followed by one or more verbs and (pro)nominal object (if appropriate), and optional oblique constituents. Thus, the clause word order was most likely VO, possibly with VSO as the unmarked order (Ross, 1988: 385–386). This order is still found in Oceanic languages in the Southeast Solomonic group and further afield in Polynesia. Curiously, this word order is also found in Papuan Kuot, quite different from its present-day Oceanic neighbors in New Ireland, which are all SVO, and any other Island Melanesian Papuan languages. Papuan languages, of whatever stock, generally have the clausal order SOV, with exceptions in three areas along the northern rim of the greater New Guinea area, which was the pathway of the migrating Austronesian speakers three to four millennia ago: the Bird’s Head, some of the Torricelli languages, the Bismarck Archipelago languages, and Bilua in the Solomon Islands. It seems an unavoidable conclusion that the SVO order is the result of contact with Austronesian languages, although this is less clear for the Torricelli languages than for the other areas (Foley, 2000: 383); but see also Laycock (1973), quoted in Wurm et al. (1975: 952). Conversely, as Ross (1988) and Lynch et al. (2002) have pointed out, there are many 132
Oceanic languages that have adopted the Papuan SOV word order along with postpositions, both in the North New Guinea and Papuan Tip clusters of Oceanic. The canonical Papuan SOV word order is found in the Papuan languages of Bougainville and the three other Papuan languages of the central Solomon Islands, Touo (Baniata), Lavukaleve, and Savosavo. Having the negative adverb in clause-final position also appears to be a Papuan feature, associated with verbfinal word order (cf. Dunn et al., 2002: 37 and Reesink, 2002). It also occurs in Papuan languages with SVO word order, and in a number of Austronesian languages, including Oceanic subgroups of the Admiralties and some of the North New Guinea linkage (Hovdhaugen and Mosel, 1999). The departure from the common pattern of verb-final word order together with clause-final negation is likely to be the result of language contact.
While the above discussion shows that it is possible to uncover recent social contact situations between language groups by investigation of contact-induced language change, more ancient language contact situations, whether Papuan–Oceanic or Papuan–Papuan, can be much harder to discern. Information on past linguistic forms and on past socio-cultural situations is very difficult to recover. The “structural phylogeny” method proposed in Dunn et al. (2005) and outlined above aims to uncover ancient connections between the Papuan languages of Island Melanesia (c.f. figure 8.2). While that paper was equivocal on the topic of whether the “historical connection” revealed is due to genealogical inheritance or derived through contact, the paper leaned towards genealogy as being the most likely origin of the stronger signals of connectedness. Here we test an obvious counter-hypothesis: that the connections identified between languages derive from shared patterns of language contact, i.e., reticulation, rather than phylogeny. This is not the first computational approach towards understanding and detecting language contact. Minett and Wang (2003) tested two methods of identifying areas of borrowing by mathematical means. Warnow et al. (2005) investigated properties of character evolution and parallel development, and both Warnow et al. (2005) and Nakhleh et al. (2005) reported work to develop methods that can handle tree structures and network structures in a single model, building on the work and data of Ringe et al. (2002). However, all of these attempts relied on the linguist to identify the borrowing at the data coding stage. The approach presented here takes a higher-level perspective, comparing patterns of similarity between abstract structural features of the target languages, using
The Languages of Island Melanesia
the same dataset as for the phylogenetic parsimony analysis shown in figure 8.2.
Spatial–Structural Correlation in Papuan and Oceanic Languages The correlation between geographic distance and structural distance for the languages of Island Melanesia is illustrated in figure 8.3. For each pair of languages the geographic separation was calculated from the geometric center of the region where the language is currently spoken, and the structural distance was calculated as the percentage of disagreement between valid pairs of feature values in the Dunn et al. (2005) database. The language pairs were divided into three groups: Oceanic–Oceanic pairs, Papuan–Papuan pairs, and mixed Oceanic–Papuan pairs. For each group, R2 was calculated. Accretion of structural distance would be expected to tail off asymptotically at some geographic separation, representing a background noise level, determined by the specific content of the feature set in the database. A cubic spline was found to fit the data appropriately (shown in figure 8.3). This figure
shows that greater structural distance correlates with greater geographic distance, and, as might be expected, languages close to each other share more structural features than those further apart. The Papuan pairs are generally structurally further apart than the Oceanic pairs, but the correlation between structural and geographic distance is constant. Papuan pairs appear to correlate best with each other within about 600 km. The mixed Papuan–Oceanic pairs also show a slight correlation between spatial and structural distance, but this correlation is very weak (R2 = 0.062) compared to the Oceanic–Oceanic pairs or the Papuan–Papuan pairs. This shows that the structural influence of Papuan and Oceanic languages on each other is not systematic enough to produce an unambiguous signal, and thus supports the hypothesis that the geographic–structural correlation in the Papuan set is independent, and not a result of “copying” the Oceanic cline by borrowing structures from geographically diverse Oceanic neighbors.
Spatial–Structural Correlation by Archipelago To investigate regional patterns of contact, the same geographic–structural correlation was plotted by greater archipelago: “Bismarcks” (New Britain and New Ireland), “Bougainville,” and “Solomons,” as shown in figure 8.4. It is evident from this figure that that the weak signal of contact that is present between Papuan and Oceanic languages is due entirely to languages of the Bismarcks and Solomons, while the languages of Bougainville show no discernible contact effects at this level of analysis. The cline as such gives no clue to the direction of contact, but as discussed above there is clear linguistic evidence, particularly in the Bismarcks, that Papuan languages
Figure 8.3 Pairwise plot of correlations between geographic and structural distances of Island Melanesian languages. Language pairs are plotted separately for Oceanic pairs (circles), Papuan pairs (triangles), and mixed pairs (gray dots). Each subset is fitted by a line of best fit (cubic spline) with confidence intervals marked by dashed lines.
Figure 8.4 Correlation between geographic and structural distance of mixed Papuan– Oceanic pairs of languages (the dots from figure 8.2) by archipelago. Each archipelago group consists of the Papuan languages of that archipelago paired with all the Oceanic languages in the sample.
133
core studies in northern island melanesia
have influenced Oceanic languages, while the reverse direction of influence is also well attested. The fact that the contact signal is two-way provides further support for the hypothesis that the geographic clades formed by the Papuan languages in the parsimony analysis (figure 8.2) are not simply the results of shared Oceanic structures imposed on otherwise diverse Papuan languages.
Discussion and Conclusions We have presented an overview of the characteristics of Oceanic and Papuan languages of Island Melanesia, and shown how linguistics can contribute information not only on the history of languages but also on the history of the speakers of those languages. The Comparative Method as applied to the relatively recently developed Oceanic language family has been able to reconstruct words for items of material and nonmaterial culture, environment, and numerous other aspects of life, which must therefore have been present in the society where the ancestral form of the language was spoken. Traditional linguistic methods have also been able to point to situations of contact and exchange between Papuan- and Oceanic-speaking populations. By reconstructing the history of diversification from the common ancestor, the CM can also suggest interpretations of migrations and population history. In this case, it indicates at least two major waves of Oceanic speakers through Island Melanesia. By applying the notion of most diversity in the place of origin, it has been possible to place the homeland of Oceanic on the north coast of New Britain. Traces of the original homeland of Proto-Oceanic are difficult to observe today, because of the overlay of Western Oceanic languages. Lynch et al. (2002: 97) supported their argument for the location of the POc homeland on the north coast of New Britain with Nichols’ (1992) reasoning about spread zones and residual zones. She observed that in many parts of the world particular areas are the recurrent routes of language dispersal, and language families replace one another in these areas through history, creating low (linguistic) genetic diversity. Other zones see remnant languages accumulating, making for very great diversity in terms of linguistic origins. Lynch et al. (2002: 97–98) made the point that the dispersal center for Western Oceanic having been in this area in itself constitutes an argument for the earlier POc dispersal center also having been in approximately the same place. It is possible that this area was a spread zone in pre-Austronesian times, too, being on the most obvious route from New Guinea into Island Melanesia. The diverse Papuan languages, at least some of which have most likely been in the region for tens of millennia,
134
are less accessible to conventional linguistic methods. This is because the time depth is so great that any such indicators of common descent that the CM requires, and which may have been present in the past, have been obscured by the natural processes of language change that are constantly at work in all human languages. Techniques for systematically exploring structural similarity have therefore been developed, on the assumption that such features are more stable over time than vocabulary. Here we have shown how this technique can be used also to investigate the extent and impact of contact between unrelated languages. The structural phylogenetic method has been shown to be capable of identifying known phylogenies (Dunn et al., 2005), and its results some way beyond the time limit of the CM most likely represent actual phylogenetic relations, but ways of distinguishing contact-induced similarity from true phylogenetic relatedness at the very greatest time depths remain elusive. Although Wurm’s work has not met with general acceptance, some of his hypotheses are worth repeating here. Wurm et al. (1975) said of the first Trans New Guinea expansion in New Guinea Island that it moved “into the south-eastern tail-end of the New Guinea mainland [and] drove out another, earlier language group there which moved on to Rossel Island in the Louisiade Archipelago, and to the New Britain–New Ireland area, perhaps superimposing itself upon even earlier languages there.” (p 942). The few typological features that Wurm et al. mentioned for this older stratum include nominal classification systems or genders, plural marking on nouns, and a prevalence of verb-stem changes according to object and subject marking. Pawley (chapter 3), based on much more careful linguistic work, also suggests that the Trans New Guinea family diversified and spread across large parts of New Guinea island, overlaying languages already spoken in those regions, and estimates the time of this process to around 8,000–10,000 years ago. The relative youth of the Trans New Guinea family as compared to the Island Melanesian languages is suggested also by the fact that Ross (2005) was able to identify a diagnostic set of pronominal forms that allows preliminary identification of Trans New Guinea languages (although he does warn that the classifications are tentative, needing further corroboration from other linguistic evidence). For Island Melanesia, on the other hand, Ross (2005) found no evidence in the pronominal paradigms for a single genealogical ancestor, but recognized five small families and three isolates. This points to a more complex story, and/or a greater depth of time in Island Melanesia than that within the Trans New Guinea family. The most enigmatic result of the “structural phylogenetics” performed on the Papuan languages of Island Melanesia is the fact that the archipelago clusters appear
The Languages of Island Melanesia
out of sequence in relation to the geography. Instead of Bismarcks–Bougainville–Solomons, the Solomons languages can be made to form a clade either with the Bismarcks or the Bougainville languages. A first hypothesis was that the Bismarcks and Solomons languages show greater affinity as a result of greater influence of the Oceanic languages. This would also fit with the fact that Bougainville Papuan languages are in closer proximity to each other than most of the others. The pairwise comparison of Papuan and Oceanic languages by archipelago (figure 8.4) showed that Papuan languages of the Bismarcks and Solomons languages exhibit slight effects of contact with Oceanic languages, while those of Bougainville did not. The contact signal is weak, but it is clear from independent data that it is the result of influence both from Papuan languages on Oceanic languages and vice versa. The notion of contact with Oceanic languages as a major determinant of the structural interrelations among Papuan languages can therefore be excluded (contra Terrell et al., 2001). The distribution of language clusters and the absence of a contact signal for Bougainville calls for an explanation, but at the present stage of research only speculation is possible. At first glance, Bougainville appears to have been a residual zone in Nichols’ (1992) sense, somehow bypassed by the migration route from the Bismarcks to the Solomons. The interior of Bougainville is mountainous and one might imagine it as a stronghold from which the earlier settlers could keep newcomers at bay. However, such a scenario does not take all the facts into account. For example, Bougainville and much of the Solomon Islands were united in a single landmass (“Greater Bougainville”) from the beginning of human settlement until the end of the last glacial maximum at around 8,000 years ago, and there is no immediately apparent reason why there should not have been free movement and contact throughout that area. Further, population densities would not have supported territorialization in the early period, and seafaring techniques would be unlikely to permit travel over distances as long as required to bypass Bougainville altogether (chapter 2). Moreover, Summerhayes (chapter 2) shows that there appears to have been virtually no contact between the Bismarcks and Bougainville (and further south) between the initial settlement of Bougainville (with Buka) at or before 32,000 years ago and around 5,000 years ago (for example, the gray cuscus, Phalanger orientalis, was introduced from New Guinea to New Ireland around 20,000 years ago, but is not found in Bougainville until 5,000 years ago). Perhaps what the tree shows, then, is in fact a weak signal of an ancient unity between Bougainville and the Solomons, versus the Bismarcks, coupled with linguistic “drift” in Bougainville in the period following the geographic separation from the Solomon Islands. Indeed,
the fact that the Papuan languages in Bougainville group into two families that are still recognizable does suggest that they belong to later, local expansions. Another possibility is that the Bougainville languages are intrusive, having arrived in the territory after the other Island Melanesian Papuan languages, but it is hard to reconcile such a scenario with the geography of the area, and the fact that there are no known relations of these two families outside Bougainville. At the current stage of research, linguistics can discern two principal strata in Island Melanesia, the recognizable family of Oceanic languages on the one hand, and the ancient and diverse Papuan languages on the other. Archeologically, however, there is slightly more resolution. Between initial human settlement 40,000 years ago and the arrival of speakers of Oceanic 3,300 years ago, there were the imprints of at least two more waves. At about 20,000 years ago, various species were introduced from New Guinea and there was increased connectivity between communities, shown for instance in trade goods such as obsidian found at vast distances from the source. At 8,000–5,000 years ago, more new species and practices appeared, and there were other changes to life in that the climate warmed up, causing sea levels to rise and changes in vegetation (chapter 2; Leavesley, 2006). If we accept the scenario that there were new arrivals into New Britain and New Ireland around 20,000 years ago and presumably several times after that, and that these did not go on to Bougainville and beyond, that could also account for Bougainville with the Solomons coming out as something of a cluster relative to the Bismarcks. This would appear to be consistent with the distribution of mitochondrial haplogroup M27, which is particularly common in Bougainville and the Solomons, as well as the mitochondrial profile of New Britain where M28 is very dominant compared to New Ireland, Bougainville, and the Solomons (Merriwether et al., 2005). However, it should be noted that the Solomon Islands languages are not much more different from the Bismarcks languages than they are from the Bougainville languages. Perhaps this points to carriers of M28 arriving in New Britain in a steady trickle, being assimilated into the existing populations and their languages, rather than in large groups that could have maintained separate cultures and languages. The position of Yélıˆ Dnye in between the Bougainville and the Solomons clades is possibly consistent with Wurm’s (1982) suggested Yele–Solomons stock, and the ideas of Wurm et al. (1975) that Trans New Guinea languages pushed earlier languages off most of the mainland, resulting in languages of an older stratum being found in the periphery of the area. However, the time depth of the Trans New Guinea family is only around 10,000 years (chapter 3). This is problematic here, as
135
core studies in northern island melanesia
population densities had gone up quite a lot by this time (Leavesley, 2006: 193; Pavlides, 2006). There is no evidence that there was ever a sailing route across from Milne Bay to the Solomons or Bougainville, and it seems unlikely that a group of people could have come through all the way from New Guinea to the Solomon Islands overland, and still have their language intact enough for a relationship to Yélıˆ Dnye to be established at the end of it. Ross’s (2001b) Yele–New Britain suggestion (based on pronoun forms) is more plausible in this respect, as New Britain is much closer to the presumed point of departure. As regards the wider region, Wurm et al. (1975) said of one of their proposed early Papuan populations in New Guinea “[i]t seems tempting to suggest that this farflung substratum which may perhaps have surviving primary manifestations in some members of the West Papuan Phylum […], in the Torricelli Phylum and perhaps also in the East Papuan Phylum [Island Melanesia], may outline the earlier presence in the New Guinea area, of an old language type […] to be later overrun and reduced to substratum level by subsequent language migrations.” Wurm et al. characterized the languages in question as having a prevalence of certain pronoun forms, an overt two-gender system, a tendency to prefixing, number marking on nouns, verb stem suppletion and alternation in connection with object and subject marking, and the absence of medial verb forms (pp 940–941).2 Reesink (2005b: 206) also saw a possible connection between the Island Melanesian languages along the New Guinea north coast and through to West Papua in present-day Indonesia. Casting her net even wider, Nichols (1997, 1998) hypothesized two or three different linguistic provinces in Australia–New Guinea and the wider area of the Pacific Rim. Such connections are also explorable using the phylogenetic methods presented above. To arrive at better resolution in the results of studies such as those presented in this chapter, linguistic case studies to determine the relative time-stability of individual structural features are needed, since, although anything can be borrowed, some types of linguistic item are much less likely to spread laterally than others, and more research could establish likely “markers” of linguistic genetic relatedness in deep time. While the body of data about Island Melanesian Papuan languages has been growing, more in-depth descriptions are still required to address issues of contact and phylogeny in greater detail, as well as to explore cross-disciplinary issues. To establish whether there is a linguistic correlate of the strong genetic separation between North and South Bougainville (Friedlaender et al., 2005), for instance, we would need data for more North Bougainville languages than Rotokas, and several New Britain languages and the Reefs/Santa Cruz languages also remain understudied.
136
Increasing interdisciplinary collaboration, such as the alliance of linguistics and bioinformatic methods applied in the structural phylogenetics study, allows for new forms of hypothesis generation and evaluation, and promises advances in our understanding of the history of the earliest Island Melanesians.
Acknowledgments This work, as part of the European Science Foundation EUROCORES Programme OMLL, was supported by funds from the Nederlandse Organisatie voor Wetenschappelijk Onderzoek (NL), Vetenskapsrådet (SE), and the EC Sixth Framework Programme under Contract no. ERAS-CT-2003-980409. Additional fieldwork data used in this study were provided by Stuart Robinson (Rotokas), Tonya Stebbins (Mali), William Thurston (Aneˆm), Claudia Wegener (Savosavo). Assistance with coding of Oceanic languages from published sources was given by S. Nordhoff, V. Rodrigues, and K. Ahlzén. For permission to use unpublished materials we thank Robyn Davies and Lisbeth Fritzell (Ramoaaina), Lee Erickson (Notsi), Don Huchisson (Sursurunga), Stellan Lindrud (Kol), Lloyd Milligan (Mangseng), Greg and Mary Pearson (Uvol), and Hideki Ohtsuba (Meramera). We also thank Robert Foley, Marta Mirazon-Lahr, Stephen Levinson, and Gunter Senft for discussions on this chapter.
Notes 1. An asterisk (*) marks a reconstructed form. All cited language material is given in italics; abbreviations used in examples are found in Appendix I. 2. The basic sets of pronouns used by Wurm (1975b) have not survived the critical inspection by colleagues, as they allowed for too many alternative realizations in living languages (Voorhoeve 1987).
References Allen G. 1987. Halia grammar. Ukarumpa: Summer Institute of Linguistics. Allen G, Hurd C. 1965. Languages of the Bougainville district. Port Moresby: Department of Information and Extension Services. Bugenhagen RD. 1994. Language change on Umboi Island. In: Dutton T, Tryon DT, editors. Language contact and change in the Austronesian world. Berlin, Germany: Mouton de Gruyter. pp 43–108. Campbell L. 1997. Amerindian personal pronouns: A second opinion. Language 73: 339–51. Campbell L. 1998. Historical linguistics: An introduction. Edinburgh, UK: Edinburgh University Press.
The Languages of Island Melanesia
Campbell L. 2003. How to show languages are related: Methods for distant genetic relationship. In: Joseph BD, Janda RD, editors. The handbook of historical linguistics. Oxford, UK: Blackwell. pp 262–82. Capell A. 1962. Oceanic linguistics today. Current Anthropology 3(4): 371–96. Dunn M, Reesink G, Terrill A. 2002. The East Papuan languages: A preliminary typological appraisal. Oceanic Linguistics 41: 28–62. Dunn M, Terrill A, Reesink G, Foley RA, Levinson SC. 2005. Structural phylogenetics and the reconstruction of ancient language history. Science 309: 2072–5. Firchow IB, Firchow J. 1969. An abbreviated phoneme inventory. Anthropological Linguistics 11: 271–6. Foley WA. 1986. The Papuan languages of New Guinea. Cambridge, UK: Cambridge University Press. Foley WA. 2000. The languages of New Guinea. Annual Review of Anthropology 29: 357–404. Forster P, Toth A. 2003. Toward a phylogenetic chronology of ancient Gaulish, Celtic, and Indo-European. PNAS 100: 9079–84. Frahm RM. 1998. Baniata serial verb constructions. MA thesis. Linguistics, The University of Auckland. Friedlaender JS, Gentz F, Friedlaender FR, Kaestle F, Schurr T, Koki G, McDonough J, Smith L, Cerchio S, Mgone C, Merriwether A. 2005. Mitochondrial genetic diversity and its determinants in Island Melanesia. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts: Studies in the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 693–716. Goulden R. 1996. The Maleu and Bariai languages of West New Britain. In: Ross M, editor. Studies in Languages of New Britain and New Ireland: 1: Austronesian languages of the North New Guinea cluster in Northwestern New Britain. Canberra, Australia: Pacific Lingistics. pp 63–144. Gray RD, Atkinson QD. 2003. Language-tree divergence times support the Anatolian theory of Indo-European origin. Nature 426: 435–9. Gray RD, Jordan FM. 2000. Language trees support the express-train sequence of Austronesian expansion. Nature 405: 1052–5. Greenberg JH. 1971. The Indo-Pacific hypothesis. In: Sebeok TA, editor. Current trends in linguistics, vol 8: Linguistics in Oceania. The Hague: Mouton and Co. pp 807–71. Greenberg JH. 1987. Language in the Americas. Stanford, CA: Stanford University Press. Harrison SP. 2003. On the limits of the comparative method. In: Joseph BD, Janda RD, editors. The handbook of historical linguistics. Oxford: Blackwell. pp 213–43. Hashimoto K. nd. Ata grammar essentials. Ukarumpa: Summer Institute of Linguistics. Henderson J. 1995. Phonology and grammar of Yele, Papua New Guinea. Canberra, Australia: Pacific Linguistics B–112.
Hovdhaugen E, Mosel U, editors. 1999. Negation in Oceanic languages. Munich, Germany: Lincom Europa. Johnston RL. 1980. Nakanai of New Britain, the grammar of an Oceanic language. Canberra, Australia: Pacific Linguistics. Kirch P. 1997. The Lapita peoples. London: Blackwell. Laycock DC. 1973. Sepik languages—checklist and preliminary classification. Canberra, Australia: Pacific Linguistics B–25. Leavesley M. 2006. Late Pleistocene complexities in the Bismarck archipelago. In: Lilley I, editor. Archaeology of Oceania; Australia and the Pacific Islands. Malden MA/Oxford/Carlton Vic: Blackwell Publishing. pp 189–204. Levinson SC. 2005. Evolution of culture in a microcosm. In: Levinson SC and Jaisson P, editors. Evolution and culture. Cambridge, MA: MIT Press. pp 1–41. Levinson SC. In preparation. A grammar of Yélˆı Dnye, the Papuan Language of Rossel Island. Lincoln PC. 1978. Reef–Santa Cruz as Austronesian. In: Wurm SA, Carrington L, editors. Second International Conference on Austronesian Linguistics: Proceedings. Canberra, Australia: Pacific Linguistics C–61. pp 929–67. Lindström E. 2002. Topics in the grammar of Kuot, a nonAustronesian language of New Ireland, Papua New Guinea. PhD. Department of Linguistics: Stockholm University. Lynch J, Ross M, Crowley T. 2002. The Oceanic Languages. London: Curzon Press. Maddieson I, Levinson SC. In preparation. The phonetics of Yélˆı Dnye, the language of Rossel Island. McMahon A, McMahon R. 2003. Finding families: Quantitative methods in language classification. Transactions of the Philological Society 101: 7–55. Merriwether A, Hodgson JA, Friedlaender FR, Allaby R, Cerchio S, Koki G, Friedlaender JS. 2005. Ancient mitochondrial M haplogroups identified in the Southwest Pacific. Proceedings of the National Academy of SciencesUSAmerica 102(37): 13034–9. Milligan L. 1992. Mangseng grammar essentials. Ukarumpa: Summer Institute of Linguistics. Minett JW, Wang WS-Y. 2003. On detecting borrowing: Distance-based and character-based methods. Diachronica 20: 289–331. Mosel U. 1984. Tolai syntax and its historical development. Canberra, Australia: Pacific Linguistics B92. Nakhleh L, Ringe D, Warnow T. 2005. Perfect phylogenetic networks: a new methodology for reconstructing the evolutionary history of natural languages. Language 81: 382–420. Nichols J. 1992. Linguistic diversity in space and time. Chicago, IL: University of Chicago Press. Nichols J. 1996. The comparative method as heuristic. In: Durie M, Ross M, editors. The comparative method reviewed: Regularity and irregularity in language change. New York/Oxford: Oxford University Press. pp 39–71.
137
core studies in northern island melanesia
Nichols J. 1997. Sprung from two common sources: Sahul as a linguistic area. In: McConvell P, Evans N, editors. Archeology and linguistics: Global perspectives on ancient Australia. Melbourne, Australia: Oxford University Press. pp 135–68. Nichols J. 1998. The origin and dispersal of languages: Linguistic evidence. In: Jablonski N, Aiello LC, editors. The origin and diversification of language. San Francisco, CA: California Academy of Sciences. pp 127–70. Nichols J, Peterson DA. 1996. The Amerind personal pronouns. Language 72: 336–71. Nichols J, Peterson DA. 1998. Amerind personal pronouns: A reply to Campbell. Language 74: 605–15. Ohtsuba H. 1996. Meramera grammar essentials. Manuscript. Ukarumpa: Summer Institute of Linguistics. Osmond M. 1998. Horticultural practices. In: Ross M, Pawley A, Osmond M, editors. The lexicon of Proto Oceanic: The culture and environment of ancestral Oceanic society. Vol I: Material culture. Canberra, Australia: Pacific Linguistics C–152. pp 115–42. Osmond M, Ross M. 1998. Household artefacts. In: Ross M, Pawley A, Osmond M, editors. The lexicon of Proto Oceanic: The culture and environment of ancestral Oceanic society. Vol I: Material culture. Canberra, Australia: Pacific Linguistics C–152. pp 67–114. Pavlides C. 2006. Life before Lapita: New developments in Melanesia’s long-term history. In: Lilley I, editor. Archaeology of Oceania: Australia and the Pacific Islands. Malden MA/Oxford/Carlton Vic: Blackwell Publishing. pp 205–27. Pawley A, Ross M. 1995. The prehistory of Oceanic languages, a current view. In: Bellwood P, Fox JJ, Tryon D, editors. The Austronesians. Canberra: Department of Anthropology, Research School of Asian and Pacific Studies, Australian National University. pp 39–74. Rankin RL. 2003. The comparative method. In: Joseph BD, Janda RD, editors. The handbook of historical linguistics. Oxford, UK: Blackwell. pp 183–212. Rath D. 1986. Mengen grammar essentials.Manuscript. Ukarumpa, Papua New Guinea: Summer Institute of Linguistics. Reesink G. 2002. Clause-final negation: Structure and interpretation. In: McGregor WB, editor. The interaction of data, description, and theory in linguistics: Functional perspectives. Special issue of functions of language. Amsterdam: John Benjamins. pp 239–68. Reesink G. 2005a. Sulka of East New Britain: A mixture of Oceanic and Papuan traits. Oceanic Linguistics 44: 145–93. Reesink G. 2005b. West Papuan languages: Roots and development. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts: Studies in the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 185–218. Rexová K, Frynta D, Zrzav y´ J. 2003. Cladistic analysis of languages: The Indo-European classification based on lexico-statistical data. Cladistics 19: 120–7.
138
Ringe D, Warnow T, Taylor A. 2002. Indo-European and computational cladistics. Transactions of the Philological Society 100: 59–129. Robinson S. 2006. The phoneme inventory of the Aita dialect of Rotokas Oceanic Linguistics 45: 206–9 Ross M. 1988. Proto Oceanic and the Austronesian languages of Western Melanesia. Canberra, Australia: Pacific Linguistics C–98. Ross M. 1994. Areal phonological features in north central New Ireland. In: Dutton T, Tryon DT, editors. Language contact and change in the Austronesian world. Berlin, Germany: Mouton de Gruyter. pp 551–72. Ross M. 1996. Contact-induced change and the comparative method: cases from Papua New Guinea. In: Durie M, Ross M, editors. The comparative method reviewed. New York: Oxford University Press. pp 180–217. Ross M. 1998a. Proto-Oceanic adjectival categories. Oceanic Linguistics 37: 85–119. Ross M. 1998b. Possessive-like attributive constructions in the Oceanic languages of northwest Melanesia. Oceanic Linguistics 37: 234–76. Ross M. 1999. Exploring metatypy: How does contactinduced typological change come about? Keynote talk given at the Australian Linguistic Society’s annual meeting, Perth, 1999. Ross M. 2001a. Contact-induced change in Oceanic languages in north-west Melanesia. In: Aikhenvald A, Dixon RMW, editors. Areal diffusion and genetic inheritance: problems in comparative linguistics. Oxford: Oxford University Press. pp 134–66. Ross M. 2001b. Is there an East Papuan phylum? Evidence from pronouns. In: Pawley A, Ross M, Tryon D, editors. The boy from Bundaberg: Studies in Melanesian linguistics in honour of Tom Dutton. Canberra, Australia: Pacific Linguistics. pp 301–21. Ross M. 2002. Bali-Vitu. In: Lynch J, Ross M, Crowley T, editors. The Oceanic languages. London: Curzon Press. pp 362–86. Ross M. 2005. Pronouns as a preliminary diagnostic for grouping Papuan languages. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan pasts: Studies in the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra, Australia: Pacific Linguistics. pp 15–65. Ross M, Durie M. 1996. Introduction. In: Durie M, Ross M, editors. The comparative method reviewed: irregularity and regularity in linguistic change. New York: Oxford University Press. pp 3–38. Ross M, Pawley A, Osmond M, editors. 1998. The lexicon of Proto Oceanic: The culture and environment of ancestral Oceanic society, Vol 1, Material culture. Canberra, Australia: Pacific Linguistics C–152. Ross M, Pawley A, Osmond M. 2003. The lexicon of Proto Oceanic: The culture and environment of ancestral Oceanic society, Vol 2. Canberra, Australia: Pacific Linguistics. pp 545. Spriggs M. 1997. The Island Melanesians. London: Blackwell.
The Languages of Island Melanesia
Stebbins T. 2002. Noun class markers in Mali-Baining. Paper presented at the 5th International conference on Oceanic linguistics. January 2002, Canberra, Australia. Terrell J, Kelly K, Rainbird P. 2001. Foregone conclusions? In search of “Papuans” and “Austronesians.” Current Anthropology 42: 97–124. Terrill A. 2002. Systems of nominal classification in East Papuan languages. Oceanic Linguistics 41: 63–88. Terrill A. 2003. Linguistic stratigraphy in the central Solomon Islands: Lexical evidence of early Papuan/ Austronesian interaction. Journal of the Polynesian Society 112: 369–401. Thomason SG. 2001. Language contact: An introduction. Washington DC: Washington University Press. Thomason SG, Kaufman T. 1988. Language contact, creolization and genetic linguistics. Berkeley, CA: University of California Press. Thurston WR. 1982. A comparative study of Aneˆm and Lusi. Canberra, Australia: Pacific Linguistics B–83. Thurston WR. 1987. Processes of change in the languages of north-western New Britain. Canberra, Australia: Pacific Linguistics B–99. Thurston WR. 1994. Renovation and innovation in the languages of north-western New Britain. In: Dutton T, Tryon D, editors. Language contact and change in the Austronesian world. Berlin, Germany: Mouton de Gruyter. pp 573–609. Trask RL. 1996. Historical linguistics. London: Arnold/ New York: Oxford University Press. Voorhoeve CL. 1987. Worming one’s way through New Guinea: the case of the peripatetic pronouns. In: Laycock DC, Winter W, editors. A world of language: Papers presented to Professor S.A. Wurm on his 65th birthday. Canberra, Australia: Pacific Linguistics. pp 709–27.
Warnow T, Evans SN, Ringe D, Nakhleh L. 2005. A stochastic model of language evolution that incorporates homoplasy and borrowing. In: Clackson J, editor. Phylogenetic methods and the prehistory of languages. Cambridge: Cambridge University Press. Watkins C. 2000. Indo-European and the Indo-Europeans. In: American heritage dictionary of the English language. Boston, MA: Houghton Mifflin. Online resouce: www.bartleby.com/61/. [2006-11-16]. Wurm SA. 1975a. The East Papuan Phylum in general. In: Wurm SA, editor. Papuan languages and the New Guinea linguistic scene. Canberra, Australia: Pacific Linguistics C–38. pp 783–804. Wurm SA. 1975b. Personal pronouns. In: Wurm SA, editor. Papuan languages and the New Guinea linguistic scene. Canberra, Australia: Pacific Linguistics C–38. pp 191–216. Wurm SA. 1978. Reefs–Santa Cruz: Austronesian, but ...! In: Wurm SA, Carrington L, editors. Second International Conference on Austronesian Linguistics: Proceedings. Canberra, Australia: Pacific Linguistics C–61. pp 969–1010. Wurm SA. 1982. Papuan languages of Oceania. Tübingen: Gunter Narr Verlag. Wurm SA, Laycock DC, Voorhoeve CL, Dutton TE. 1975. Papuan linguistic prehistory, and past language migrations in the New Guinea area. In: Wurm SA, editor. New Guinea area languages and language study. Canberra, Australia: Pacific Linguistics. pp 935–60.
139
core studies in northern island melanesia
Appendix I
Abbreviations Used in Examples
In Gloss
Legend
In Gloss
Legend
– . 1 2 3 ART BODY PART CAUS CONT DU EDIBLE F FOC FOOD FUT GENERAL HAB
separator of stem and affix or multiple affixes separator of senses of indivisible word form 1st person 2nd person 3nd person article inalienable possessive classifier causative continuous dual number food possessive classifier feminine gender focus marker food possessive classifier future tense general possessive classifier habitual aspect
INDEF IRR M N NEUTRAL NF NOM OBJ PERF POSS PREP PREPV REAL RECIP SG SUBJ TRANS
indefinite irrealis masculine gender neuter gender neutral possessive classifier non-finite verb form nominalization marker object perfect tense possessive marker preposition prepositional verb realis reciprocal singular number subject transitivization marker
Appendix II
Languages in the Database (see further, Dunn et al. 2005, Supporting Online Material).
Language Name
Island/Archipelago
Major Affiliation
Anêm Ata Bali-Vitu Banoni Bilua Buin Gapapaiwa Kairiru Kaulong Kilivila Kokota Kol Kuot Lavukaleve Mali Baining Motuna (Siwai) Nalik Nasioi Rotokas Roviana Savosavo Siar Sisiqa Sudest Sulka Taiof Takia Touo (Baniata) Tungag Yabem Yélˆı Dnye
New Britain New Britain New Britain Bougainville Solomons Bougainville Milne Bay New Guinea New Britain Milne Bay Solomons New Britain New Ireland Solomons New Britain Bougainville New Ireland Bougainville Bougainville Solomons Solomons New Ireland Solomons Milne Bay New Britain Bougainville New Guinea Solomons New Ireland New Guinea Milne Bay
Papuan Papuan Oceanic Oceanic Papuan Papuan Oceanic Oceanic Oceanic Oceanic Oceanic Papuan Papuan Papuan Papuan Papuan Oceanic Papuan Papuan Oceanic Papuan Oceanic Oceanic Oceanic Papuan Oceanic Oceanic Papuan Oceanic Oceanic Papuan
140
Major Subgroup
Meso-Melanesian Meso-Melanesian
Papuan Tip North New Guinea North New Guinea Papuan Tip Meso-Melanesian
Meso-Melanesian
MesoMelanesian MesoMelanesian MesoMelanesian Papuan Tip Meso-Melanesian North New Guinea Meso-Melanesian North New Guinea
9 Inferring Prehistory from Genetic, Linguistic, and Geographic Variation Keith Hunley, Michael Dunn, Eva Lindström, Ger Reesink, Angela Terrill, Heather Norton, Laura Scheinfeldt, Françoise R. Friedlaender, D. Andrew Merriwether, George Koki, and Jonathan S. Friedlaender
Introduction A primary goal of this volume is to improve our understanding of the prehistory of Northern Island Melanesia. This task is impeded by the fact that human groups behave in ways that frequently obscure or erase evidence of past population events. Despite these behaviors, researchers commonly attempt to reconstruct the past by examining the fit of biological and cultural data to simple models of population history. Even when data do not strictly conform to these models, this approach frequently provides important insights into past population processes. In this spirit, our strategy in this chapter is to formally test the fit of Northern Island Melanesian biological and linguistic data to two essentially polar models of population history and then to systematically examine the causes of any observed departures from these models. The first model, termed “population fissions,” is one in which populations split as they expand to occupy new territories. After splitting, the daughter populations remain relatively isolated, and, as a result, biological and linguistic evolution is branching, or tree-like, in nature (Cavalli-Sforza and Edwards, 1964 Cavalli-Sforza and Piazza, 1975; Cavalli-Sforza et al., 1988). If this model applies to Island Melanesia, reconstructing its history is in principle a simple matter of reconstructing the history of population fissions. In fact, this model may be well suited to Island Melanesia, both because the region is
comprised of islands, and also because travel within the larger islands with rugged interiors is physically difficult. These geographic factors might contribute to population isolation subsequent to splitting, and hence to tree-like evolution. The second model, termed “isolation by distance,” is one in which people exchange genes and features of their languages with their geographic neighbors, principally through local marital exchange. The accumulation of this exchange over time will produce a correlation between population genetic, linguistic, and geographic distances (Kimura and Weiss, 1964; Malécot, 1948; Wright, 1943). If this model applies to Island Melanesia, any previous history of tree-like evolution would have been erased. As a result, our ability to reconstruct the deeper population history of the region, including its initial colonization, would be impeded. On the other hand, we would have instead gained important insights into the details of migration rates and patterns in the region. Because people have lived in Island Melanesia for tens of thousands of years (Leaveseley et al., 2002; Pavlides and Godsen, 1994; Wickler and Spriggs, 1988), it is also possible that biological and linguistic patterns in the region conform to this model. On the surface, these models may appear overly simplistic to anthropologists. Nonetheless, an advantage of testing the fit of data to these models is that deviations from them, when viewed in concert with other types of
141
core studies in northern island melanesia
anthropological data, may provide important information about past population events. For example, CavalliSforza and colleagues (1988) examined the population fissions model by comparing the fit of global gene and language trees. They found that the trees fit well in all but four instances. Using historical and archaeological data, they explained these four deviations in terms of language replacement and population admixture. Isolation by distance has also been tested at different geographic levels (Barbujani et al., 1995; Eller, 1999; Fuselli et al., 2003; Relethford, 2004; Serre and Pääbo, 2004). For example, Fuselli et al. (2003) recently tested the fit of mitochondrial sequence variation to the isolation by distance model in Native South America. They identified different patterns of genetic and geographic correspondence between eastern Andean and western Amazonian populations. While the genetic data did not strictly fit isolation by distance in either region, their results led them to conclude that the pattern of genetic and geographic correspondence in the eastern region was caused by population expansion followed by a series of founder events, while that in the west was caused by long-range dispersion and gene flow. With this background in mind, testing the fit of the Northern Island Melanesian genetic, linguistic, and phenotypic data described in previous chapters to these two idealized models should be useful both as an exercise in testing their utility, and in gaining a better perspective on past population processes in the region. For the population fissions model, we test the hypothesis that trees constructed from genetic and linguistic distance data represent the genetic structure of Island Melanesian populations. These populations are identified by their linguistic affiliations, referring to Austronesian- and Papuan-speaking populations. The Austronesian-speaking groups are all linguistic descendants of a relatively recent linguistic dispersal, whereas the Papuan-speaking groups are a residual category descended from possibly heterogeneous languages spoken in the region prior to the Austronesian dispersal (chapter 8). For the isolation by distance model, we test the hypotheses that population genetic, linguistic, and phenotypic distances are correlated with geographic distance. We then examine causes of departures from the predictions of both models in light of ethnographic, archeological, and historic data. Finally we discuss the relevance of our results in terms of the prehistory of the region.
Materials and Methods Materials The genetic, linguistic, and phenotypic (skin and hair reflectance) data used in the following analyses are based
142
on the materials from the “core” northern Island Melanesian sample described and analyzed in chapters 4, 5, 6, and 8. Sample details are listed in table 9.1. Our sample differs somewhat from those in previous chapters because we excluded populations containing less than 15 individuals. In addition, for the autosomal single nucleotide polymorphisms (SNPs), we removed haplotypes containing fewer than 5 nucleotides, and for the mtDNA sequences, we removed haplotypes containing more than 5% missing nucleotides. We also excluded the Y SNPs described in chapter 5 in order to avoid problems associated with ascertainment bias. We stress that this chapter examines genetic variation only at the population level. The results and conclusions of this chapter should be interpreted in combination with the analyses presented in chapters 4–6, which are primarily concerned with the description of haplotypes and their distributions across the region.
Analytical Methods Population Fissions Model To test the hypothesis of genetic treeness for a given dataset, we first estimate the genetic differences within and between all populations for a given genetic system, e.g., the autosomal SNPs. These are termed the “realized” genetic differences (Nei, 1987). Next, we construct a population gene tree (NJ, Saitou and Nei, 1987) from these differences and then use the methods described below to determine whether the within and between population genetic differences estimated from this NJ tree are statistically indistinguishable from the realized genetic differences. If they are, the tree is said to “fit” the genetic data. When a tree fits, we conclude that its pattern of branching can be taken to represent the history of the fissions of the groups represented in the tree. For example, imagine that figure 9.1 represents the NJ tree constructed from the autosomal SNPs. If this tree perfectly fits the realized pattern of within- and betweenpopulation autosomal SNP variation, we would conclude that an Oceanic-speaking group and a Papuan-speaking group were produced by a fission of a common ancestral group, and that both groups subsequently evolved independently of one another. In addition, the branching pattern within each language group would indicate that all of these populations were subsequently isolated from one another and then evolved independently in a perfect branching fashion. In this hypothetical example, if we were indeed able to demonstrate that population genetic evolution was tree-like, we would have learned a great deal about the population prehistory of the region. The following sections provide a more formal description of this model-fitting approach.
Inferring Prehistory from Genetic, Linguistic, and Geographic Variation
Table 9.1
Sample Details mtDNA
Austosomal SNP
Y STR
Skin Reflectance
Population
Langauge group
Island
n
p
n
p
n
SSRDa
n
Mean
Aita Anêm Ata Buka Kaket Kol Kove Kuot Lavongai Loso Madak Mali Mamusi Mangseng Melamela Mengen Mussau Nagovisi Nalik Nakanai Nasioi Notsi Rotokas Saposa Siwai Sulka Teop Tigak Tolai
Papuan Papuan Papuan Austronesian Papuan Papuan Austronesian Papuan Austronesian Austronesian Austronesian Papuan Austronesian Austronesian Austronesian Austronesian Austronesian Papuan Austronesian Austronesian Papuan Austronesian Papuan Austronesian Papuan Papuan Austronesian Austronesian Austronesian
Bougainville New Britain New Britain Bougainville New Britain New Britain New Britain New Ireland New Hanover New Britain New Ireland New Britain New Britain New Britain New Britain New Britain Mussau Bougainville New Ireland New Britain Bougainville New Ireland Bougainville Bougainville Bougainville New Britain Bougainville New Ireland New Britain
52 17 58 15 59 57 17 62 18 16 31 58 62 17 21 23 16 15 24 65 30 17 19 21 19 28 17 26 77
0.012 0.018 0.009 0.020 0.012 0.017 0.011 0.004 0.008 0.015 0.009 0.014 0.014 0.017 0.022 0.018 0.008 0.001 0.009 0.014 0.015 0.008 0.021 0.008 0.011 0.014 0.016 0.006 0.019
34 47 48 47 41 10 27 49 35 25 27 29 17 45 27 27 16
0.088 0.130 0.156 0.135 0.087 0.102 0.182 0.085 0.134 0.198 0.170 0.039 0.174 0.201 0.018 0.212 0.189
18 34 44 9 39
3.1 3.4 4.3 3.1 4.0 3.3 4.7 3.8 4.6 4.4 3.7 4.0 4.8 3.5 4.1 4.6 4.3 5.3
36 55 68 13 45 53 102 17 26
91.4 69.6 67.7 89.7 65.0 76.0 77.1 70.3 78.0
79 14 37 34
68.9 64.7 69.2 65.0
27 98 23
72.2 66.1 74.1
4.5 4.6 3.7 4.6 4.9
48 24 27 18
86.7 93.7 72.6 68.3
25 32 43 15 19 24 43 11 14 23
17 36 14 26 33 18 22 49
Structural Linguisticb (Rotokas) x x (Halia) x x x x x (Nakanai) x x (Uvol) x x x x x (Loso) x x x x x x (Saposa) x x
a
Sum of Squared Number of Repeat Differences between haplotypes. Structural linguistic features were scored for these populations. x indicates that linguistic data was sampled directly from the population indicated in the first column. Where linguistic data was lacking, data from another closely related language group was used as a proxy where possible (substitute language shown in parentheses). All data including hair reflectance (not listed) available upon request from author. b
Figure 9.1 Tree showing hypothetical Papuan–Austronesian division. Equal branch lengths signify equal evolutionary rates.
Tree Construction. To construct genetic trees, we first estimated genetic distances between populations for each type of genetic data. For the autosomal SNPs and mtDNA sequences, genetic distances consisted of the adjusted net number of nucleotide substitutions between haplotypes (Meyer et al., 1999; Nei, 1987; Tamura and Nei, 1993). For the Y chromosome short tandem repeat polymorphisms (STRs), we estimated population pairwise RST distances (Slatkin, 1995). We then constructed NJ trees from these genetic distances and employed the statistical analyses described below to test the hypothesis that each tree represented the genetic structure of the Island Melanesian populations. We also constructed a language NJ tree from the proportion of pairwise linguistic differences between pairs of populations (Dunn et al., 2005; see also chapter 8). We then
143
core studies in northern island melanesia
employed the statistical analyses described below to test the hypothesis that this language tree represented the genetic structure of the Island Melanesian populations. The rationale for testing the fit of the language tree is that, even if the genetic data are not tree-like, patterns of genetic and linguistic exchange among populations may have been similar. If so, the language tree would strongly resemble one or more of the gene trees and would have a similar fit to that tree. All the resulting constructed trees are depicted in figure 9.2. Statistical Analyses. We first constructed a matrix of realized within- and between-population genetic differences for each genetic system. In this case, our measure of genetic difference is gene identity (Nei, 1987). We next attempted to fit the NJ trees for each genetic system to their respective realized gene identity matrices using maximum likelihood (Cavalli-Sforza and Piazza, 1975; Hunley and Long, 2005; Long and Kittles, 2003; Urbanek et al., 1996). We also attempted to fit the language tree to each genetic system. This maximum likelihood fitting method produces a matrix of “expected” gene identities contingent on the assumption that the tree being fitted accurately represents the true genetic relationships among populations. The fit of the realized to expected gene identities is estimated with a likelihood ratio statistic Λ, which under the idealized circumstances of a large number of independent polymorphic sites is distributed as a chi-squared random variable (Cavalli-Sforza and Piazza, 1975). The degrees of freedom associated with this statistic is equal to r(r + 1)/2 minus the number of parameters specified by the fitted tree, where r is the number of populations sampled. If a tree fits, Λ will be low relative to its degrees of freedom.
Test of Isolation by Distance We evaluated the alternative model, isolation by distance, in two ways. First, for the autosomal SNPs and mtDNA ˆ ST genetic distances between sequence data, we estimated Φ population pairs (Excoffier et al., 1992). To visualize the ˆ ST/(1−Φ ˆ ST) against the results, we plotted the quantity Φ logarithm of geographic distance. Isolation by distance predicts a linear relationship between these transformed variables (Rousset, 1997; Slatkin and Maddison, 1990). This procedure was repeated for the Y chromosome STRs using RST as the measure of between-population genetic distance (Slatkin, 1995). We then examined the correlations between distances within geographic (islands) and linguistic (Papuan vs. Austronesian) subgroups of each plot. We applied a similar procedure to test isolation by distance for the phenotypic and linguistic data. We estimated phenotypic distances using methods described by Relethford and colleagues (Relethford, 1996; Relethford
144
and Blangero, 1990; Relethford et al., 1997; WilliamsBlangero and Blangero, 1992). Linguistic distances were estimated as the proportion of pairwise linguistic differences between pairs of populations. We then applied least squares regression to the pairwise population linguistic differences and phenotypic distances vs. geographic distance. To visualize the results, we constructed plots of linguistic vs. geographic and phenotypic vs. geographic distances. We then examined correlations between distances within geographic and linguistic subgroups of each plot.
Additional Population Genetic Analyses Based on the results of the above analyses, we applied several additional tests to further explore the relationship between the genetic, phenotypic, linguistic and geographic data. First we constructed plots of genetic vs. linguistic distances. Second, we applied analysis of molecular variance (AMOVA, Excoffier et al., 1992) to linguistic and geographic subdivisions of the genetic data. We also applied F-statistic analyses to the full sample and to each island for the skin and hair reflectance data (Relethford and Blangero, 1990; Relethford et al., 1997; Relethford, 1996). Third, we estimated genetic variation within each population using methods described in Tamura and Nei (1993) and Slatkin (1995). We then applied non-parametric Monte Carlo simulations to test the null hypothesis that the within population variation for the Papuan-speaking populations was the same as that for the Austronesianspeaking populations. Finally, we applied multidimensional scaling (MDS, Kruskal and Wish, 1978) separately to genetic, linguistic and phenotypic distances. Where appropriate, significance levels were adjusted to take into account the fact that we conducted multiple tests. Adjustments were made using the sequential-rejective approach described by Holm (1979). Computations for the population genetic analyses were performed using Arlequin (Schneider et al., 2000), Phylip (Felsenstein, 2005) and RMET (Relethford, 1996) and SPSS (SPSS Inc. v8.0).
Results Neighbor Joining Trees The NJ trees reveal several interesting details about the structure of Northern Island Melanesian populations. We begin by examining each tree for linguistic and geographic (island) structure, and then discuss the results of fitting these trees to the genetic data. The autosomal SNP NJ tree in figure 9.2A shows that most populations are connected either by short or
Inferring Prehistory from Genetic, Linguistic, and Geographic Variation
Figure 9.2 Neighbor joining trees constructed from genetic and linguistic distances *Papuan-speaking populations. A, autosomal SNPs; B, mtDNA; C, Y chromosome STRs; D, structural linguistic. Circles, Bougainville; squares, New Britain; triangles, New Ireland.
145
core studies in northern island melanesia
negative branches. Negative branches indicate a pattern of between-population genetic variation that is inconsistent with branching, or tree-like, evolution. The substantial number of negative branches in this tree (12) and other trees immediately suggests that population genetic evolution has not been tree-like in the region. In addition, the populations in the autosomal SNP tree do not cluster by language, and with the exception of Bougainville, the populations do not cluster by islands. However, even for Bougainville, the Saposa and Teop branches are negative. The mtDNA NJ tree in figure 9.2B again has many negative branches. However, in this case, the populations do tend to cluster by islands. For example, the New Britain populations generally form a distinct cluster, and the New Ireland populations also tend to form a second cluster. On the other hand, the Bougainville populations are distributed throughout the tree. The Y chromosome STR NJ tree in figure 9.2C also has many negative branches. In addition, the populations generally do not cluster by language or geography. As discussed below, we believe the Y chromosome STR data contain limited information, which may contribute to a lack of discernable patterning in this tree. The language tree constructed from structural linguistic distances is presented in figure 9.2D. This tree differs markedly from the other trees in two respects. First, all of the branches are positive. Second, with two exceptions, the language tree is the only one that separates Papuan- and Austronesian-speaking populations. These exceptions are the Papuan-speaking Anêm and Ata. These two New Britain populations are both surrounded by Austronesian-speaking populations (see map in chapter 8), in contrast to the other Papuan-speaking populations of New Britain, which are clustered together in the northeast portion of the island. An additional feature of the language tree is that it lacks geographic structure, contrasted, for example, to the mtDNA tree, which, as indicated above, contains substantial geographic structuring. This result suggests extensive linguistic coancestry and exchange between islands. An important result is that the language tree does not resemble any of the gene trees. The lack of similarity between the language and genetic trees supports the notion of distinctive population genetic vs. linguistic evolutionary histories.
Population Fissions The maximum likelihood procedure we used was unable to estimate an expected gene identity matrix for any of the trees. This result means either that population genetic evolution has not been tree-like, or that the genetic data contain insufficient information to capture existing treeness.
146
This latter explanation may well apply to the autosomal SNPs and Y chromosome STRs because these data are comprised of relatively few loci. However this explanation probably does not apply to the mtDNA sequences. These d-loop sequences are comprised of 977 nucleotides, 176 of which are polymorphic. Of these polymorphic nucleotides, 31 occur only in Papuan-speaking populations or Austronesian populations, and 25 occur in multiple populations restricted to a single island. This pattern of structured variation suggests that the mtDNA sequences would contain sufficient information to resolve treeness if these genetic data were indeed tree-like. Hence, the reason why none of the trees fits the mtDNA sequence distance data is very likely because mtDNA evolution has not been tree-like.
Isolation by Distance Figure 9.3A–C plots the transformed genetic and geographic distances for the autosomal SNP, mtDNA, and Y chromosome STR markers. Table 9.2 lists the correlations and their statistical significance for the full plots and for sub-groupings of the data. Several plots indicate a statistically significant but weak relationship between genetic and geographic distances. For example, figure 9.3A plots the transformed distances for the autosomal SNPs. The plot indicates that only 8% (0.292) of the variation in genetic distance is explained by geographic distance. A closer inspection of the plot confirms that the Aita are responsible for a large portion of this correlation. When the Aita are excluded, the magnitude of the correlation drops substantially. The autosomal SNP–geography correlation coefficient was higher within islands (0.40), but this correlation was not statistically significant. It is interesting to note that this correlation coefficient is almost identical to that previously measured for blood protein frequency variation in Bougainville (Friedlaender, 1971). Figure 9.3B plots the transformed distances for the mtDNA sequences. The correlation for the full plot is not statistically significant. However, there is a statistically significant relationship between mtDNA genetic and geographic distance between Papuan-speaking populations on both New Britain and Bougainville. This correlation is largely caused by a single population on each island: the Anêm on New Britain and the Aita on Bougainville. The correlations become non-significant when these populations are removed from the analyses. Figure 9.3C plots the transformed genetic and geographic distances for the Y chromosome STRs. In this case, the overall correlation is statistically significant but still relatively weak. Much of the strength of the correlation is the result of an association among Austronesianspeaking populations within islands. In this case, though
Inferring Prehistory from Genetic, Linguistic, and Geographic Variation
Figure 9.3 Plots of transformed genetic vs. geographic (A–C) and untransformed linguistic and phenotypic vs. geographic distances (D–F). See color insert.
the correlations are weak, they are more robust in that they are not caused only by one or two populations. The untransformed linguistic and geographic distances are plotted in figure 9.3D. There is a significant but weak relationship between these distances. A closer inspection of the plot reveals three separate strata; one for Papuan-speaking populations, one for Austronesian speakers, and one for the between Papuan–Austronesianspeaking population comparisons. All the strata reveal
similar weak correlations. Importantly, the plots reveal that the linguistic distances between Austronesian-speaking populations are lower than those for Papuan-speaking populations; this is predicted given that the Austronesian languages are closely related enough to form a reconstructable subgroup, whereas the Papuan group is ancient, and not amenable to lexical reconstruction. In addition, the correlation is weak but significant for both within-island and between-island comparisons.
147
core studies in northern island melanesia
Table 9.2
Correlations of Genetic, Linguistic and Phenotypic Distance with Geographic Distance Correlations with Geographic Distance Autosomal SNP
All populations Within Islands Between Islands Within Islands, Austronesian only Within Islands, Papuan only
mtDNA
Y STR
Language
Skin
Hair
r
na
r
n
r
n
r
n
r
n
r
n
0.29* 0.4 0.29*
136 34 102
0.14 0.2 −0.09
378 122 256
0.27* 0.23 0.03
253 94 159
0.31* 0.27* 0.29*
300 98 202
0.68* −0.15 0.63*
190 52 138
0.1 −0.25 0.2
153 42 111
0.36
17
0.2
37
0.41
37
0.22
27
−0.17
30
−0.3
23
na
1
0.66*
21
−0.13
10
0.24
17
na
6
na
1
a
Number of pairwise population comparisons. *Adjusted p-value = 0.004. na: insufficient number of populations.
Figure 9.3E plots the untransformed skin reflectance phenotype and geographic distances. Unlike so many of the previous associations, there is a strong correlation between the distances. As the plot indicates, this correlation is caused by phenotypic differences between islands. Figure 9.3F plots the untransformed hair reflectance and geographic distances. The plot reveals no correlation between these distances whatsoever.
Additional Analyses: Genetic–Linguistic Correlations Figure 9.4 plots genetic distances against linguistic differences for the three genetic systems, and table 9.3 lists the correlations and their statistical significance for the full plots and for sub-groupings of the data. The autosomal SNP correlation is relatively high but, as figure 9.4A
Figure 9.4 Plots of genetic vs. linguistic distances. See color insert.
148
Inferring Prehistory from Genetic, Linguistic, and Geographic Variation
Table 9.3
Correlation of Genetic with Linguistic Distance Linguistic–Genetic Correlations Autosomal SNP
All populations Within Islands Between Islands Within Islands, Austronesian only Within Islands, Papuan only
mtDNA
Y STR
r
n
r
n
r
n
0.43* 0.10 0.49* 0.39 na
120 32 88 16 1
0.38* 0.45* 0.33* −0.09 0.35
300 98 202 28 18
0.14 0.24 0.07 0.20 −0.18
210 79 131 28 10
*Adjusted p-value = 0.004.
quite differently in Island Melanesia. This difference is at least partly the result of differing histories of linguistic and genetic coancestry and exchange in the region.
reveals, the correlation is heavily influenced by the Aita. The correlations for the mtDNA–linguistic data are more robust (figure 9.4B). They are not caused by one or a few populations, and the correlation is significant both for populations on the same and on different islands. However, we emphasize that the magnitude of all of these correlations is relatively weak. For example, the largest correlation coefficient in table 9.3 is 0.49 (between islands for autosomal SNPs). This means that genetic distance explains at most about 24% (0.492) of the variation in linguistic difference. However, again in this case the correlation is strongly influenced by the Aita. The correlation coefficient decreases to 0.17 (r2 = 0.03) when the Aita are removed from the analysis, and the correlation is no longer statistically significant. Most of the correlation coefficients in table 9.3 are substantially smaller than this maximum value of 0.49. In addition, none of the Y chromosome STR correlations are statistically significant. These results indicate that linguistic and genetic variations are patterned
Table 9.4
ΦSR ΦST ΦRT
ΦSR ΦST ΦRT
The AMOVA results, presented in table 9.4, all indicate that a substantial portion of the total genetic variation is apportioned between populations in the total sample (ΦST) and between populations within islands (ΦSR). For example, the ΦST values for the three genetic systems ranged from 0.1461 for Y chromosome STRs to 0.3361 for the mtDNA. The ΦSR values ranged from 0.0617 for the autosomal SNPs to 0.2174 for the mtDNA. In addition, a substantial portion of variation is apportioned between islands (ΦRT) for each genetic system. These ΦRT values ranged from 0.0719 for the Y chromosome STRs to 0.1517 for the mtDNA. All values are statistically significant. These values are similar to those reported in chapters 4–6.
Analysis of Molecular Variance Autosomal SNP
ΦSR ΦST ΦRT
Analysis of Molecular Variance and F Statistics
mtDNA
Y Chromosome STR
Geography
Language
Geography
Language
Geography
0.0617* 0.1984* 0.1458* Geography Austronesian 0.0457* 0.1503* 0.11 Geography Papuan na
0.1458* 0.1657* −0.0238
0.2174* 0.3361* 0.1517* Geography Austronesian 0.1030* 0.1832* 0.0894* Geography Papuan 0.3067* 0.4575* 0.2174*
0.2930* 0.2959* 0.0041
0.0800* 0.1461* 0.0719* Geography Austronesian 0.0593* 0.1279* 0.0730* Geography Papuana 0.1310* 0.1682* 0.043
Language 0.1360* 0.1260* −0.0116
Geography: populations divided into islands. Language: populations divided into Papuan- and Austronesian-speaking. a Sample contains only 1 population each from Bougainville and New Ireland. *Adjusted p-value = 0.01.
149
core studies in northern island melanesia
Slight differences are the result of variation in the samples used in each chapter (described in the methods section). For each genetic system, more variation is apportioned between populations (ΦST and ΦSR) for the Papuanspeaking populations than for the Austronesian-speaking populations. This result most likely reflects either longer isolation of or less migration between Papuan- than Austronesian-speaking populations. For the mtDNA, a larger portion of the variation is also apportioned between islands for the Papuan sample (ΦRT). This result suggests more recent common ancestry or greater between-island female migration for Austronesian-speaking populations. In contrast to the finding that a great deal of genetic variation is apportioned geographically, the AMOVA analyses indicate that none of the genetic variation is apportioned linguistically (Language ΦST). This result suggests that any pre-existing genetic distinctions between these groups have been blurred by intermixture between them over the last 3,300 years. For skin reflectance, FST was high for the full sample (0.5224), but low within islands (average for New Britain, New Ireland, and Bougainville = 0.0503), and even lower between Papuan- and Austronesian-speaking groups (0.0131). For hair reflectance, compared to the skin data, FST was considerably lower for the full sample (0.0916), somewhat higher within islands (average = 0.1488) and similarly very low between Papuan- and Austronesian-speaking groups (0.0147). Hence, as with the genetic data, substantial phenotypic variation is apportioned geographically but not linguistically.
Within-Population Genetic Variation Table 9.1 lists the various measures of within-population genetic variation for each genetic system. Monte Carlo simulations indicate that within population genetic variation is similar for Austronesian- and Papuan-speaking populations. This result suggests similar effective population sizes for Papuan- and Austronesian-speaking groups, lending weight to the hypothesis that variation in the AMOVA and FST results reported above is caused by differences in either the time of separation or variation in migration rates between Papuan- and Austronesian-speaking populations.
Multidimensional Scaling MDS represents an alternative to trees for viewing population relationships. Figure 9.5 contains the MDS plots for the genetic, linguistic, and phenotypic data. In each plot, the populations are coded to reflect linguistic status and geographic location. To varying degrees, the autosomal SNP, mtDNA, and linguistic plots are characterized by a closer clustering of the Austronesian-speaking populations
150
versus a relatively more dispersed situation for the Papuan-speaking populations. The fact that Papuanspeaking populations do not cluster together in the same way that the Austronesian populations do, and the fact that many Papuan-speaking populations are genetically closer to Austronesian-speaking populations than they are to each other, explains why the prior AMOVA results do not capture this very interesting linguistic pattern. This is a comprehensible and reasonable result, given the relatively long tenure of Papuan-speaking populations in the region. This long tenure has produced substantial linguistic and genetic differences between Papuan-speaking groups that are not observed for the Austronesian-speaking groups. These plots also reveal varying degrees of island-level geographic structure. This structure is most pronounced for skin reflectance and least pronounced for the Y chromosome STRs.
Discussion: Why Don’t the Models Fit Better? Population Fissions As mentioned at the outset, the population fissions model is described by Cavalli-Sforza and colleagues in several important publications (Cavalli-Sforza, 1997; CavalliSforza et al., 1988, 1992, 1994). The model has been used to investigate the evolutionary history of human populations at the global (Cavalli-Sforza et al., 1988; Ramachandran et al., 2005) and regional levels (Cavalli-Sforza et al., 1994). It assumes that populations descend from a common ancestral group through a series of successive fissions. Following these fissions, populations evolve genetically and linguistically independently of one another. The model is relatively easy to test by examining the fit of gene and language trees to observed patterns of population genetic variation. Several recent publications have tested the model and systematically examined causes of departures from its predictions (Cavalli-Sforza et al., 1988, 1992; Cavalli-Sforza and Piazza, 1975; Hunley and Long, 2005; Long and Kittles, 2003; Urbanek et al., 1996). We chose to examine this model in Northern Island Melanesia in order to test whether the geography of the region might result in tree-like evolution. If the data had fit the population fissions model well, we would have learned a great deal about the order and timing of population fissions associated with colonization and subsequent population events in the region. However, we found that none of the data fit this model. More specifically, we found that none of the gene or language trees could be fit to the pattern of within- and between-population genetic variation estimated from three types of genetic data. This indicates a substantial lack of treeness in the genetic
Inferring Prehistory from Genetic, Linguistic, and Geographic Variation
Figure 9.5 Multidimensional scaling plots for genetic, linguistic, and phenotypic distances. Stress values are indicated in the lower left portion of the plot. Circles, Bougainville; squares, New Britain; triangles, New Ireland. Papuan-speaking populations are shaded. This plot differs somewhat from the Y SNP MDS plot in chapter 5, as would be expected, given the small number of loci used here.
151
core studies in northern island melanesia
variation in this region. As mentioned, the trees may fit poorly because the genetic data contain insufficient information to capture existing treeness, but a better explanation, at least for the mtDNA data, is that population genetic evolution has generally not been tree-like in the region. The most likely reason for this lack of treeness is that, even if the initial and subsequent population radiations into the region were tree-like, newly formed groups were not subsequently isolated and therefore did not evolve independently of one another. Archeological studies document inter-island trade both before (Allen, 1996; Leaveseley et al., 2002; Summerhayes and Allen, 1993) and following (chapter 2 and references therein) the apparent Austronesian intrusion into Island Melanesia ~3,300 years ago. Undoubtedly, genetic exchange accompanied this trade. The lack of genetic partitioning between Papuan- and Austronesian-speaking populations also suggests substantial genetic exchange between these groups on each island.
Isolation by Distance Sewa1l Wright (1943) introduced the expression “isolation by distance” to describe the tendency of geographically proximate populations to exchange genes. Kimura and Weiss (1964) and Malécot (1948) subsequently extended the model to deal with migration in one and two spatial dimensions. These last two versions of the model assume homogeneity of effective population sizes, migration rates, and migration distances. Under these assumptions, both versions predict an exponential, monotonic decline in genetic similarity with increasing geographic distance. Because these assumptions are routinely violated in human populations, it can be difficult to meaningfully interpret observed correspondences that occur between genetic and geographic distances in many world regions. Though we identified several statistically significant correlations with geography in Northern Island Melanesia, these correlations were weak, and frequently were caused by just one or two populations. Violations of the assumptions of the model may contribute to its lack of fit, e.g., differences in effective population sizes as well as temporal and spatial variation in migration rates and distances between populations should contribute to the weakness of the observed correlations. For example, table 9.1 indicates substantial differences in the amount of within-population genetic variation for all genetic systems. Migration rates also vary dramatically across the region (Friedlaender, 1975). Also, we suspect that population movement has played a considerable part in preventing or obscuring genetic, linguistic, and geographic correspondence. For example, Austronesian intrusion
152
likely displaced Papuan-speaking groups, e.g., the Austronesian-speaking Tolai may have caused the speakers of the Papuan Kaket and Mali (Baining) languages to move inland from their previous coastal location (Capell, 1967). There also has been a southern displacement of the Kuot in New Ireland by northern Austronesian speakers (Lindström, personal communication). In addition, the Papuan-speaking Anêm recently moved to a more western location. No doubt more recent outside contact also affected population locations and interactions among indigenous groups (Friedlaender, 1975; Oliver, 1991). So, in the end we can reject both simple models of population history. Treeness may never have existed in the region, or it may have been disrupted by indigenous population processes, including contractions, expansions, fusions, fissions, genetic exchange, and movements. These processes also disrupt treeness at larger geographic scales. For example, Hunley and Long (2005) showed that population genetic evolution has not been tree-like in Native North America. Their study concluded that any treeness that may have existed in North America was disrupted by a combination of local mate exchange and population movements, much like those observed in Island Melanesia. This pattern of reticulate exchange augmented by population movement is no doubt ubiquitous in human groups, and should not be ignored. In fact, we expect that these processes will disrupt tree-like evolution at all but the most extreme geographic scales (Ramachandran et al., 2005) or in newly colonized, geographically heterogeneous regions in which insufficient time has elapsed to erase an early history of fissions (Cavalli-Sforza et al., 1988, 1994). At the outset, we considered the possibility that Island Melanesia might be one such region, but clearly the island structure of the region and the ruggedness of the topography within islands have not resulted in tree-like evolution. Nonetheless, despite that fact that evolution has not been tree-like, our results indicate that seas have formed a relative barrier to genetic exchange. For example, the ΦRT values for each genetic system indicate that a substantial portion of genetic variation is apportioned between islands. This result might not have been anticipated given that a great deal has been made of the ease of seagoing travel in this region (Irwin, 1992).
What Do Our Analyses Tell Us about the Prehistory of Island Melanesia? In spite of the lack of fit of either model, the results of our analyses indicate that the genetic, linguistic, and phenotypic data contain important information about the prehistory of the region. For example, the MDS plots (figure 9.5) of autosomal SNP, mtDNA, and linguistic
Inferring Prehistory from Genetic, Linguistic, and Geographic Variation
data indicate that Austronesian-speaking populations are relatively similar to one another both genetically and linguistically. These plots also emphasize the relatively large genetic and linguistic differences between Papuanspeaking populations. Additional analyses confirmed significantly greater genetic distances between Papuanthan Austronesian-speaking populations for all genetic systems. For example, the average number of pairwise mtDNA sequence nucleotide differences between Papuanspeaking populations included in this study is 5.55. This compares with the Austronesian-speaking population value of 1.63. This difference is significant at the 0.0005 level (independent sample t-test). The Papuan– Austronesian difference is equally dramatic for the autosomal SNPs and Y chromosome STRs. These results are consistent with an early migration (or migrations) of proto-Papuan speakers followed by the formation of non-tree like but substantial population structure. Because of their relatively long tenure in the region, Papuan-speaking populations have diverged substantially both genetically and linguistically. Our results are also consistent with a more recent migration of Austronesian speakers into Island Melanesia, followed again by the formation of substantial but non-treelike population structure. However in this case, insufficient time has elapsed to produce the magnitude of genetic and linguistic difference observed for the Papuan-speaking populations. Our results also indicate that substantial genetic and linguistic exchange occurred between Papuan- and Austronesian-speaking populations following the more recent arrival of the latter. For example, AMOVA and FST results indicate that essentially none of the genetic or phenotypic variation is apportioned between these two linguistic groups. Nonetheless, this exchange has been insufficient to erase the genetic and linguistic signatures of distinctive Papuan and Austronesian migrations. One additional interesting finding of this study is that skin color variation is different between islands, but is essentially uniform across populations within islands (though see Norton et al., 2006, for important exceptions). It is possible that a combination of within-island migration and natural selection for dark skin have produced relative inter-linguistic homogeneity in skin color, while maintaining dramatic inter-island differences. Finally, while our results paint a picture of Island Melanesian population history that is consistent with other genetic and archeological data, the patterns of variation in the different types of genetic data employed in this study are inconsistent, particularly in the Y chromosome STRs. For example, the MDS plot in figure 9.5C indicates that Y chromosome STR variation is patterned neither geographically nor linguistically. We strongly suspect that this absence of pattern reflects a lack of information in these
genetic data (i.e., the Y STR dataset). Nonetheless, we must consider the possibility that the Y chromosome data reflect aspects of population history not captured by other loci used in this and other studies. Higher-resolution genetic data, in combination with archeological, linguistic, and ethnographic data are required to address this issue, and to more fully explore the prehistory of the region.
References Allen J. 1996. The pre-Austronesian settlement of Island Melanesia. In: WH Goodenough, editor. Prehistoric settlement of the Pacific. Philadelphia, PA: American Philosophical Society. pp 11–27. Barbujani G, Sokal R, and Oden N. 1995. Indo-European origins: A computer-simulation test of five hypotheses. American Journal of Physical Anthropology 96: 109-132. Capell A. 1967. A lost tribe in New Ireland. Mankind 6: 499–509. Cavalli-Sforza LL, Contributors Menozzi P, Piazza A. 1994. The history and geography of human genes. Princeton: Princeton University Press. Cavalli-Sforza LL. 1997. Genes, peoples, and languages. Proceedings of the National Academy of Sciences USA 94: 7719–24. Cavalli-Sforza L, Edwards A. 1964. Analysis of human evolution. Proceedings of the 11th International Congress of Genetics 2: 923–33. Cavalli-Sforza LL, Piazza A. 1975. Analysis of evolution: Evolutionary rates, independence and treeness. Theoretical Population Biology 8: 127–65. Cavalli-Sforza LL, Piazza A, Menozzi P, Mountain J. 1988. Reconstruction of human evolution: Bringing together genetic, archaeological, and linguistic data. Proceedings of the National Academy of Sciences USA 85: 6002–6. Cavalli-Sforza LL, Minch E, Mountain JL. 1992. Coevolution of genes and languages revisited. Proceedings of the National Academy of Sciences USA 89: 5620–4. Dunn M, Terrill A, Reesink G, Foley RA, Levinson SC. 2005. Structural phylogenetics and the reconstruction of ancient language history. Science 309: 2072–5. Eller E. 1999. Population substructure and isolation by distance in three continental regions. American Journal of Physical Anthropology 108: 147-159. Excoffier L, Smouse P, Quattro J. 1992. Analysis of molecular variance inferred from metric distances among DNA haplotypes: Application of human mitochondrial DNA restriction data. Genetics 131: 479–91. Felsenstein J. 2005. PHYLIP (Phylogeny Inference Package) version 3.6. Seattle, Washington DC: Distributed by the author. Department of Genome Sciences, University of Washington. Friedlaender J. 1971. Biological divergences in south-central Bougainville: An analysis of blood polymorphism gene frequencies and anthropometric measurements
153
core studies in northern island melanesia
utilizing tree models, and a comparison of these variables with linguistic, geographic, and migrational “distances.” American Journal of Human Genetics 23: 253–70. Friedlaender J. 1975. Patterns of human variation: The demography, genetics and phenetics of Bougainville Islanders. Cambridge, MA: Harvard University Press. Fuselli S, Tarazona-Santos E, Dupanloup I, Soto A, Luiselli D, Pettener D. 2003. Mitochondrial DNA diversity in South America and the genetic history of Andean highlanders. Molecular Biology and Evolution 20: 1682–91. Holm S. 1979. A simple sequentially rejective multiple test procedure. Scandanavian Journal of Statistics 6: 65–70. Hunley K, Long JC. 2005. Gene flow across linguistic boundaries in Native North American populations. Proceedings of the National Academy of Sciences USA 102: 1312–7. Irwin G. 1992. The prehistoric exploration and colonization of the Pacific. Cambridge: Cambridge University Press. Kimura M, Weiss G. 1964. The stepping stone model of population structure and the decrease of genetic correlation with distance. Genetics 49: 561–76. Kruskal JB, and Wish M. 1978. Multidimensional Scaling. Beverly Hills: Sage Publications. Leaveseley M, Bird M, Fifield L, Hausladen P, Santos G, Tada MD. 2002. Buang Merabak: Early evidence for human occupation in the Bismarck archipelago, Papua New Guinea. Australian Archaeology 54: 55–7. Long JC, Kittles RA. 2003. Human genetic diversity and the nonexistence of biological races. Human Biology 75: 449–71. Malécot G. 1948. Les Mathématiques de l′hérédité. Paris, France: Masson et Cie. Meyer S, Weiss G, Haeseler Av. 1999. Pattern of nucleotide substitution and rate heterogeneity in the hypervariable regions I and II of human mtDNA. Genetics 152: 1103–10. Nei M. 1987. Molecular evolutionary genetics. New York: Columbia University Press. Norton HL, Friedlaender JS, Merriwether DA, Koki G, Mgone CS, and Shriver MD. 2006. Skin and hair pigmentation variation in Island Melanesia. American Journal of Physical Anthropology 130: 254–68. Oliver D. 1991. Black islanders: A personal perspective on Bougainville 1937–1991. Honolulu, Hawaii: University of Hawaii Press. Pavlides C, Gosden C. 1994. 35,000-year-old sites in the rain forests of west New Britain, Papua New Guinea. Antiquity 69: 605–10. Ramachandran S, Deshpande O, Roseman CC, Rosenberg NA, Feldman MW, and Cavalli-Sforza LL. 2005.
154
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 USA 102: 15942–7. Relethford JH. 1996. Genetic drift can obscure population history: Problem and solution. Human Biology 68: 29–44. Relethford JH. 2004. Global patterns of isolation by distance based on genetic and morphological data. Human Biology 76: 499-513. Relethford J, Blangero J. 1990. Detection of differential gene flow from patterns of quantitative variation. Human Biology 62: 5–25. Relethford J, Crawford M, Blangero J. 1997. Genetic drift and gene flow in post-Famine Ireland. Human Biology 69: 443–65. Rousset F. 1997. Genetic differentiation and estimation of gene flow from F-statistics under isolation by distance. Genetics 145: 1219–28. Saitou N, Nei M. 1987. The neighbor-joining method: A new method for reconstructing phylogenetic trees. Molecular Biology and Evolution 4: 406–25. Schneider S, Roessli D, Excoffier L. 2000. Arlequin ver. 2000: A software for population genetics data analysis. Geneva, Switzerland: Genetics and Biometry Laboratory. Serre D, Pääbo S. 2004. Evidence for gradients of human genetic diversity within and among continents. Genome Research 14: 1679–85. Slatkin M. 1995. A measure of population subdivision based on microsatellite allele frequencies. Genetics 139: 457–62. Slatkin M, Maddison WP. 1990. Detecting isolation by distance using phylogenies of genes. Genetics 126: 249–60. Summerhayes G, Allen J. 1993. The transport of Mopir obsidian to late Pleistocene New Ireland. Archeological Oceania 28: 145–9. Tamura K, Nei M. 1993. Estimation of the number of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees. Molecular Biology and Evolution 10: 512–26. Urbanek M, Goldman D, Long JC. 1996. The apportionment of dinucleotide repeat diversity in Native Americans and Europeans: A new approach to measuring gene identity reveals asymmetric patterns of divergence. Molecular Biology and Evolution 13: 943–53. Wickler S, Spriggs M. 1988. Pleistocene human occupation of the Solomon Islands, Melanesia. Antiquity 62: 703–6. Williams-Blangero S, Blangero J. 1992. Quantitative genetic analysis of skin reflectance: A multivariate approach. Human Biology 64: 35–49. Wright S. 1943. Isolation by distance. Genetics 28: 114–38.
part iii Regional Studies and Conclusion
This page intentionally left blank
10 Animal Translocations, Genetic Variation, and the Human Settlement of the Pacific Elizabeth Matisoo-Smith
Introduction Studying the biological variation of the animals and plants transported by humans as they moved into and through the Pacific has become an attractive alternative approach to reconstructing human population origins, migrations, and interactions in the region. They provide a proxy for identifying human origins and mobility. This commensal approach has, to date, focused primarily on the animals associated with the Lapita expansion, the pig, dog, chicken, and rat. The results show the clear value of the approach and its potential for application in other regions and with other organisms. They also raise new questions regarding animal/human relationships in the Pacific and elsewhere.
Animal Translocations in the Pleistocene The introductions of the dog, pig, chicken, and rat to Remote Oceania were claimed to be associated with the “transported landscapes” of the Lapita cultural complex (Kirch, 1997: 47; 2000: 111), but as Summerhayes (chapter 2) and others point out, Near Oceania provides the earliest evidence of animal translocation (Flannery and White, 1991; Heinsohn, 1998, 2001; White, 2004). The Pleistocene introduction of the Northern Common Cuscus (Phalanger orientalis) from New Guinea to New Ireland around 20,000 BP is perhaps the earliest evidence for translocation worldwide (Grayson, 2001). Within a few thousand years of this initial introduction, this species was introduced to other islands in the Bismarck Archipelago. Other species native
to New Guinea were also transported throughout the Bismarcks and Admiralties during the late Pleistocene, including rats (Rattus mordax and Rattus praetor along with the black-tailed Melomys rat, Melomys rufenscens), another species of cuscus (the Admiralty Cuscus, Spilocuscus kraemeri), and a bandicoot (Echympera kalubu) (Flannery, 1995; White et al., 2000; Summerhayes, 2003; White, 2004; Leavesley, 2005). These early animal translocations could hardly be described as evidence for “domestication” in the traditional sense, but they do provide clear indicators of human manipulation or management of the environment. While the study of such early animal translocations provides new evidence of human interaction across Near Oceania, they are potentially more important in developing hypotheses concerning the arrival of animals from outside the immediate Sahul region, particularly regarding questions about “domestication” and alternative animal– human relationships in the Pacific.
Post-Pleistocene Animal Introductions Pig Introductions Among the recent attempts to reconstruct the late Pleistocene/early Holocene movement of animals around the archipelagos of Near Oceania and Island Southeast Asia, those concerning the initial introduction time of the pig to New Guinea from Southeast Asia have been particularly contentious (Spriggs, 1997; Bulmer, 1975, 1982; Goreki et al., 1991; Allen, 2000; Green, 2000; Bellwood
157
regional studies and conclusion
and White, 2005). Since pigs are generally considered domesticated animals, their presence in archeological sites has often been interpreted as an indication of agricultural societies, with pigs acting as a particularly useful storage for agricultural surplus. As Summerhayes (chapter 2) discusses, the argument for pigs being in the New Guinea highlands as early as the late Pleistocene/early Holocene is based on reports of a few pig bones and isolated teeth from the sites of Kafiavana, Yuku, and Kiowa (Bulmer, 1975, 1982). Goreki et al. (1991) and Swadling (1997) argue for pigs on the coast around 5,000–6000 BP, which would also be preLapita. Because of the association with intensive agriculture in particular, both the direct and indirect evidence for pig in early Holocene contexts in the highlands takes on special significance (e.g. the Kuk swamp pig wallows and stake holes identified by Golson and Hughes, 1980). Many question the reliability of both the dates for the material and the contexts in which they were discovered (Hedges et al., 1995, Spriggs, 1997). The limited number of early pig remains in New Guinea, and their absence in Australia, suggests that if pigs were present in late Pleistocene/Early Holocene New Guinea, they were not there in large numbers. If pigs were introduced early, some have concluded that a date of around 6,000 BP is probably the earliest acceptable timeframe (Allen, 2000; Green, 2000). However, I suggest the controversy surrounding the introduction of pig has become something of a preoccupation, and it is perhaps more appropriately viewed within the developing context of other animal dispersals in the region. By the time of even the earliest proposed pig dates in the New Guinea highlands, Pacific peoples had been moving animals around and altering their environments for at least 10,000 years, since 20,000 BP. We know that interactions increased during the early Holocene and more animals were being moved around the region. The northern pademelon (Thylogale browni), another wallaby-like herbivore, was introduced to New Ireland by 8,400 BP (White et al., 1991) and, by 8,000 BP, there is evidence of translocation of New Guinea fauna (specifically a wallaby belonging to the Dorcopsis genus) westward across the Wallace Line to the island of Gebe, in the Moluccas (Bellwood et al., 1998). Why would picking up a pig (or more likely, a piglet) or two from west of the Wallace Line and bringing it back to New Guinea be such a different and unlikely event? In addition, why must transporting a pig be interpreted as evidence of “domestication” when moving around a wallaby or a cuscus is not? There is little doubt that pigs were a part of the Lapita Cultural Complex and it is in Lapita sites that pig remains are found in any numbers (though still not large numbers). It is also clear that the archeological evidence for the initial Lapita expansion into Remote Oceania includes pig bones, and that they may be part of its full-blown
158
“Neolithic” signature in some locations. Pig remains appear in small numbers in most Lapita sites in Near and Remote Oceania; however, they have not been recorded in Lapita sites in New Caledonia (Sand, 1996) and large numbers of pig bones do not generally occur in sites until late in the Lapita chronology. As detailed in the next section, whether the Lapita-associated pigs came from New Guinea or Island Southeast Asia is not clear and, given the argument for early pig introductions to New Guinea, it is possible that their origin may be as complicated as the origin of the Lapita people themselves.
Molecular Evidence for Pig Origins and Introductions In an attempt to determine the location(s) of pig domestication and identify the relationships of pig populations worldwide, Larson et al. (2005) analyzed 663 bp of the mitochondrial control region from 686 wild and domestic pigs (Sus scrofa). A few S. scrofa samples from the Pacific (Hawaii, Vanuatu, the Admiralty Islands, and New Guinea) as well as Island Southeast Asian samples of the species Sus celebensis, S. barbatus, and S. verrucosus were included in the analyses. The study suggests there may have been multiple centers for pig domestication, with the most basal lineages of Sus scrofa (and therefore its origin) in western island Southeast Asia. It also confirmed previous analyses (Okumura et al., 1996; Watanobe et al., 2001; Robins et al., 2003) that showed very distinct Asian and European Sus scrofa clades. Once the recently introduced European lineages in the Pacific were removed from the analyses, all of the Pacific pigs fit within the greater Asian pig clade, but as a monophyletic branch, distinct from the other Asian scrofa groups. The only non-Pacific population that was also found in this “Pacific clade” was a sample from Halmahera in Wallacea. No specific origin for the “Pacific clade” could be clearly determined, leading the authors to suggest “either the existence of indigenous S. scrofa in Wallacea or an early human-mediated introduction from elsewhere in ISEA currently not sampled in our study” (Larson et al., 2005: 1621). Significantly, the analyses were inconsistent with the view that Pacific pigs were derived from either the Taiwanese wild boar or from the indigenous Sulawesi wild boar (Sus celebensis) as had been suggested by Groves (1981). Like many genetic studies, sample origin and distribution is a potential problem in this study. Most of the Pacific pig samples came from historic museum collections and, given the impact that the introduction of modern pig breeds has had in the Pacific in the years since European arrival (Hide, 2003), may not necessarily represent the indigenous prehistoric pigs introduced to the Pacific. However, recent analyses of pig remains that are clearly associated with undisturbed and undoubtedly prehistoric archeological sites from both Near and Remote Oceania
Animal Translocations, Genetic Variation, and the Human Settlement of the Pacific
(unpublished data from our University of Auckland labs) also fit within this monophyletic Pacific clade, suggesting that the museum pig samples are at least maternally related to the ancient pig populations introduced by Pacific peoples. The question still begs: Where did these Pacific pigs come from? Are they related to early Papuan pigs from Near Oceania, or did “Lapita peoples” or others introduce them from elsewhere in Island Southeast Asia? The connection to Halmahera is intriguing. Bellwood and White (2005: 381) pointed out that “a substantial program of archaeology in Maluku Utara has revealed that pigs were completely absent here until after 3500 years BP” (N.B. Maluku Utara Province in northern Wallacea includes Halmahera, just to the west of the Bird’s Head of New Guinea). Therefore, they argue “If pigs were taken initially to New Guinea from Halmahera, they must also postdate 3500 BP there …” This does suggest that pig introductions to Halmahera from elsewhere were most likely associated with the “Neolithic expansion” and that if pigs were introduced to New Guinea prior to 3,500 BP, they must have come through some other route. However, the lack of pig in Halmahera prior to 3,500 BP does not allow one, on that basis alone, to reject the early New Guinea pig dates. It also means other sites in Wallacea and Island Southeast Asia should be examined for potential source populations and alternative routes of dispersal. Another interesting alternative is that the Halmahera pig was introduced there from New Guinea after 3,500 BP. The origin of the New Guinea pigs therefore still remains unclear. As mentioned, the molecular data derived from the Pacific pig samples make it clear they do not come from Taiwan. Unfortunately, no pig samples from the Philippines were included in the genetic study, although our research group is currently analyzing a significant number of samples from throughout the archipelago. Unfortunately, given the limited number and lack of diagnostic features of the early New Guinea pig remains, we cannot tell on morphological grounds alone whether or not they resemble the later Lapita-associated pig remains. Perhaps the practice of large-scale agricultural production that begins/arrives after 3,500 BP allowed for the rapid and dramatic increase in a small pig population already present in New Guinea, resulting in the clear and continuous appearance in the archeological record from then on. Alternatively, as suggested by Spriggs (1997: 95), it is certainly possible that there were two introductions of pigs into Near Oceania, with wild pigs introduced during the early Holocene to New Guinea and domestic pigs later, around 3,500 BP, to the Bismarck Archipelago and from there both back to New Guinea proper and out into the rest of Oceania. As discussed below, genetic analyses of other “Lapita-associated” animals, the dog and the rat, also indicate that multiple and pre-Lapita introductions to the Pacific were possible if not likely. However, given
the unique animal/human relationships in the Pacific, particularly pre-Lapita, the traditional morphological markers of “domesticated” animals may not occur until they are truly dependent upon and manipulated by humans— a situation that in some cases may never have happened in the Pacific. Therefore, molecular evidence may be the only means of addressing these relationships. DNA may still remain in the “early” pig teeth and ancient DNA analyses could then provide an answer to this question in the future.
Dog Introductions Archeological evidence for the introduction of the dog into the Pacific region is limited. The earliest archeological evidence for dingoes in Australia is about 3,500 BP (Gollan, 1985) and about the same for the first appearance of dogs in Timor (Bellwood, 1997). Like the pig situation, there is some suggestion that dogs may be present possibly as early as 6,000 BP in the New Guinea highlands at Kiowa (Bulmer, 1975). Once again, the early appearance of dogs in Near Oceania is not without debate, primarily because early dog remains are rare and generally consist of individual teeth. Explanations based on contemporary ethnographic evidence may suggest why dog remains are rare in archeological sites in New Guinea (see Bulmer, 2001, for a discussion and review). For example, Ralph Bulmer suggested that among the Kalam, since dogs were not eaten but used primarily in hunting and were said to possess spirits that linger after death, their remains were thrown into rivers, which were passages to the underworld. Similarly, dogs were considered to carry strong magic and, therefore had to be buried away from villages, again making archeological recovery unlikely (Bulmer, 1967, 1976). With the exception of the dog remains recorded from the upper parts of the disturbed midden at Akari (Swadling, 1997; Green, 2000; Bulmer, 2001), dog remains do not appear on the New Guinea coast until much later. Irwin (1977) records dog jaw fragments from Mailu possibly as early as 2,000 BP, and Sue Bulmer (1978) records the burial of a puppy under a house floor from the Taurama site dated, based on pottery style, to between 1,000 and 2,000 BP. Lilley (1986) reports dog bone from Sio, in the Vitiaz Strait region, dated to approximately the same period. Though it is generally suggested that dogs were part of the Lapita introduction, when the archeological reports are examined, even Lapita sites in Near Oceania do not provide conclusive evidence for dog remains. Kirch (1988) reports dog for the Talepakemelai (ECA) site in Mussau, but in a more recent publication on the archeological investigations in the Mussau Islands, there is no mention of any identified dog remains (Kirch, 2001). A small number of pig bones were identified from the early sites, but
159
regional studies and conclusion
otherwise only “medium mammal” bone is mentioned. Bulmer (2001) lists dog bones from Kamgot (ERA); however, a review of the faunal lists for the site shows only a single dog tooth, though there is some unidentified medium mammal listed (unpublished faunal lists provided by G. Summerhayes). Bulmer (citing Gosden et al., 1989) also states that there is dog bone from Apalo (FOJ) a beach midden site on Kumbun Island in the Arawe group, but no details are given. Kirch and Yen (1982) report dog (along with pig, chicken, and rat) in the early parts of the archeological sequence from Tikopia, which dates from 2,800 BP. It turns out that the material from the early phase represents one dog bone (Nagaoka, 1988). If the context is without question, this would be the earliest evidence of dogs in the Pacific Islands, but it is hardly conclusive. Interestingly, dog does not appear in the middle of the sequence but reappears in small amounts from 1,200 AD, a period represented by the Polynesian occupation phase. Flannery et al. (1988) report that dogs, along with pigs, were present on Buka by 1,860 BP. Further east in Remote Oceania, dogs are similarly elusive at early dates. To date, no Lapita dog remains have been identified in Vanuatu (Stuart Bedford, personal communication, 2005) or New Caledonia (Sand, 1996; White et al., 2000), and in Fiji dog “may be present” in Lapita phases at Yanuka and Naigani, but does not appear until about 1,000 BP in Lakeba, when pigs also turn up (Clark and Anderson, 2001). Dog bone, if present, is rare in Lapita sites in Tonga (Burley, 1998). Overall, the archeological evidence for Lapita dog or for any dog prior to 2,000 BP in the Pacific is weak to non-existent, other than for the dingo evidence in Australia at 3,500 BP.
Molecular Evidence for Dog Origins and Introductions Molecular genetic studies on dog evolution have only recently begun. The first mtDNA analysis of modern dogs was published in 1997 (Vilà et al., 1997), and since that time several papers addressing the question of origins, both of dog domestication and of particular dog populations, have followed (Leonard et al., 2002; Savolainen et al., 2002, 2004). While initial estimates of divergence of dogs from wolves suggested domestication around 100,000 BP, recent estimates of around 15,000 BP fit much more closely with archeological estimates for early animal domestication (Savolainen et al., 2002). Interestingly, the patterns of phylogeographic variation suggest the origin of the domestication of dogs is East Asia. Regarding the Pacific, the first study by Vilà et al. (1997) included a sample of four dingoes and two New Guinea singing dogs in a total of 140 domestic dogs representing 67 breeds. All six Australian/New Guinea samples had the same sequence for the 261 bp of HVR1 sequenced and were identified as belonging
160
to haplotype D18. This haplotype belonged to a clade that included other “ancient” dogs like basenjis and greyhounds. The only other dogs that carried the D18 haplotype were a Siberian husky and a crossbreed. Concerning the origin of the Australian dingo, Savolainen et al. (2004) conducted further studies of 582 bp of control region sequence in 211 dingoes collected from throughout Australia. A total of 20 haplotypes were identified in dingoes, but they differed by a maximum of only two substitutions. Only two of those 20 haplotypes were shared with any other dog breeds, with the remaining 18 being unique to dingoes. The main haplotype found in 56% of dingoes (A29) was one of the two types also found in other dogs. All of the non-dingo A29 dogs came from East or Southeast Asia (2/18 from East Siberia, 4/96 from Japan, 2/7 from Indonesia), the American Arctic (6/25) or New Guinea (1/2). In a minimum spanning network analysis, all other dingoes and most other Asian dog lineages radiated out in a star-pattern from the A29 lineage. This suggests that all dingo lineages, along with most other Asian domestic dog lineages, are derived from the A29 haplotype, most likely from an ancient domestic East or Southeast Asian dog. This contradicts suggestions, based on morphological grounds, of an Indian origin for the dingo (Corbett, 1985; Gollan, 1985). Interestingly, one of the two New Guinea singing dog sequences was also in the A29 group, and the other was a unique sequence that was, again, just one substitution from the A29 haplotype. Molecular dating of phylogeny split times is problematic. Given the mean distance of 0.190 substitutions (SD = 0.007) for all other dingo sequences to the A29 lineage, and a mutation rate based on a wolf–coyote split of 1 MYA, the probable time of introduction of the dingo to Australia was estimated to be 4,600 to 5,400 BP. However, with the possibility of a wolf–coyote split dating to 2 MYA, the calculated dates for dingo introduction could be between 4,600 and 10,800 BP (Savolainen et al., 2004). Putting aside the numerous debates and issues about estimates of mutation rates and calculations of dates of divergence, the suggested timing for the introduction of the dingo, and most likely the New Guinea singing dog, sits right in that familiar early- to mid-Holocene timeframe identified for other animal introductions into the region. Besides the New Guinea singing dog, a number of shorter d-loop sequences from archeological dog remains from the Pacific were included in the analyses of Savolainen et al. (2004). In most cases, the samples came from confirmed, pre-European archeological contexts and included samples from Hawaii (n = 4), the Cook Islands (n = 2), and New Zealand (n = 13). Two haplotypes were identified in those samples. One was a common lineage (identical to D2 in Vilà et al., 1997) that was also found in
Animal Translocations, Genetic Variation, and the Human Settlement of the Pacific
Asian-derived breeds (Chinese crested, chow chow, and Tibetan terrier) and a basenji. Interestingly, further analyses of archeological dog remains in our laboratories at the University of Auckland have identified the same D2 haplotype in ancient DNA extracted from the Taurama puppy remains from New Guinea, dated to between 1,000 and 2,000 BP, and in a modern dog bone from the village of Ban Chiang in Thailand. The other haplotype, A75, was found in all three Polynesian archeological dog populations, but in only two Indonesian dogs from the 600+ domestic dogs studied worldwide. From the molecular evidence, we suggest that there were at least three dog introductions to the Pacific— the A29 haplotype to Australia and New Guinea, the D2 haplotype found on the coast of New Guinea and out into Remote Oceania, and the A75 haplotype, which has only been identified in Indonesia and Polynesia. The estimated dates of divergence suggest that the A29 haplotype may have been a mid-Holocene, pre-Lapita introduction, with a later introduction of the D2 and A75 coming from Southeast Asia. Dogs with these two haplotypes could well have been carried out into Remote Oceania by the Lapita people, but it is possible that at least one of these lineages was introduced later, around 1,000–2,000 BP, when there is an increase in the numbers of dogs in the archeological record. Analyses of aDNA from dated archeological dog remains, and in particular, any Lapitaassociated confirmed dog remains, might allow us to test this possibility at some point in the future.
Rat Introductions While there may be debates about the definition of “domestication” and its applicability to different animal introductions in the Pacific, few would argue that the introductions of wallabies, dogs, pigs, and chickens were unintentional. The transport of rats around the Pacific, on the other hand, is often compared to the transport of European rodents on historic sailing vessels and has therefore generally been assumed similarly unintentional (Taylor et al., 1982; Kirch, 1997). However, I have argued that rat introductions to islands in Polynesia, and most likely in Remote Oceania in general, were intentional because it was an important food item (Matisoo-Smith, 1994). The Pacific rat, Rattus exulans, is the third most widely distributed rat worldwide, after Rattus rattus and Rattus norvegicus. It generally does not like wet conditions and has not been recorded stowing away on any European sailing ships during historic times. Its origin is thought to be around the Isthmus of Kra (Thailand) and its distribution beyond Southeast Asia is limited to the islands associated with Austronesian dispersal—from the Andaman Islands through Near Oceania and out into the extremes of Remote Oceania (Tate, 1935; Roberts, 1991). Because of
this distribution, the introduction of R. exulans in the Pacific has generally been associated with Lapita. This association, however, has not been demonstrated with certainty and R. exulans is not the first or only rat moved around the Pacific prehistorically. The case of the large spiny rat, Rattus praetor, makes the R. exulans situation appear somewhat less singular. White et al. (2000) provide an excellent review of the distribution of R. praetor across the Pacific. It is listed as indigenous to New Guinea (Flannery, 1995) and appears on New Ireland in late Pleistocene deposits, from 13,000 BP onwards, in Panakiwuk cave (Marshall and Allen, 1991), but only appears in the top two horizons at Balof, which date to 3,000–4,000 BP. R. praetor is reported from Pamwak on Manus at levels dated to between 9,000 and 5,000 BP. Elsewhere in the Pacific, including the Solomon Islands, it appears to be a much later introduction in the range of 2,000–3,000 BP. R. praetor remains have been recovered in the earliest archeological sites in Vanuatu and in Fiji, but it was never introduced to New Caledonia (White et al., 2000). Though the arrival of R. exulans and its distribution through the Pacific is generally associated with Lapita settlement, the precise dating of its introduction to Near Oceania is problematic, as sites from the period just prior to Lapita arrival are rare. Three exulans bones were found at Panakiwuk, New Ireland, in layers dating to 8,000– 13,000 BP, but given the assumed Lapita connection, they were thought to be in such old layers because of site disturbance (Marshall and Allen, 1991). One problem in assessing both the number of rat species present and their arrival date in a site is the recovery technique employed by archeologists, especially the use of large screen sizes for sieving excavated material. As Nagaoka (1988) points out, large screen sizes may result in the loss of small faunal remains so animals such as rats may be under- or unrepresented. Another difficulty is the species identification of rodent remains. Generally, rodent remains in Pacific contexts are classified by size into small, medium, and large. Small rats have been assumed to be R. exulans, medium to be R. praetor, and large, R. mordax or some other rodent species. However, our molecular studies have shown that there is significant overlap in size of R. exulans and other rat species in the Pacific (MatisooSmith and Allen, 2001), and that even with complete fresh carcasses, morphologically based species identifications in rats in Near Oceania is problematic (Robins et al., manuscript submitted for publication).
Molecular Evidence for Rat Origins and Introductions Our study of mtDNA variation in Pacific rat species began with research on extant populations of R. exulans from central and east Polynesia (Matisoo-Smith, 1994; Matisoo-Smith et al., 1998). We chose the rat to begin our
161
regional studies and conclusion
commensal studies for several reasons. First, it was the one Pacific commensal that was successfully established on nearly all islands and archipelagos in Polynesia that have any evidence of prehistoric human occupation. The distributions of the dog and pig are much more irregular in both time and space, and while the chicken appears to be more widely distributed across the Pacific than pigs and dogs, it is not present in pre-colonial New Zealand. Second, R. exulans is a clear, reproductively isolated, species from the later European-introduced rats, Rattus norvegicus and Rattus rattus, and can be distinguished both morphologically and molecularly from other rat species. Polynesian dogs, pigs, and chickens, on the other hand, belong to the same species as European-introduced dogs, pigs, and chickens (Canis familiaris, Sus scrofa, and Gallus gallus), and the populations quickly interbred, causing potential distributional confusion. Finally, R. exulans had a significant role in Polynesian mythology and oral tradition, particularly in New Zealand and other East Polynesian locations, where it was often represented as a fellow traveler (Matisoo-Smith, 1994) (see figure 10.1). The results of this first molecular study confirmed that the patterns of variation in current exulans populations are useful in identifying human settlement and interaction patterns within Polynesia (Matisoo-Smith et al., 1998). We then attempted to see if we could extract mtDNA
Figure 10.1 Carving from the meeting house Tanenuiarangi, University of Auckland depicting Ruanui, captain of one of the founding canoes recognized by New Zealand Maori, shown with the kiore (Rattus exulans) traveling on his shoulder. See color insert.
162
from archeological remains, thereby addressing the issue of lineage extinction and dating the appearance of specific lineages (Matisoo-Smith et al., 1999). Once we realized we could indeed extract DNA from archeological exulans remains, the door was open to expand our study in both time and space. To address the contentious issue of Lapita origins, we expanded our study to focus on how Polynesian and other Remote Oceanic samples related to those from Near Oceania and Southeast Asia (Matisoo-Smith and Robins, 2004). We analyzed 240 bp of mtDNA d-loop from 131 exulans samples from Island Southeast Asia and the Pacific. The analysis included sequences obtained from bones from museum collections (n = 33), archeological sites (n = 87), and some fresh tissue (n = 11). Phylogenetic analyses identified three very distinct haplogroups. Haplogroup I included samples from the Philippines, Borneo, and Sulawesi. Haplogroup II lineages were distributed from the Philippines, through the Moluccas, New Guinea, and the Solomon Islands. The third haplogroup had a distribution virtually confined to Remote Oceania (see figure 10.2). We were surprised to find, given the Lapita settlement scenario that stresses relationships between Near and Remote Oceania, that our Near and Remote Oceanic exulans samples were so different. In fact, there were no Near Oceanic mtDNA lineages in Remote Oceania and visa versa, but both lineages were present in Halmahera (Wallacea) to the west. Studies of morphological variation of Pacific exulans (Tate, 1935; Motokawa et al., 2004) report a similar lack of continuity between Near and Remote Oceania. The apparent lack of relationship we identified may suggest distinct introductions from Halmahera to Near and Remote Oceania, which would contradict the archeological and linguistic evidence linking Near and Remote Lapita communities. Alternatively, it may be the result of two introductions of R. exulans to Near Oceania— an early one to the main islands, followed by a Lapita introduction to small islets; that is, locations that have not been sampled to date in either morphological or genetic studies. If one lineage of exulans was introduced and established on the large islands earlier than 3,500 BP and Lapita peoples later introduced a new lineage to the previously uninhabited islands, then exulans from the smaller islets, like Mussau or other offshore islands in the Bismarcks, should have the Remote Oceanic lineages and be distinct from those found on the larger islands. If they are found to be the same as the mainland samples, then Remote Oceanic exulans must be getting from Halmahera to Remote Oceania by some other route and the settlement of the region is even more complex than we presently believe. We are currently embarking on collection trips for both modern and archeological R. exulans samples from the islands and islets off the north coast of New Guinea to address this issue.
Figure 10.2 An unrooted Neighbor Joining (NJ) tree and map of the Pacific showing location of samples and associated haplogroups. Bootstrap values for main branches are shown. Ellipses and symbols identify regions associated with particular haplogroups: filled circle, haplogroup I; asterisk, haplogroup II; open triangle, haplogroup IIIa; filled triangle, haplogroup IIIb. Haplotype R33 represents the three Thai samples. From Matisoo-Smith and Robins (2004).
163
regional studies and conclusion
In addition to studying mtDNA variation in R. exulans, we are also now examining the relationships between Pacific populations of ancient R. praetor (which are no longer extant in Remote Oceania), and the Asian rat, Rattus tanezumi, which was distributed throughout Micronesia prehistorically and generally pre-dates the appearance of R. exulans there. Complete mitochondrial DNA sequencing of the various species of rats found from Southeast Asia through Oceania may also allow for better identification of both extant and archeological murid species (Robins et al., n.d.).
Chicken Introductions Of all of the so-called Lapita-introduced “domesticated” animals, the one species that so far has not been claimed to be a possible pre-Lapita introduction is the red jungle fowl (Gallus gallus) or chicken. Unfortunately, the record of chicken remains in Lapita sites is no more informative than that for dogs or pigs. Chicken bones may be confused with other bird bones, and only small numbers of Gallus gallus bones have been identified in any Lapita site, with the exception of those in Tonga, where many are recorded (Steadman, 1993; Steadman et al., 2002; Storey, 2004). One Gallus gallus bone was reported from Watom Island, East New Britain, and a minimum number of 19 individuals is reported for Tikopia (Nagaoka, 1988). A PhD student, Alice Storey, in our laboratory at the University of Auckland, is currently undertaking analyses of aDNA from Pacific chicken bones. Preliminary results suggesting preferential preservation of aDNA in avian compared to mammal remains from open coastal sites are promising for analyses of Lapita associated faunal remains (Storey, 2004).
Linguistic Evidence of Animal Introductions As Pawley (chapter 3) has addressed more fully, linguistic evidence does provide some suggestions as to origins and timings of introductions of commensal animals as well as other components of the Lapita cultural complex. While the linguistic evidence suggests that indeed Lapita peoples carried the pig to the Pacific, the fact that it did not seem to be part of the Trans-New Guinea farming complex does not necessarily mean that it was not introduced as a wild animal at some point prior to Lapita. As Lynch (1991: 426) points out, there are two proto-Oceanic words for pig, *boRok and *bwo(e), which may suggest two different terms for wild vs domestic pigs. The fact that pigs were not brought by Lapita peoples to New Caledonia is also supported by linguistic data, as all of the New Caledonian and Loyalty Islands
164
terms for pigs are derived from the proto-Polynesian form, *puaka. The lack of a Proto-Oceanic reconstruction for dogs has been claimed by Hudson (1991), but Lynch (1991) states that while there are “no secure reflexes of the PAN *asu … reflexes of the Proto-Oceanic *guan are found in the Shortland Islands and in parts of Ysabel, Guadalcanal and Makira.” He goes on to point out that in the western Solomons, the reconstructions *siga and *pasi cannot be safely attributed to proto-Oceanic, so “(f)urther comparisons with forms in the New Guinea region is needed before any conclusive statements can be made” (1991: 427). Therefore we cannot at this point rule out the possibility of an earlier introduction of dogs to the highlands or elsewhere based on linguistic evidence. The use of the proto-Polynesian forms related to *kulii in most of the western Pacific is consistent with our findings that the Lapita–dog connection is questionable. Interestingly, while there is no archeological evidence for dogs in New Caledonia, the term for dog in the north of the Grande Terre is the POC reflex *guan (Lynch 1991). In New Guinea there are, of course, multiple genera and species of native rat but very little linguistic information on identification. In Kalam classification, R. exulans is specified as kopyak, which more generally refers to “dirty rats,” those found around the house or latrine, or kopyak walcogon, identifying them as the little rats or “squeakers.” Because they were classified as “dirty” they were generally not considered as food items except for old women and children, who would eat them if the rats were found far enough away from habitation sites (Bulmer and Menzies, 1972). According to Ross Clark (personal communication) there are two words reconstructed for protoAustronesian: *tikus which is mostly in western Indonesia, and *labaw which is much more widespread. The proto-Oceanic reconstruction is *kasupe or *kusupe, which occurs in North-Central Vanuatu, the Southeast Solomons, Western Fiji, various parts of New Guinea, and probably also in South Vanuatu and New Caledonia. Locally in Northern Vanuatu there is another innovative form *karivi. Polynesian languages don’t have any of the above forms for rat, but have three different forms of their own—*kumaa, *kimoa, and *kiole. Clark (1994, and personal communication) provides the linguistic evidence for chicken in the Pacific and suggests that there is no firm proto-Oceanic reconstruction for red jungle fowl or chicken, though a number of Oceanic languages use a generic term for bird, *manuk, when referring to fowl. A more widely used term, particularly referring to male fowl is the onomatopoeic *kokorako, a possible proto-Oceanic reconstruction. The proto-Polynesian reconstruction for chickens, *moa, which has no external cognates is found throughout Polynesia.
Animal Translocations, Genetic Variation, and the Human Settlement of the Pacific
Other Commensal Animal Research In addition to the analyses of the distribution and genetic variation in the four commensal animals associated with Lapita dispersal, researchers have also used this approach to track likely stowaway species. Austin (1999) analyzed genetic variation in a New Guinea native lizard, Lipinia noctua, to see if it could inform human colonization issues and the ultimate origins of Polynesians. His small sample set showed that L. noctua in Remote Oceania were monophyletic and the result of a rapid colonization from a single source—New Guinea—evidence he suggested was support for the “fast-train” hypothesis for Polynesian origins. However, since this lizard is a New Guinea native, and is not present in Island Southeast Asia, this result unfortunately could not address hypotheses on the issue of ultimate origins of the Lapita peoples, which was the proposed goal of the study (see Terrell, 2000, for even further issues with the Austin paper). If we learned more about the genetic distributions of a number of such unintended introductions in Oceania, these could provide a most instructive contrast to distributions of intentional introductions. As discussed in more detail in chapter 12, in recent years, Douglas Yen’s pioneering work on taro and sweet potato (1974, 1985, 1990) has been further extended by several young researchers studying genetic variation in Pacific food crops. Hinkle (2004) has investigated the distribution and genetic variation in Cordyline, or Hawaiian Ti plant, important throughout Polynesia for its leaves and as a food plant. It is believed that Pacific colonists brought the Cordyline with them as they moved eastwards across the Pacific. Clarke (2003, Clarke et al., 2006) has been addressing the issue of the introduction of plants into the Pacific from the west—specifically tracking the genetic origins of Pacific varieties of sweet potato (Ipomoea batatas) and bottle gourd (Legenaria siceraria) that it is believed have a South American origin. Genetic analyses of ancient plant remains, which are rare but occasionally preserved in archeological sites, is a most exciting development (Erickson et al., 2005).
Discussion and Conclusions So what does commensal animal evidence suggest about human interactions/origins? While the evidence for preLapita introductions of dog and pig is still debatable due to the limited nature of skeletal material and questions of context, the evidence for the Lapita introduction of dog is not much better. Table 10.1 provides a sample of key Lapita sites in both Near and western Remote Oceania, highlighting several points with regard to commensal animal remains.
Evidence from a range of fields now suggests that many of the elements formerly associated strictly with a Lapita intrusion are present in Near Oceania prior to the appearance of full-blown Lapita—including arboriculture, agriculture, pottery and even the mitochondrial DNA marker, the “Polynesian motif” (see chapter 9). This has led to the various explanations, like Green’s Triple I model, for Lapita origins, which allows for incorporation of indigenous (Near Oceanic) traits combined with introduced (i.e. Island Southeast Asian) characteristics and the development of new innovations. It is probably now safer not to assume any specific origin, including Southeast Asian, for any component of the Lapita cultural complex unless there is direct and convincing evidence for such an origin. The significance of early pig remains in Lapita sites should be put in a larger context. While pigs were apparently present from the earliest Lapita contexts, their economic significance or interpretation is not clear. Pigs are not present in large numbers until later in the archeological sequence. Is this evidence for full-scale domestication and breeding early on? The frequency of pig bone alone probably suggests not, but then we must develop scenarios for the cultural and economic conditions under which large populations of pigs develop. In addition, does it take a few hundred years for pig populations to expand so that remains are recoverable in the archeological record? It is also unclear as to whether or not the mode of “domestication” of pigs in the western Pacific would ever produce such a predicted pattern of archeological evidence. This review also underlines the apparent lack of archeological evidence for Lapita dogs. There is only a single dog bone or individual tooth in most cases where dog has been reported in Lapita sites, and the evidence is about the same for pre-Lapita dogs in Near Oceania. It is not until about 2,000 BP that we see indisputable evidence for dogs and, at that point, we see dog remains suddenly throughout the Pacific. As with the pig situation, is this simply the time when dog populations reach a point at which they become archeologically identifiable or is this an indicator of something else—perhaps a later introduction of dogs or a change in dog–human interactions? The apparent total lack of prehistoric dog (not just Lapita sites) in the Vanuatu and New Caledonian sites is curious. Given this lacuna, identifying the origins of Polynesian dogs becomes more complex, since there is clear archeological evidence for dog from early in the settlement of central east Polynesia. Was there a specific decision against the adoption of the dog by the peoples of Vanuatu and New Caledonia? Given the large sizes of the islands in those archipelagos, it would be difficult to see why this might be the case. Dogs would be valuable for hunting if not as a food source. The lack of dog remains
165
166
Table 10.1
Selected Early Lapita Sites and Presence of “Lapita-Associated” Animal Remains
Region
Site
Date BP
Pig
Dog
Rat
Chicken
Ref
Mussau
ECA, ECB, EKP
3,550–2,770
Yes, small #s of teeth and post cranial
(See note 1)? None identified—med mammal?
None reported
Yes, NISP 10 in ECA
Watom
SAC
2,900–2,000
Yes, abundant
None identified
Yes? NISP 1
Anir
ERA, ERC
Nissan
DFF
3,380–2,950, 2,950–2,360 3,300–
1 tooth from ERA TP14 layer 1 None
Buka
DAI
2,190–
Yes, some teeth and post cranial Yes, small amounts from 3,300–2,900 increasing later Yes, from bottom
Reef/Santa Cruz
RF-2
3,200–2,800
Yes
None
None recorded, but rat gnawing on bone in earliest layera Yes, exulans and possibly larger Murid Yes, exulans in layers 3,100–2,500 BP, praetor since 860 BP Yes, praetor from 1,860 BP no exulansa Yes, exulans, plus medium & large murid—prob praetor & mordax
Kirch 1988, 2001; Nagaoka 1988; Steadman and Kirch 1998 Smith 1998 (2000); Green 1998 (2000)
Tikopia
TK-4
2,900–
Yes, NISP 6
Yes? NISP 1
Yes, exulans (NISP 205) and praetor
Yes, MNI = 19
Vanuatu
numerous
From 3,000
Yes, from early Lapita
None identified
Yes, exulans and praetor
Yes
New Caledonia
numerous
From 3,000
None
None
Yes, exulans only
None
Fiji
Yanuka VL 16/81
2,980–
Yes? MNI 1
Yes? MNI 2
Rat bone identified, probably exulans
Yes? MNI = 2
Fiji
Lakeba 101/ 7/196, 197
2,935–
None before 1,000 BP
None before 1,000 BP
Yes in 197, from 2,700 BP, MNI = 20
Yes, early Lapita MNI = 19
Fiji
Naigani VL 21/5
2,850–
None identified
? possible dog fragments
Yes, exulans
None identified
Yes, from 1,860 BP
Note 1: Kirch (1988) lists 3 dog bones identified from ECA, but Kirch 2001 shows no record of dog bone from ECA. a May be an issue of screen size—only 5 mm sieve used. b Only mammal bone reported.
None identified None reportedb
Summerhayes 2001 and pers comm. 2005 Flannery et al., 1988
None reportedb
Flannery et al., 1988
Yes
Green 1991; R. Green pers comm.— Rattus material studied in our laboratories Nagaoka 1988; Kirch & Yen 1982; Flannery et al., 1988 Bedford 2000, pers comm. 2005, species ID of rats in our lab (mtDNA) Sand 1996, 2001; White et al., 2000, C. Sand pers comm. 2005 Nagaoka 1988; Clark and Anderson 2001 Best 2002; Clark and Anderson 2001; Nagaoka 1988 G. Irwin pers com. 2005; Clark and Anderson 2001; Kay 1984
Animal Translocations, Genetic Variation, and the Human Settlement of the Pacific
probably does not have a taphonomic explanation, or be the result of selective preservation or a recovery problem, as both Vanuatu and New Caledonia have numerous sites where faunal remains are present. The faunal situation in New Caledonia, however, is unique in a number of ways. Of all of the animals transported through the Pacific, the archeological evidence suggests that New Caledonia only received R. exulans. There is no evidence for dog, pig, chicken, or R. praetor. The faunal record in New Caledonia, however, is full of turtle, bird, fish, crocodilian, and other reptilian remains, so again it is unlikely that there is a taphonomic explanation for this situation. What does this suggest about the degree and timing of contact and interaction between New Caledonia and other archipelagos? Does this pattern simply reflect specific cultural choices made by the inhabitants of the region? One explanation for the unusual distribution of faunal remains in New Caledonia and elsewhere may be that our interpretations are based solely on the identification of the remains. Many sites record “unidentified medium mammal” in cases where the bones are devoid of any features that allow identification of the species. This is where molecular tools may be of particular value in future research. While aDNA studies are expensive and destructive and thus should not be used ad hoc, when applied to specific questions they can provide essential information, including species identification. Given that determining the presence or absence of particular animal remains has significant implications for questions of Pacific prehistory, this is an important area for future development. There is also the question of field collection and the reliability of the excavated remains themselves, as illustrated by the problems of screen size, sampling, and quantification (Nagaoka, 1988). Yet faunal remains provide key information about human interactions, subsistence strategies, and even population origins. Highlighting these issues and the potentials of molecular data in particular for addressing some key archeological questions should lead to more focused and in-depth collaborations between field archeologists, faunal analysts, and molecular archeologists. In doing so, we can further our understanding of prehistory and past connections in Island Melanesia and the Pacific in general.
Acknowledgments I thank Stuart Bedford, Sue Bulmer, Ross Clark, Moira Doherty, Roger Green, Judith Robins, Christophe Sand, Matthew Spriggs, Alice Storey, Glenn Summerhayes, and Peter White for providing valuable information and useful discussions on this topic and manuscript.
References Allen J. 2000. From beach to beach: The development of maritime economies in prehistoric Melanesia. Modern Quaternary Research Southeast Asia 16: 137–76. Austin CC. 1999. Lizards took the express train to Polynesia. Nature 397: 113–14. Bellwood P. 1997. Prehistory of the Indo-Malaysian Archipelago. Honolulu, Hawaii: University of Hawaii Press. Bellwood P, White JP. 2005. Domesticated pigs in eastern Indonesia. Science 309: 381. Bellwood P, Nitihaminoto G, Irwin G, Gunadi, Waluyo A, Tandirjo D. 1998. 35,000 years of prehistory in northern Moluccas. Modern Quaternary Research Southeast Asia 15: 233–73. Bulmer R. 1967. Why is the cassowary not a bird? A problem of zoological taxonomy among the Karem of the New Guinea highlands. Man 2(1): 5–25. Bulmer R. 1976. Selectivity in hunting and in disposal of animal bone by the Kalam of the New Guinea Highlands. In: Sieveking G, Longworth H, Wilson KE, editors. Problems in economic and social archaeology. London: Duckworth. pp. 169–86. Bulmer RNH, Menzies JI. 1972. The lexicon of proto-oceanic. Volume 3 (Flora & Fauna). Karam classification of marsupials and rodents. Journal of the Polynesian Society 81(4): 472–99. Bulmer S. 1975. Settlement and economy in prehistoric Papua New Guinea. Journal de la Société Océanistes 31: 8–75. Bulmer S. 1978. Prehistoric culture change in the Port Moresby region. Ph.D dissertation. Port Moresby, Papua New Guinea: University of Papua New Guinea. Bulmer S. 1982. Human ecology and cultural variation in prehistoric New Guinea. In: Gresswit J, editor. Biogeography and ecology of New Guinea. The Hague, Netherlands: DR Junk. pp 169–206. Bulmer S. 2001. Lapita dogs and singing dogs and the history of the dog in New Guinea. In: Clark GR, Anderson AJ, Vunidilo T, editors. The archaeology of Lapita dispersal in Oceania—Papers from the Fourth Lapita Conference, June 2000. Canberra, Australia: Pandanus Books. pp 183–201. Burley DV. 1998. Tongan archaeology and the Tongan past, 2850–150 BP. Journal of World Prehistory 12(3): 337–92. Clark G, Anderson A. 2001. The pattern of Lapita settlement in Fiji. Archaeology in Oceania 36: 77–88. Clark R. 1994. The word is the bird: Evolution, migration and extinction of Oceanic ornighonyms. In: Pawley A, Ross MD editors. Austronesian terminologies: Continuity and change. Pacific Linguistics Series C No. 127. Canberra, Australia: Department of Linguistics, Research School of Pacific and Asian Studies, The Australian National University. Clarke A. 2003. Using DNA sequence data to trace the origins and dispersal of sweet potato and bottle gourd in Polynesia: Implications for human mobility.
167
regional studies and conclusion
Paper presented, Australasian Society for Human Biology Annual Conference. Auckland, New Zealand. Clarke A, Burtenshaw MK, McLenachan PA, Penny D. 2006. Reconstructing the origins and dispersal of the Polynesian bottle gourd (Lagenaria siceraria). Molecular Biology and Evolution 23(5): 893–900. Corbett LK. 1985. Morphological comparisons of Australian and Thai dingoes: A reappraisal of dingo status, distribution and ancestry. Proceedings of the Ecological Society of Australia 13: 277–91. Erickson DL, Smith BD, Clarke AC, Sandweiss DH, Tuross N. 2005. An Asian origin for a 10,000-year-old domesticated plant in the Americas. Proceedings of the National Academy of Sciences USA 102(51): 18315–20. Flannery TF. 1995. Mammals of the Southwest Pacific and Moluccan Islands. Sydney, Australia: Reed Books. Flannery TF, White JP. 1991. Animal translocations: Zoogeography of New Ireland Mammals. National Geographic Research and Exploration 7: 96–113. Flannery TF, Kirch PV, Specht J, Spriggs M. 1988. Holocene animal fauna from archaeological sites in Island Melanesia. Archaeology in Oceania 23(3): 89–94. Gollan K. 1985. Prehistoric dogs in Australia: An Indian origin? In: Misra VN, Bellwood P, editors. Recent advances in Indo-Pacific prehistory. Oxford, UK: Oxford University Press. pp 439–43. Golson J, Hughes JP. 1980. The appearance of plant and animal domestication in New Guinea. Journal de la Société Océanistes 69: 294–303. Gorecki P, Mabin M, Campbell J.1991. Archaeology and geomorphology of the Vanimo Coast, Papua New Guinea: Preliminary results. Archaeology in Oceania 26: 119–22. Gosden C, Allen J, Ambrose W, Anson D, Golson J, Green R, Kirch P, Lilley I, Specht J, Spriggs M. 1989. Lapita sites in the Bismarck Archipelago. Antiquity 63: 561–86. Grayson DK. 2001. The archaeological record of human impacts on animal populations. J World Prehistory 15(1): 1–68. Green RC. 1991. A reappraisal of the dating for some Lapita sites in the Reef/Santa Cruz Group of the Southeast Solomons. Journal of the Polynesian Society 100: 197–207. Green RC. 2000. Lapita and the cultural model for intrusion, integration and innovation. In: Anderson A, Murray T, editors. Australian Archaeologist— Collected papers in honour of Jim Allen. Canberra, Australia: Coombs Academic Publishing. pp 372–92. Groves C. 1981. Ancestors for the pigs: Taxonomy and phylogeny of the genus Sus. Australian National University, Technical Bulletin No. 3, Canberra, Australia: Department of Prehistory, Research School of Pacific Studies. Hedges REM, Housley RA, Bronk Ramsey C, van Klinker GJ. 1995. Radiocarbon dates from the Oxford AMS system, Archaeometry data list 20. Archaeometry 37: 417–30. Heinsohn T. 1998. Captive ecology. Nature Australia 26(2): 36–43.
168
Heinsohn T. 2001. Human influences on vertebrate zoogeography: Animal translocation and biological invasions across and to the east of Wallace’s Line. In: Metcalfe JM, Smith B, Morewood M, Davidson I, editors. Faunal and floral migrations and evolution in Southeast Asia-Australasia. Lise, The Netherlands: AA Balkema. pp 154–70. Hide R. 2003. Pig husbandry in New Guinea: A literature review and bibliography. Canberra, Australia: Australian Centre for International Agricultural Research. ACIAR Monograph No. 108. Hinkle AE. 2004. Distribution of a male sterile form of “Ti” (Cordyline fruticosa) in Polynesia and its ethnobotanical significance. Journal of the Polynesian Society 113(3): 263–90. Hudson E. 1991. All ngggaun: Oceanic terms for ‘dog’. In: Harlow R, Hooper R, editors. VICAL I: Papers from the Fifth International Conference on Austronesian Linguistics, Part One. Auckland, New Zealand: Linguistic Society of New Zealand. pp 283–304. Irwin GJ. 1977. The emergence of Mailu as a central place in the prehistory of coastal Papua. Ph.D. dissertation. Canberra, Australia:.The Australian National University. Kirch PV. 1988. The Talepakemalai Lapita site and Oceanic prehistory. National Geographic Research and Exploration 4: 328–42. Kirch PV. 1997. The Lapita peoples: Ancestors of the Oceanic world. Cambridge, MA: Blackwell. Kirch PV. 2000. On the road of the winds: An archaeological history of the Pacific Islands before European contact. Berkeley, CA: University of California Press. Kirch PV. 2001. Lapita and its transformations in Near Oceania: Archaeological investigations in the Mussau Islands, Papua New Guinea, 1985–88. Volume 1 Introduction, excavations, chronology. No 59 Contributions of the Archaeological Research Facility. Berkeley, CA: University of California. Kirch PV, Yen DC. 1982. Tikopia: The prehistory and ecology of a Polynesian Outlier. B.P. Bishop Museum Bulletin 238. Larson G, Dobney K, Albarella U, Fang M, Matisoo-Smith E, Robins J, Lowden S, Finlayson H, Brand T, Willerslev E, Rowley-Conwy P, Andersson L, Cooper A. 2005. Worldwide phylogeography of wild boar reveals multiple centers of pig domestication. Science 307: 1618–21. Leavesley M. 2005. Prehistoric hunting strategies in New Ireland, Papua New Guinea: The evidence of the cuscus (Phalanger orientalis) remains from Buan Merabak Cave. Asian Perspectives 44(1): 207–18. Leonard JA, Wayne RK, Wheeler J, Valez R, Guillen S, Vila C. 2002. Ancient DNA evidence for Old World origin of New World dog. Science 298: 1613–16. Lilley I. 1986. Prehistoric exchange in the Vitiaz Strait, Papua New Guinea. Ph.D. dissertation. Canberra, Australia: The Australian National University. Lynch J. 1991. Pigs and dogs in Island Melanesia. In: Pawley A editor. Man and a half: Essays in Pacific anthropology and ethnobiology in honour of
Animal Translocations, Genetic Variation, and the Human Settlement of the Pacific
Ralph Bulmer. The Polynesian Society Memoir No. 48, Auckland, New Zealand: The Polynesian Society. Marshall B, Allen J. 1991. Excavations at Panakiwuk Cave, New Ireland. In: Allen J, Gosden C, editors. Report of the Lapita homeland project. Occasional papers in prehistory 20. Canberra, Australia: Department of Prehistory, Research School of Pacific Studies. pp 59–91. Matisoo-Smith E. 1994. The human colonisation of Polynesia. A novel approach: Genetic analyses of the Polynesian Rat (Rattus exulans). Journal of the Polynesian Society 103: 75–87. Matisoo-Smith E, Allen JS. 2001. Name that rat: Molecular and morphological markers in Pacific rodent remains. International Journal of Osteoarcheology 11: 34–42. Matisoo-Smith E, Robins J. 2004. Origins and dispersals of Pacific peoples: Evidence from mtDNA phylogenies of the Pacific rat. Proceedings of the National Academy of Sciences USA 101(24): 9167–72. Matisoo-Smith E, Roberts RM, Allen JS, Irwin GJ, Penny D, Lambert DM. 1998. Patterns of human colonisation in Polynesia revealed by mitochondrial DNA from the Polynesian Rat. Proceedings of the National Academy of Sciences USA 95(25): 15145–50. Matisoo-Smith E, Sutton DG, Ladefoged TN, Lambert DM, Allen JS. 1999. Prehistoric mobility in Polynesia: MtDNA variation in Rattus exulans from the Chatham and Kermadec Islands. Asian Perspectives 38(2): 186–99. Motokawa M, Lin L-K, Lu K-H. 2004. Skull geographic variation in the Polynesian Rat, Rattus exulans. Raffles Bulletin of Zoology 52(2): 653–63. Nagaoka L. 1988. Lapita subsistence: The evidence of non-fish archaeofaunal remains. In: PV Kirch, TL Hunt editors. Archaeology of the Lapita cultural complex: A critical review. Seattle, WA: Thomas Burke Memorial Washington State Museum Research Report No. 5. pp 117–33. Okumura N, Ishiguro N, Nakano M, Hirai K, Matsui A, Sahara M. 1996. Geographic population structure and sequence divergence in the mitochondrial DNA control region of the Japanese wild boar (Sus scrofa leucomystax), with reference to those of domestic pigs. Biochemical Genetics 34: 179–89. Roberts M. 1991. Origin, dispersal routes and geographic distribution of Rattus exulans (Peale) with special reference to New Zealand. Pacific Science 45: 123–30. Robins J, Matisoo-Smith E, Ross H. 2003. The origins of the feral pigs on the Auckland Islands. Journal of the Royal Society of New Zealand 33(2): 561–9. Robins J, Matisoo-Smith E, Ross H (n.d.) Identifying Rattus species using small fragments of mitochondrial DNA. Molecular Ecology (submitted). Sand C. 1996. Recent developments in the study of New Caledonia’s prehistory. Archaeology in Oceania 31: 45–71. Savolainen P, Zhang Y-P, Luo J, Lundeberg J, Leitner T. 2002. Genetic evidence for an East Asian origin of domestic dogs. Science 298: 1610–13. Savolainen P, Leitner T, Wilton AN, Matisoo-Smith E, Lundeberg J. 2004. A detailed picture of the origin of
the Australian dingo obtained from the study of mitochondrial DNA. Proceedings of the National Academy of Sciences USA 101(33): 12387–90. Spriggs MJT. 1997. The Island Melanesians. Oxford, UK: Blackwell. Steadman DW. 1993. Biogeography of Tongan birds before and after human impact. Proceedings of the National Academy of Sciences USA 90: 818–22. Steadman DW, Kirch PV. 1998. Biogeography and prehistoric exploitation of birds in the Mussau Islands, Bismarck Archipelago, Papua New Guinea. Emu, 98: 13–22. Steadman DW, Plourde A, Burley DV. 2002. Prehistoric butchery and consumption of birds in the Kingdom of Tonga, South Pacific. Journal of Archaeological Science 29: 571–8. Storey AA. 2004. Save me a drumstick: Molecular taphonomy, differential preservation and ancient DNA from the Kingdom of Tonga. Unpublished MA dissertation, Simon Fraser University, Burnaby, Canada. Summerhayes G. 2003. The rocky road: The selection and transport of Admiralites obsidian to Lapita communities. Australian Archaeology 57: 135–42. Swadling P. 1997. Changing shorelines and cultural orientations in the Sepik-Ramu, Papua New Guinea: Implications for Pacific prehistory. World Archaeology 29: 1–14. Tate GHH. 1935. Rodents of the genera Rattus and Mus from the Pacific Islands. Bulletin of the American Museum of Natural History 68: 145–78. Taylor JM, Calaby JH, Van Deusen HM. 1982. A revision of the genus Rattus (Rodentia, Muridae) in the New Guinean region. Bulletin of the American Museum of Natural History 173: 177–336. Terrell JE. 2000. Anthropological knowledge and scientific fact. American Anthropologist 102(4): 808–17. Vilá C, Savolainen P, Maldonado JE, Amorim IR, Rice JE, Honeycutt RL, Crandall KA, Lundeberg J, Wayne RK. 1997. Multiple and ancient origins of the domestic dog. Science 276: 1687–9. Watanobe T, Ishiguro N, Okamura N, Nakano M, Matsui A, Hongo H. 2001. Prehistoric introduction of domestic pigs onto the Okinawa islands: Ancient mitochondrial DNA evidence. Journal of Molecular Evolution 55: 222–31. White JP. 2004. Where the wild things are: Prehistoric animal translocation in the circum New Guinea archipelago. In: Fitzpatrick SM editor. Voyages of discovery: The archaeology of Islands. Westport: Praeger. White JP, Flannery TF, O’Brien R, Hancock RV, Pavlish L. 1991. The Balof shelters in New Ireland. In: Allen J, Gosden C editors. Report of the Lapita homeland project. Occasional Papers in Prehistory 20. Canberra, Australia: Department of Prehistory, Research School of Pacific Studies. pp 46-58. White JP, Clark J, Bedford S. 2000. Distribution, present and past, of Rattus praetor in the Pacific and its implications. Pacific Science 54(2): 105–17.
169
regional studies and conclusion
Yen DE. 1974. The sweet potato and Oceania: An essay in ethnobotany. Honolulu, Hawaii: Bishop Museum Press. Yen DE. 1985. Wild plants and domestication in Pacific islands. In: Misra VN, Bellwood P, editors. Recent advances in Indo-Pacific prehistory. New Delhi, India: Oxford/IBH Publishing. pp 315–26.
170
Yen DE. 1990. Environment, agriculture and the colonization of the Pacific. In: Yen DE, Mummery JMJ, editors. Pacific production systems: Approaches to economic prehistory. Canberra, Australia: Department of Prehistory, Australian National University. pp 258–77.
11 Viral Phylogeny and Human Migration in the Southwest Pacific Jill Czarnecki, Jonathan S. Friedlaender, and Gerald Stoner
Introduction While the earlier chapters of this book show how powerfully archeological, historical linguistic, and human genetic data are being used to reconstruct human prehistoric migrations in the Southwest Pacific, other, less traditional sources of information can also contribute. In particular, other organisms inextricably involved with humans may offer insight into human population dynamics. Besides dogs, pigs, and Pacific rats, all of which were intentionally carried along on ancient ocean voyages (Matisoo-Smith et al., 1997; Matisoo-Smith, 2002; Matisoo-Smith and Robins, 2004; see also chapter 10), certain human viruses, and in particular the JC virus, are now being used as markers of ancient population migrations. Agostini et al. (1997) outlined four essential requirements for using a viral marker to trace human population migrations. First, it must have evolved early enough in human history so that as it has accumulated mutations and differentiated over time, distinct genotypes have come to be associated with particular populations or regions. Second, in the ideal circumstance, the virus should be endemic worldwide, providing the maximum amount of information on a variety of human migrations. Third, the virus should be genetically relatively stable so that rapid change will not distort more ancient associations. However, it must not be so uniform that only a limited number of strains have developed. Finally, and perhaps most critically, viral transmission must be essentially limited to immediate family or community members, in order to preserve population associations from one generation to
the next. The more viral infection mimics human genetic transmission from parent to offspring, the better. Five viruses have been used in the study of human population migrations with varying degrees of success, and they each illustrate the strengths and weaknesses of this approach. Human T-lymphotropic virus type 1 (HTLV-1) is an oncoretrovirus that causes adult T-cell leukemia, neuromyelopathy, and a variety of other illnesses (Mahieux et al., 1997). Several different subtypes have been found that correlate to geographic regions, including Africa, Melanesia, Australia, and East Asia, making it possible to both construct a phylogeny of the virus and to estimate a molecular clock from the viral mutation rates although a date of origin for the virus has not yet been determined (Black, 1997; Mahieux et al., 1997; Miura et al., 1994; Yanagihara, 1994; Yanagihara et al., 1995). It has a stable genome with little variation between strains (Mahieux et al., 1997), and it is transmitted via breast milk, thus creating a maternal transmission pattern similar to that of mitochondrial DNA (Black, 1997). Its major drawback is that it is not broadly endemic and therefore is informative in only a few geographical areas (Sugimoto et al., 1997). In the Southwest Pacific, the incidence of HTLV-1 infection is 0% in Irian Jaya and ranges from 0 to 14.6% in Papua New Guinea (Takao et al., 2000; Tanggo et al., 2000), so it is not highly informative for variation within the region. Human papillomavirus type 16 (HPV-16) is the predominant type of genital papillomavirus found in human populations worldwide (Chan et al., 1992). It causes various
171
regional studies and conclusion
types of epithelial neoplasia, with genital cancer being the best known. HPV-16 variants have been studied from at least 25 different regions (Ho et al., 1993), and these studies suggest the coevolution of different strains with three major continental groups: Africans, Europeans, and East Asians, with a possible African origin (Ho et al., 1993). The virus is thought to have arisen more than 200,000 years ago, it has been identified in patients worldwide, it is genetically stable, and it is transmitted via intimate (genital) contact between individuals (Ho et al., 1993). The major problem in using HPV-16 to reconstruct ancient migrations is that the variants that apparently correlate best to continental distinctions seem to have been obscured by more recent migrations occurring since the sixteenth century (Sugimoto et al., 1997). Although very little HPV16 research has been carried out in the Southwest Pacific, the initial results there illustrate this problem. One study of 8 affected women in New Caledonia showed that all carried HPV-16 isolates of European origin (Watts et al., 2002). This suggests a very recent introduction there of HPV-16. Herpes simplex virus type 1 (HSV-1) infection can manifest itself as oral or genital herpes, the latter being a sexually transmitted disease (Umene and Sakoaka, 1997). HSV-1 strains have been shown to have distinct patterns in geographically separate countries and ethnic groups (Umene and Sakoaka, 1997). For example, areas of Japan and Korea show distinct differences that correlate to the known population history of the region (Umene and Sakoaka, 1997). The virus has apparently existed in the human population for approximately 100,000 years (Major et al., 1992), it is widespread worldwide, it contains stable RFLPs that can be used as diversity markers (Umene and Sakoaka, 1997), and it is spread via close sexual contact. As with HPV-16, because sexual contact is not necessarily limited to particular communities, HSV-1 is problematic in tracing human population migrations. Since it is a sexually transmitted disease, many people will not harbor the virus, making positively testing sample sizes small (that is, prevalence rates will be low). In the Southwest Pacific, studies have characterized a different HSV viral type, HSV-2, infection in Papua New Guinea (Rezza et al., 2001), but little work has been done on the prevalence of HSV-1 infection. Further study of this virus will elucidate its utility in tracing population migration within this region. Varicella zoster virus (VZV), another herpesvirus, appears to have been present in the primate lineage for 60–70 million years (Wagenaar et al., 2003). VZV was shown to have distinct Asian and European phylogenetic clades (Wagenaar et al., 2003) and may prove to be another informative virus for tracing human population history in the future. In the Southwest Pacific, VZV has been identified in certain populations in Papua New Guinea
172
(Cook, 1985) and likely has appreciable prevalence rates elsewhere. While no work on the strain characterization or phylogeny of this virus in the Southwest Pacific has appeared, again it may prove to be informative with further testing.
JC Virus (JCV) JC virus (carrying the initials of the first positively identified subject), the last of the five, is proving to be far more useful in population history studies and consequently is the focus of the rest of this chapter. As part of the Polyomavirus subfamily, it is a DNA virus with a doublestranded covalently closed circular genome of approximately 5,000 base pairs and an icosahedral capsid made up of three viral proteins (Eckhardt, 1991). There are two human polyomaviruses, BKV and JCV, which are highly homologous to Simian virus 40 (SV40), all of which persist in the kidneys of the host species. JCV was first isolated in 1971 (Padgett et al., 1971) and was found to be the causative agent of a rare demyelinating disease, progressive multifocal leukoencephalopathy (PML), which affects immunocompromised individuals, predominantly patients with acquired immune deficiency syndrome (AIDS). PML is a disease of the central nervous system, causing brain lesions and typically leading to death within 6 months from time of onset (Frisque and White, 1992). As a result of this disease association, JCV has been studied extensively. JCV exists in two forms, called archetype and PML type, the latter of which causes the disease. The PML type results from a rearrangement of the archetype viral genome in response to a deficient immune system. In contrast to the PML type, JCV archetype establishes itself in the kidney and urinary tract of healthy individuals for the long term with no negative effect on the host (Boldorini et al., 2005). JCV archetype is truly endemic in humans. A screen of healthy individuals in Wisconsin showed that 65% of individuals had developed anti-JCV antibodies by age 14, with the number increasing to 75% in the 50–59-year cohort (Frisque and White, 1992). However, the mode of JCV transmission still remains unknown. It appears that the virus is transmitted fairly strictly within families, usually from parent to offspring, but not exclusively from the maternal side. An Okinawan study showed that American soldiers who were long-term residents there had not passed their distinctive strains of JCV on to the native population (Kato et al., 1997). Zheng et al. (2004a) also found that in four out of five families studied, identical JCV DNA sequences were found in both the parent and the child. The fifth family did not have identical JCV sequences, but the sequences found in the parent and child were closely related phylogenetically, suggesting
Viral Phylogeny and Human Migration in the Southwest Pacific
that the initial strain may have disappeared from the parent due to frequent genome changes. In spite of remaining questions on the nature of its transmission, JCV fulfills the four requirements to make it useful as a marker of population migration. Based upon the relationship between Asian and Native American JCV strains and dating from archeological evidence, JCV is believed to have been coevolving along with the major lineages of early humans possibly since their migration out of Africa some 100,000 years ago (Agostini et al., 1997). It is ubiquitous, existing in approximately 80% or more of the human population (Walker and Frisque 1986; Ault and Stoner 1992; Sugimoto et al., 1997; Agostini et al., 2001a). The archetype JCV is relatively stable. Although no direct ascertainment of the JCV mutation rate has been made, it is clearly fast enough for isolated geographic areas to develop distinguishing mutations, but is not so fast that these relationships are obscured. And whatever the exact mechanism, JCV is transmitted within families or close communities, preserving geographic/ethnic associations.
Table 11.1 Worldwide Distribution of JCV Genotypesa Shown with Major Geographic Domains and a Detailed Geographic Breakdown Major Geographic Domain Europe
Africa
Asia
Oceania
JCV Methodology JCV phylogenetic analysis is described in detail by Agostini et al. (2001b) and is similar to the methods used in the study of mitochondrial DNA. Briefly, the virus is extracted from urine samples and then the viral DNA is extracted from the virus. The viral DNA is then subjected to polymerase chain reaction (PCR) to amplify the DNA in sufficient quantity for DNA sequencing. The JCV genome contains many informative polymorphic sites that correlate to specific locations (Cui et al., 2004). These informative sites are used to determine a viral genotype. In the study of JCV, the term genotype refers to the evolutionary events, or mutations, that have taken place in a geographically defined population (Agostini et al., 2001b) and therefore allow correlation of the virus variants to a particular geographic region (table 11.1). JCV genotype distribution studies can be used to identify migration patterns of humans carrying these viruses. In addition to gathering information from genotype distribution, sequence information gained from JCV can be used to generate phylogenetic trees that give insight into the relationships between JCV strains worldwide, and also within smaller geographic areas. The phylogenetic branching order, branch length, and clustering of particular viral strains all are particularly informative aspects. The genotypes used in the classification of JCV are based on variations in the DNA sequence, rather than on any antigenic difference between strains, since sequence variation has not been shown to cause any antigenic epitopes that are serologically distinguishable (Agostini et al., 2001b). In order to divide JCV strains into genotypes
JCV Genotype
Specific Geographic Domain
1A 1B 4 3A 3B 6 2A1 2A2 2A3 2B 2D1 2D2 7A 7B1 7B2 7C1 7C2 2E 8A 8B
E. & N. Europe W. & SW Europe Europe Africa W. Africa W. & Central Asia Japan, S. Korea Americas S. Korea Eurasia, USA China/Mongolia S. India SE Asia & S. China China, Korea China China S. India Oceania PNG PNG & Melanesia
a
Nomenclature from Stoner and colleagues at the National Institutes of Health. See Czarnecki (2003) for corresponding nomenclature from the University of Tokyo.
and subtypes, the percentage of sequence variation between strains is used. Thus far, different JCV strains have been shown to differ by up to 2.6% of their genome (the highly variable regulatory region is excluded) (Agostini et al., 2001b). In order to be considered a major genotype, the strains must differ by 1% or more (Agostini et al., 2001b). To be recognized as subtypes, strains within a genotype must still differ by more than 0.3% of their sequence, be closely related phylogenetically, and correlate to a geographical region. Several major genotypes are limited to major geographical regions such as Asia and Africa, as shown in table 11.1. Early JCV phylogenetic trees were based on sequencing of a 610 bp fragment from the intergenic region (Sugimoto et al., 1997; Guo et al., 1998; Chang et al., 1999; Saruwatari et al., 2002; Sugimoto et al., 2002; Miranda et al., 2003), while more recent ones use whole genome sequences (Jobes et al., 1998, 2001; Agostini et al., 2001a; Fernandez Cobo et al., 2001, 2002; Yanagihara et al., 2002; Zheng et al., 2004b, Cui et al., 2004). Interpretations of both sets of phylogenetic trees have been controversial because it appeared at first that a European strain might be ancestral to the rest, contrary to the generally accepted “Out of Africa” model for modern human origins. JCV phylogenies derived by Sugimoto et al. (1997) showed African
173
regional studies and conclusion
and Asian JCV types being more closely related to each other than either is to the European type, as did a subsequent phylogenetic analysis by Hatwell and Sharp (2000). As a result of these analyses, the usefulness of JCV as a marker of human migration was questioned (Wooding, 2001). However, Jobes et al. (1998) disagreed with this interpretation as put forth by Sugimoto et al. (1997). While they agreed that the European Type 1 does appear to have diverged early on, that is not necessarily evidence for it being ancestral to all JCV types. Since there may be no clearly identifiable root, the long European branch lengths might be the result of faster mutation rates. They also argued that European types have a wide geographic distribution and therefore may not have been under the same normalizing selective constraints as other smaller, isolated, or genetically distinct populations (Jobes et al., 1998). The authors also mention another possible explanation that European Type 1 could be evolving at a rate different from other types. More recently, Pavesi (2003) conducted neighbor-joining analysis on the slow-evolving sites of JCV, thought to be the most accurate reflection of early viral evolution, and concluded that an African origin of JCV was most likely. By contrast, his phylogenetic analysis on the fast-evolving sites of JCV generated a tree much like that described by Hatwell and Sharp (2000). His results support an origin of European Types 1 and 4 from the ancestral type of African Type 6, with a separate diversion of the Asian types from this same African Type. In more recent publications, Pavesi (2004, 2005) has further supported the African origin of all JCV genotypes. Taken together, these studies show that JCV is, in fact, a viable marker for ancient population relationships. The fact remains that all JCV phylogenies constructed to date, whether based on whole genome sequences or intergenic region sequences, clearly differentiate African, European, and Asian clades. This is clearly shown in the maximum parsimony tree shown in figure 11.1.
Worldwide JCV Strain Distributions As shown in figure 11.1 and table 11.1, JCV strains are characterized by sharply constrained regional distributions, with Types 3A, 3B, and 6 in Africa, Types 1A, 1B, and 4 in Europe, Types 8A, 8B, and 2E in the Southwest Pacific, and the remaining types in Asia. Based on the fact that all of these JCV genotypes only differ by 1–2.6% of their genome, it has been difficult to construct a reliable molecular clock for this virus (Agostini et al., 2001b). Despite this fact, the general worldwide JCV genotype distribution and the phylogenetic trees constructed thus far fit well with current knowledge on the dispersal of modern humans out of Africa some 100,000 years ago (Agostini et al., 2001b). Additionally, the close relationship between Asian and North American JCV genotypes points to major JCV
174
genotypes being established well before the migration of humans into the Americas up to 30,000 years ago (Agostini et al., 1997). Taken together, this evidence makes it possible that JCV genotypes dispersed from Africa with early human migrations some 100,000 years ago (Agostini et al., 2001b).
Southeast Asia and the Southwest Pacific As shown in figure 11.1, a good deal is now known concerning the distribution of JCV strains in the Southwest Pacific (Agostini et al., 1997; Jobes et al., 1999, 2001; Ryschkewitsch et al., 2000; Yanagihara et al., 2002; Czarnecki, 2003; and Czarnecki et al., manuscript). JCV Type 3, which is distributed across Africa and as far east as northern India, is ancestral to all the Asiatic and Southwest Pacific variants. The first two distinctive clades from this African root are types 8A and 8B, and all the other sub-branches appear to be subdivisions of 8B (i.e., the major Asian genotypes 2 and 7). Types 8A and 8B stand out because they are each limited in their distributions to the Southwest Pacific and have a great deal of internal diversity. As their long branches and high bootstrap values indicate, they form very distinctive JCV clades, and it is misleading that their nomenclature suggests a subtype-like relationship between them, when they are clearly distinct genotypes within this region. Each of them appears to be old because of their high bootstrap values and their internal branching. The phylogenetic relationship among the 8A strains from several studies indicates its greatest diversity is within the Eastern Highlands of New Guinea, with a less diverse set distributed along the New Guinea coast. The frequency distribution of 8A also suggests a highlands center, if not origin, for this strain. A total of 51% of highlands strains detected tested positive for 8A, vs. only 12% on the coast (figure 11.2) (Czarnecki, 2003). Significantly, 8A has not been found outside New Guinea to date. An intriguing feature of Type 8A genotypes is the presence of a 21 bp deletion in the agnoprotein gene (Jobes et al., 1999). This deletion, found in all 8A genotypes sequenced and restricted to this genotype, is the only known agno-gene deletion observed in any JCV genotype, and the first stable deletion in the coding region of a healthy individual. It might possibly confer some selective advantage to the virus. Prediction of the mRNA secondary structure of the deleted agnoprotein shows the elimination of a large stem and loop structure, which could possibly “alter (upregulate) late gene expression by relieving secondary structural constraints” (Jobes et al., 2001). This deletion could also change codon usage, increasing accuracy and efficiency of late gene translation (Jobes et al., 2001). In any case, this deletion sets 8A apart distinctively and makes it difficult to envision 8A, with the
Figure 11.1 Maximum parsimony tree constructed using worldwide JCV whole genomes. Basic groupings of African and European clades are shown. The Asian clade is further divided into subclades to show relationships among Asian genotypes. Hatching patterns show JCV isolates corresponding to specific regions in the Southwest Pacific. JCV genotypes are taken from the published literature (Agostini et al., 1997, 1998a, 1998b, 1998c, 2001; Cui et al., 2004; Czarnecki, 2003; Fernandez-Cobo et al., 2001, 2002; Frisque et al., 1984; Jobes et al., 2001; Kato et al., 2000; Loeber and Dories, 1998; Stoner et al., 2000; Yanagihara et al., 2002). 175
regional studies and conclusion
Figure 11.2 JCV genotype distribution pattern in the Southwest Pacific. The frequency of JCV genotypes throughout the Southwest Pacific is shown. Hatching patterns represent individual JCV genotypes. Frequency data are based on published literature (Agostini et al., 1997; Czarnecki, 2003; Jobes et al., 1999, 2001; Ryschkewitsch et al., 2000; Yanagihara et al., 2002).
deletion, as close to the ancestor of 8B or the other strains in that large phylogenetic branch. Because it is fixed in this strain, the deletion must have arisen early in the independent evolution of 8A. The phylogenetic relationship between the Type 8B samples does not show a particular branching order or major differences in genetic diversity from region to region within its distribution (figure 11.1). What is apparent is a stable grouping of all New Guinea 8B strains together, separate from those distributed across the western Pacific as far as western Polynesia and central Micronesia. There is also a close clustering of the Vanuatu and Solomon Islands strains, and also between strains from Tonga and Wallis & Futuna. A number of hypothetical scenarios for the origins of 8A and 8B have been proposed, either seeing them as diverging from each other in New Guinea subsequent to their introduction, or as representing separate migrations, with 8A preceding 8B (Jobes et al., 1999; Yanagihara et al., 2002). While both derive from Afro-Asian Type 3, they are clearly very separate and comparatively old strains, and their regional restriction is very reminiscent of certain mitochondrial DNA and Y chromosome
176
haplogroup distributions (see chapters 4 and 5). A compatible scenario would have both developing outside Near Oceania in some south or southeast Asian locale, and being independently introduced to Sahul—8A preceding 8B by a significant time gap, with 8B not being introduced until after the New Guinea Highlands were effectively settled, beginning approximately 10,000 years ago (chapter 2). Subsequently, 8B, distributed along the Near Oceanic coasts, could have ultimately spread to Remote Oceania, reminiscent of the mitochondrial DNA scenarios suggesting Late Pleistocene or Holocene population movements (Lapita or otherwise). This is roughly in agreement with the scenario of Yanagihara et al. (2002). The distributions of the other JCV strains in the Southwest Pacific as well as their phylogenetic positions indicate they are more recent in date and in their introduction to the region. One or two are possible candidates for being direct introductions from Southeast Asia along with the introduction of agriculture and groups that were involved in the development of the Lapita phenomenon. Type 7A is the predominant JCV genotype found in Southeast Asia, along with 7B and 7C. They are found
Viral Phylogeny and Human Migration in the Southwest Pacific
in South China, Taiwan, Malaysia, the Philippines, and Indonesia (Sugimoto et al., 1997; Chang et al., 1999; Miranda et al., 2003; Cui et al., 2004). In Taiwan, 7A predominates in urban Taipei, with 7B1 and 7C1 in low frequencies, while in the Bunun (Taiwan Aboriginal Austronesian speakers) 7C1 and 7B are more common (Sugimoto et al., 1997; Chang et al., 1999), suggesting 7C1 and 7B are older, aboriginal forms. It had been suggested that type 7A is associated with Austronesian-speaking populations and migrations into the Southwest Pacific, possibly including the Austronesian expansion (Cui et al., 2004). The problem with this interpretation is that Type 7A in the Pacific has been found infrequently in the Southwest Pacific, only in Guam and in one Hawaiian individual (Ryschkewitsch et al., 2000; Yanagihara et al., 2002). Of course, sampling from Near and Remote Oceania has been limited, so Type 7A may actually be present at a higher frequency in these regions. Another issue that this volume addresses more generally is that there was very likely more than one distinct movement of people from Southeast Asia into Near Oceania in the last 10,000 years, and that finding a strain distributed across Southeast Asia and Oceania cannot simply be equated with a single population movement (Austronesian-speakers or otherwise). There were clearly a number of impulses over the region over a long time span. Type 2E appears to be restricted to the Southwest Pacific, where it has been found in low frequency, as well as Guam in somewhat higher frequencies (figure 11.2) (Rychkewitsch et al., 2000; Jobes et al., 2001; Czarnecki, 2003). It is most closely related to the widespread Asian Type 2A, as evidenced by the sister status of these two genotypes in the phylogenetic tree. Initial interpretation of Type 2E in the Pacific suggested that it was carried by the settlers of the Austronesian expansion (Yanagihara et al., 2002). More recent interpretations suggest that Type 2E may actually reflect an earlier migration into the region before the Austronesian expansion (Cui et al., 2004). The Type 2E strains found thus far in PNG do not appear to be present at a higher frequency in either the coastal or highland regions, nor do they appear to correlate with linguistic differences between groups there (Czarnecki, 2003). Whether Type 2E or 7A reflect the Austronesian expansion is unclear. Since no Type 2E strains have been found outside the Pacific, but are closely related to other Asian strains, it is possible that it may reflect a more recent migration from Asia, as compared to the older dispersals in the Pacific of Types 8A and 8B. The low frequency of Type 2E in the Southwest Pacific adds doubt to the interpretation that it is the sole representation of the Austronesian expansion. But, as discussed with Type 7A, it is possible that the Austronesian expansion included
multiple migrations over a period of time, and may have included multiple strains of JCV such as Types 2E and 7A. Further sampling would need to be carried out to clarify this issue.
Summary and Conclusions Although not a traditional genetic marker, JC virus has proven useful in understanding the movements of humans into and within the Southwest Pacific. Based on these similarities, JCV has added strength to the hypotheses derived from the human genetic data. In terms of the oldest migrations into the Southwest Pacific, JCV distributions and trees are very reminiscent of mtDNA and Y chromosomal DNA distributions, both having very old haplogroups restricted to New Guinea and Island Melanesia but with definite distinctions between them within the region. JCV genotypes 8A and 8B both appear to be very old, much like the mtDNA haplogroups P, Q, M27, M28, and M29 (Friedlaender et al., 2005; Merriwether et al., 2005). In terms of the more recent migrations within the Southwest Pacific, there appears to be an overlay of migrations from Southeast Asia, represented by JCV genotypes 7A and 2E, much like mtDNA haplogroup B4A (the “Polynesian motif”). It appears that more sampling is needed to ascertain whether one or both of these genotypes represent any portion of the Austronesian expansion. Like the mtDNA evidence, the JCV evidence does not clarify the exact nature of the Austronesian expansion into this region. The lack of a suitable method for applying a molecular clock to the evolution of JCV makes it difficult to make any clear-cut associations between the Austronesian expansion, for example, and particular JCV genotypes. Despite this, the fact that JCV genotype distributions and trees fit so well with the current human genetic data highlight the usefulness of this virus as a marker of human migration in the Southwest Pacific.
Acknowledgments This work would not have been possible without the invaluable assistance of the following people: members of the late Dr. Gerald Stoner’s laboratory group at the National Institutes of Health, specifically Caroline Ryschkewitsch, Dr. Christopher Cubitt, Dr. Xiaohong Cui, and Dr. Gennene Mengistu; at Temple University, Dr. L. Christie Rockwell, Dr. Joseph Lorenz, and Dr. Francoise Friedlaender; and Dr. Charles Mgone, formerly of the Papua New Guinea Institute for Medical Research. Financial assistance for this project was provided by the Wenner-Gren Foundation for Anthropological Research, Inc. and the National Institutes of Health.
177
regional studies and conclusion
References Agostini HT, Yanagihara R, Davis V, Ryschkewitsch CF, Stoner GL. 1997. Asian genotypes of JC virus in Native Americans and in a Pacific Island population: Markers of viral evolution and human migration. Proceedings of the National Academy of Sciences USA 94: 14542–6. Agostini HT, Ryschkewitsch CF, Stoner GL. 1998a. JC virus Type 1 has multiple subtypes: Three new complete genomes. Journal of General Virology 79(4): 801–5. Agostini HT, Shishido-Hara Y, Baumhefner RW, Singer EJ, Ryschkewitsch CF, Stoner GL. 1998b. JC virus Type 2: definition of subtypes based on DNA sequence analysis of ten complete genomes. Journal of General Virology 79(5): 1143–51. Agostini HT, Ryschkewitsch CF, Stoner GL. 1998c. Complete genome of a JC virus genotype Type 6 from the brain of an African American with progressive multifocal leukoencephalopathy. Journal of Human Virology 1: 267–72. Agostini HT, Deckhut AM, Jobes DV, Girones R, Schlunck G, Prost MG, Frias C, Perez-Trallero E, Ryschkewitsch CF, Stoner GL. 2001a. Genotypes of JC virus in East, Central and Southwest Europe. Journal of General Virology 82: 1221–31. Agostini HT, Jobes DV, Stoner GL. 2001b. Molecular evolution and epidemiology of JC virus. In: Khalili K, Stoner GL, editors. Human polyomaviruses: Molecular and clinical perspectives. New York: Wiley-Liss. pp 491–526. Ault HT, Stoner GL. 1992. Two major genotypes of JC virus defined in progressive multifocal leukoencephalopathy brain by early and late coding region DNA sequences. Journal of General Virology 73: 2669–78. Black FL. 1997. Tracing prehistoric migrations by the viruses they carry: Human T-cell lymphotropic viruses as markers of ethnic relationships. Human Biology 69: 467–82. Boldorini R, Veggiani C, Barco D, Monga G. 2005. Kidney and urinary tract polyomavirus infection and distribution. Archives of Pathology and Laboratory Medicine 129: 69–73. Chan S, Ho L, Ong C, Chow V, Drescher B, Durst M, Meulen J, Villa L, Luande J, Mgaya HN, Bernard H. 1992. Molecular variants of human papillomavirus type 16 from four continents suggest ancient pandemic spread of the virus and its coevolution with humankind. Journal of Virology 66(4): 2057–66. Chang D, Sugimoto C, Wang M, Tsai RT, Yogo Y. 1999. JC virus genotypes in a Taiwan aboriginal tribe (Bunun): Implications for its population history. Archives of Virology 144: 1081–90. Cook IF. 1985. Herpes zoster in children following malaria. Journal of Tropical Medicine and Hygeine 88(4): 261–4. Cui X, Wang JC, Deckhut A, Joseph BC, Eberwein P, Cubitt CL, Ryschkewitsch CF, Agostini HT, Stoner GL. 2004. Chinese strains (type 7) of JC virus are AfroAsiatic in origin but are phylogenetically distinct from the Mongolian and Indian strains (type 2D) and the Korean and Japanese strains (type 2A). Journal of Molecular Evolution 58: 568–83. 178
Czarnecki J. 2003. JC virus genotypes in Papua New Guinea. Ph.D. dissertation. Philadelphia: Temple University. Czarnecki J, Friedlaender JS, Mgone CS, Stoner GL. 2006. JC virus genotypes in Papua New Guinea: Implications for the peopling of the Pacific. Manuscript. Eckhardt W. 1991. Polyomavirinae and their replication. In: Fields BN, Knipe DM, Chanock RM, editors. Fundamental virology, 2nd Edition. New York: Raven Press. pp 727–41. Fernandez-Cobo M, Jobes DV, Yanagihara R, Nerurkar VR, Yamamura Y, Ryschkewitsch CF, Stoner GL. 2001. Reconstructing population history using JC virus: Amerinds, Spanish, and Africans in the ancestry of modern Puerto Ricans. Human Biology 73(3): 385–402. Fernandez-Cobo M, Agostini HT, Britez G, Ryschkewitsch CF, Stoner GL. 2002. Strains of JC virus in Amerindspeakers of North America (Salish) and South America (Guarani), N-Dene-speakers of New Mexico (Navajo), and modern Japanese suggest links through an ancestral Asian population. American Journal of Physical Anthropology 118: 154–68. Friedlaender JS, Schurr TS, Gentz F, Koki G, Friedlaender F, Horvat G, Babbs P, Cerchio S, Kaestle F, Schanfield M, Deka R, Yanagihara R, Merriwether DA. 2005. Expanding Southwest Pacific mitochondrial haplogroups P and Q. Molecular Biology and Evolution 22: 1506–17. Frisque RJ, White III FA. 1992. The molecular biology of JC virus, causative agent of progressive multifocal leukoencephalopathy. In: Roos RP, editor. Molecular neurovirology. Totowa, NJ: Humana Press. pp 25–158. Frisque RJ, Bream GL, Cannella MT. 1984. Human polyomavirus JC virus genome. Journal of Virology 51(2): 458–69. Guo J, Sugimoto C, Kitamura T, Ebihara H, Kato A, Guo Z, Liu J, Zheng SP, Wang YL, Na YQ, Suzuki M, Taguchi F, Yogo Y. 1998. Four geographically distinct genotypes of JC virus are prevalent in China and Mongolia: Implications for the racial composition of modern China. Journal of General Virology 79(Pt 10): 2499–505. Hatwell JN, Sharp PM. 2000. Evolution of human polyomavirus JC. Journal of General Virology 81: 1191–200. Ho L, Chan SY, Burk RD, Bas BC, Fujinaga K, Icenogle JP, Kahn T, Kiviat N, Lancaster W, Mavromara-Nazos P, Labropoulou V, Mitrani-Rosenbaum S, Norrild B, Radhakrishna Pillai M, Stoerker J, Syrjaenen K, Syrjaenen S, Tay S, Villa LL, Wheeler CM, Williamson A, Bernard H. 1993. The genetic drift of human papillomavirus type 16 is a means of reconstructing prehistoric viral spread and the movement of ancient human populations. Journal of Virology 67: 6413–23. Jobes DV, Chima SC, Ryschkewitsch CF, Stoner GL. 1998. Phylogenetic analysis of 22 complete genomes of the human polyomavirus JC virus. Journal of General Virology 79 (Pt 10): 2491–8. Jobes DV, Friedlaender JS, Mgone CS, Koki G, Alpers MP, Ryschkewitsch CF, Stoner GL. 1999. A novel JC virus variant found in the Highlands of Papua New Guinea has a 21-base pair deletion in the agnoprotien gene. Journal of Human Virology 2(6): 350–8.
Viral Phylogeny and Human Migration in the Southwest Pacific
Jobes DV, Friedlaender JS, Mgone CS, Agostini HT, Koki G, Yanagihara R, Ng TCN, Chima SC, Ryschkewitsch CF, Stoner GL. 2001. New JC virus (JCV) genotypes from Papua New Guinea and Micronesia (type 8 and type 2E) and evolutionary analysis of 32 complete JCV genomes. Archives of Virology 146: 2097–113. Kato A, Kitamura T, Sugimoto C, Ogawa Y, Nakazato K, Nagashima K, Hall WW, Kawabe K, Yogo Y. 1997. Lack of evidence for the transmission of JC polyomavirus between human populations. Archives of Virology 142(5): 875–82. Kato A, Sugimoto C, Zheng HY, Kitamura T, Yogo Y. 2000. Lack of disease-specific amino acid changes in the viral proteins of JC virus isolates from the brain with progressive multifocal leukoencephalopathy. Archives of Virology 145(10): 2173–82. Loeber G, Dories K. 1988. DNA rearrangements in organspecific variants of polyomavirus JC strain GS. Journal of Virology 62: 1730–5. Mahieux R, Ibrahim F, Mauclere P, Herve V, Michel P, Tekaia F, Chappey C, Garin B, Van Der Ryst E, Guillemain B, Ledru E, Delaporte E, de The G, Gessain A. 1997. Molecular epidemiology of 58 new African human T-cell leukemia virus type I (HTLV-1) strains: Identification of a new and distinct HTLV-1 molecular subtype in Central Africa and in pygmies. Journal of Virology 71: 1317–33. Major EO, Amemiya K, Tornatore CS, Houff SA, Berger J. 1992. Pathogenesis and molecular biology of progressive multifocal leukoencephalopathy, the JC virus-induced demyelinating disease of the human brain. Clinical Microbiology Reviews 5: 49–73. Matisoo-Smith E. 2002. Something old, something new: Do genetic studies of contemporary populations reliably represent prehistoric populations of Pacific Rattus exulans? Human Biology 74(3): 489–96. Matisoo-Smith E, Robins JH. 2004. Origins and dispersals of Pacific peoples: Evidence from mtDNA phylogenies of the Pacific rat. Proceedings of the National Academies of Science USA 101(24): 9167–72. Matisoo-Smith E, Allen JS, Ladefoged TN, Roberts RM, Lambert DM. 1997. Ancient DNA from Polynesian rats: Extraction, amplification and sequence from single small bones. Electrophoresis 18(9): 489–96. Merriwether DA, Hodgson JA, Friedlaender FR, Allaby R, Cerchio S, Koki G, Friedlaender JS. 2005. Ancient mitochondrial M haplogroups identified in the Southwest Pacific. Proceedings of the National Academy of Sciences, USA 102:13034–9. Miranda JJ, Sugimoto C, Paraguison R, Takasaka T, Zheng HY, Yogo Y. 2003. Genetic diversity of JC virus in the modern Filipino population: Implications for the peopling of the Philippines. American Journal of Physical Anthropology 120: 125–32. Miura T, Fukunaga T, Igarashi T, Yamashita M, Ido E, Funahashi S, Ishida T, Washio K, Ueda S, Hashimoto K, Yoshida M, Osame M, Singhal BS, Zaninovic V, Cartier L, Sonoda S, Tajima K, Ina Y, Gojobori T, Hayami M. 1994. Phylogenetic subtypes of human T-lymphotropic virus type I and their relations to the anthropological
background. Proceedings of the National Academy of Sciences USA 91: 1124–7. Padgett BL, Walker DL, ZuRhein GM, Eckroade RJ, Dessel BH. 1971. Cultivation of papova-like virus from human brain with progressive multifocal leucoencephalopathy. Lancet 1(7712): 1257–60. Pavesi A. 2003. African origin of polyomavirus JC and implications for prehistoric human migrations. Journal of Molecular Evolution 56(5): 564–72. Pavesi A. 2004. Detecting traces of prehistoric human migrations by geographic synthetic maps of polyomavirus JC. Journal of Molecular Evolution 58(3): 304–13. Rezza G, Danaya RT, Wagner TM, Sarmati L, Owen IL, Monin P, Andreoni M, Suligoi B, Ensoli B, Pozio E. 2001. Human herpesvirus-8 and other viral infections, Papua New Guinea. Emerging Infectious Diseases 7(5): 893–5. Ryschkewitsch CF, Friedlaender JS, Mgone CS, Jobes DV, Agostini HT, Chima SC, Alpers MP, Koki G, Yanagihara R, Stoner GL. 2000. Human polyomavirus JC variants in Papua New Guinea and Guam reflect ancient population settlement and viral evolution. Microbes and Infectious Disease (9): 987–96. Saruwatari L, Sugimoto C, Kitamura T, Ohno N, Sakai E, Shresta P, Hoa BK, Phi PT, An HP, Tuyet NT, Honjo T, Kobayashi N, Zheng HY, Takasha T, Yogo Y. 2002. Asian domains of four major genotypes of JC virus, Af2, B1-b, CY and SC. Archives of Virology 147(1): 1–10. Stoner GL, Jobes DV, Ryschkewitsch CF, Fernandez- Cobo M. 2000. Evolutionary analysis of the human polymavirus JC. GenBank 2 (bases 1 to 4854), submitted (23 June 2000), Neurotoxicolocy Section, National Institute of Neurological Disorders and Stroke, Bethesda, MD. Sugimoto C, Kitamura T, Guo J, Al-Ahdal MN, Shchelkunov SN, Otova B, Ondrejka P, Chollet J, El-Safi S, Ettayebi M, Gresenguet G, Kocagoz T, Chaiyarasamee S, Thant KZ, Then S, Moe K, Kobayashi N, Taguchi F, Yogo Y. 1997. Typing of urinary JC virus DNA offers a novel means of tracing human migrations. Proceedings of the National Academy of Sciences USA 94: 9191–6. Sugimoto C, Hasegawa M, Zheng HY, Demenev V, Sekino Y, Kojima K, Honjo T, Kida H, Hovi T, Vesikari T, Schalken JA, Tomita K, Mitsunobu Y, Ikegaya H, Kabayashi N, Kitamura T, Yogo Y. 2002. JC virus strains indigenous to Northeastern Siberians and Canadian Inuits are unique but evolutionally related to those distributed throughout Europe and Mediterranean areas. Journal of Molecular Evolution 55: 322–35. Takao S, Ishida T, Bhatia KK, Saha N, Soemantri A, Kayame OW. 2000. Seroprevalence of human T-lymphotropic virus type 1 in Papua New Guinea and Irian Jaya measured using different western blot criteria. Journal of Clinical Virology 16(2): 129–33. Tanggo Y, Gultom SP, Simanjuntak T, Sibuea WH, Matsuzaki H, Yamaguchi K. 2000. Human T lymphotropic virus I in Indonesia. Very low seroprevalence in the Jakarta area: Antibodies in healthy blood donors and in various nonhemotological diseases. Intervirology 43(2): 77–9. 179
regional studies and conclusion
Umene K, Sakaoka H. 1997. Populations of two Eastern countries of Japan and Korea and with a related history share a predominant genotype of herpes simplex virus type I. Archives of Virology 142: 1953–61. Wagenaar TR, Chow VTK, Buranathai C, Thawatsupha P, Grose C. 2003. The out of Africa model of varicellazoster virus evolution: Single nucleotide polymorphisms and private alleles distinguish Asian clades from European/North American clades. Vaccine 21: 1072–81. Walker DL, Frisque RJ. 1986. The biology and molecular biology of JC virus. In: Salzman NP, editors. The papovaviridae, Vol. 1, The polyomaviruses. New York: Plenum Press. Pp 327–77. Watts KJ, Thompson CH, Cossart YE, Rose BR. 2002. Sequence variation and physical state of human papilloma virus type 16 cervical cancer isolates from Australia and New Caledonia. International Journal of Cancer 97: 868–74. Wooding S. 2001. Do human and JC virus genes show evidence of host-parasite codemography? Infection, Genetics and Evolution 1: 3–12. Yanagihihara R. 1994. Geographic-specific genotypes or topotypes of human T-cell lymphotropic virus type I
180
as markers for early and recent migrations of human populations. Advanced Virus Research 43: 147–86. Yanagihara R, Saitou N, Nerurkar VR, Song KJ, Bastian I, Franchini G, Gajdusek DC. 1995. Molecular phylogeny and dissemination of human T-cell lymphotropic virus type I viewed within the context of primate evolution and human migration. Cellular and Molecular Biology 41: S145–61. Yanagihara R, Nerurkar YR, Scheirich I, Agostini HT, Mgone CS, Cui X, Jobes DV, Cubitt CL, Ryschkewitsch CF, Hrdy DB, Friedlaender JS, Stoner GL. 2002. JC virus genotypes in the Western Pacific suggest Asian mainland relationships and virus association with early population movements. Human Biology 74(3): 473–88. Zheng HY, Kitamura T, Takaska T, Chen Q, Yogo Y. 2004a. Unambiguous identification of JC polyomavirus strains transmitted from parents to children. Archives of Virology 149: 261–73. Zheng HY, Zhao P, Suganami H, Ohasi Y, Ikegaya H, Kim JC, Sugimoto C, Takasaka T, Kitamura T, Yogo Y. 2004b. Regional distribution of two related Northeast Asian genotypes of JC virus, CY-a and -b: Implications for the dispersal of Northeast Asians. Microbes and Infection 6: 596–603.
12 Origins of Plant Exploitation in Near Oceania: A Review Robin Allaby
Introduction The majority of plant species exploited by the peoples of Near and Remote Oceania can also be found in Southeast Asia. The earliest European explorers noted this during the Cook voyages, and this led to a botanical tradition that influenced later interpretations of Pacific agricultural origins (Yen, 1991a). For example, Sauer (1952) described Southeast Asia as a “cradle of agriculture,” and subsequently, a Southeast Asian origin for crops and cultivation practices became the default explanation for Pacific agriculture. However, recent advances in archeobotany, and particularly its molecular aspect, have challenged a Southeast Asian origin for many plant species. There has been a shift towards the attribution of cultivation practices to early Holocene and even Pleistocene societies there that had previously been considered technologically less advanced “hunter-gatherers.” Perhaps the emergent model for Near Oceania is a harbinger for reconsidering agricultural origins in other regions. In order to understand the complex processes that gave rise to plant exploitation in Near and Remote Oceania, this review begins by considering the extant nature of plant exploitation in Oceania, then the recent advances in archeobotany and molecular archeobotany techniques which have changed our view of Pacific agricultural origins, before considering those plants which are now thought to have been domesticated in Near Oceania and those that were introduced later. The emphasis is on the botanical sources of evidence rather than the anthropological frameworks that had formed a cornerstone for the interpretation of cultivar origins.
Tropical Vegeculture and Arboriculture Plant exploitation in Oceania is made up of both vegeculture and arboriculture. The vegeculture primarily exploits corms, rhizomes, and roots of species of taro, yam, and sweet potato, all “storage organ plants” (for a comprehensive classification of these types of plants see Hather, 1994b). While temperate environments have supported agrarian systems that exploit cereals for starch, people in tropical environments tend to exploit storage organ crops. Storage organ crops may have been preferable to cereals because they are very amenable to propagation and are visibly obvious sources of food. In contrast to the recalcitrant cereals, it is likely that the ease of cultivation of the wild storage organ crops led to their exploitation wherever wild forms could be found (Harris, 1969). It has been a matter of debate whether storage organ crops are indigenous flora to aseasonal humid tropic conditions. Hawkes (1969) argued that storage organs are an adaptation to dry seasons, and therefore evolved outside the aseasonal humid tropics. Hather (1996) countered that storage organs may be an adaptation to longevity rather than seasonality, allowing continued plant growth through a horizontal plane. In this way, the Monocotyledonae, incapable of producing wood, may have evolved an alternative strategy to the boreal habit for supporting continued growth. In Hather’s argument, storage organ plants may be considered as native to aseasonal humid tropic environments, rather than likely introductions. However, despite New Guinea’s status as the wettest land area on the planet, there is actually a dry season in particular regions, so both arguments may be
181
regional studies and conclusion
true and the crops need not be considered alien to the New Guinea humid tropics. Therefore, these plants can be viewed as adapted native species to the New Guinea humid tropics, and likely to be encountered by Pleistocene settlers in this region. The second form of plant subsistence in Oceania is arboriculture. A wide range of tree species is exploited for their nuts and fruits, although a half dozen or so genera are of primary importance. The contemporary arboriculture is described for the Mussau Islands by Lepofsky (1992) and Santa Cruz by Yen (1974). The trees sweet almond (Canarium), coconut (Cocos), cutnut (Barringtonia), sea almond (Terminilia), screw pine (Pandanus), sandal wood (Santalum), and Pometia are all important as domesticates. A third category of exploited plants, often grouped within the arboriculture for convenience (e.g. Yen, 1995), includes bananas (Musa sp.) and sugar cane (Saccharum sp.). Since these crops are neither trees nor root crops, they are considered under the category “herbiculture” in this review for convenience. There is a continuum of plant exploitation in Near Oceania that ranges from collecting wild tubers and fruits, to growing wild forms, to growing wild forms in technologically advanced—such as irrigated—systems, to growing domesticated forms with and without such systems. Only the latter may truly be referred to as agriculture, and, for example, frequently past plant exploitation by Lapita peoples was considered “horticulture” (Kirch, 1997). Most of the plants exploited today are “domesticated” in that they have undergone genetic modification resulting in a phenotypic difference to the wild form. However, the domesticate forms are not all dependent on human intervention for propagation as is the case for most temperate cereal species. The phenotypic change resulting from the domestication process may be a loss of toxic substances (e.g. Colocasia esculenta and Barringtonia edule), or more subtle, such as the thinning of nut walls (e.g. Pandanus odoratissimus), or softening and increased potability of tissue (Piper methysticum). More extreme phenotypic changes that do necessitate human intervention for propagation include the trend toward parthenocarpy, the formation of fruit without seeds (e.g. Musa sp., Artocarpus atilis).
Sources of Evidence New sources of evidence have illuminated aspects of plant exploitation in the Pacific in formerly unattainable ways. In particular, vegeculture crops that are typically subject to very poor archeological preservation have now been characterized to a much greater degree than had been previously possible.
182
Archeobotany Traditionally archeobotanists have analyzed pollen grains and tough macrofossils such as seeds, nuts, and woody tissue such as endocarps, and as a result, non-tree species have been underrepresented. For example, the 50 or so pollen records of New Guinea favor tree species (Haberle, 1994). The remains of plants from the Arawe Islands spanning 4000 to <1000 years BP were almost exclusively tree species identified from endocarps and seeds, including Aleurites moluccana, Canarium sp., Cocos nucifera, Cordia subcordata, Dracontomelon dao, Pandanus sp., Terminalia sp. and the cycad Cycas circinalis (Matthews and Gosden, 1997). Similarly in the Mussau Islands, practically all the arboriculture of today was identified in remains from 3600 years BP (Kirch, 1989). The Balof 2 site of New Ireland yielded evidence of important trees such as Canarium (Allen et al., 1989). Several techniques have now been developed that can identify the typically more poorly preserved species of vegeculture and herbiculture. The scanning electron microscope can reveal species level distinctions in the morphology of cells and the organization of charred parenchymous storage tissue with Dioscorea bulbiferous, Cordyline terminalis and Ipomoea batatas (Hather, 1991, 1994a). The technique has been used with prehistoric samples of Ipomoea batatas in central Polynesia (Hather and Kirch, 1991). A second approach involves detecting starch grains in the soil matrix. An early and controversial report identified starch grains originating from Colocasia 28,000 years BP in the Solomon Islands (Loy et al., 1992), but the technique is now well established and regularly used to identify root crops (Piperno et al., 2000). There are some concerns regarding starch grain taphonomy (Haslam, 2004). Firstly, the grains are degraded by heat and consequently are not found in charred contexts. Secondly, degradation can occur by microorganisms, and small grains less than 5 µm diameter decrease in number over time (Therin et al., 1998; Haslam, 2004), but nevertheless, some grains do survive (Lentfer et al., 2002). The presence of plant species based on starch grain identification appears to be reliable, but smaller grains (<5 µm) tend to be less defined and frequently confused with other bodies (Haslam, 2004). Consequently, studies based on starch grains of small sizes should be cautiously interpreted. Despite their small size (5 µm), the starch grains of Colocasia esculenta Denham et al. 2003 argue that they are identifiable using this technique. The grains of sweet potato (Ipomoea batatas) are a little larger (~10 µm) and those of yams (Disocorea sp.) are all larger (20–70 µm) with the exception of D. esculenta (~5 µm) (Loy et al., 1992: figure 10). A third important new approach is the identification of phytoliths, which are opaline silica bodies formed by
Origins of Plant Exploitation in Near Oceania: A Review
living plants. A number of plant families have been found to produce large amounts of phytoliths (Piperno, 1991), including the important members of the Monocotyledonae; Musaceae, Palmae, and Zingiberaceae. However, the Araceae and the Dioscoraceae, which contain the taro and yam genera respectively, are not known for their phytolith production. Non-monocot families that are heavy phytolith producers include the Bursereaceae, Curcubitaceae, Moraceae, and Piperaceae.
Molecular Archeobotany A range of new biomolecular approaches to phylogeny, phylogeography and population genetics methods on genetic data sets have been applied to domesticated plant origins. Some of the molecular data types are useful for phylogenetic analysis, allowing for the identification of primitive (plesiomorphic) and derived (apomorphic) character states, which combined with phylogeographic data can yield information about the direction of crop movements in the past as well as evolutionary events. Other data types are useful as measures of genetic diversity of populations, which can be a powerful approach to identifying centers of diversity, which may be inferred to equate to centers of crop origin. Nucleic acid sequences are particularly appropriate for phylogenetic analysis. However, there is no single genetic locus that is the obvious choice for study. Plant mtDNA is highly recombinant and has a very low rate of change, unlike the mammalian condition (Palmer and Herbon, 1988, Wolfe et al., 1987). The low rate of change of plant organelle DNA makes nuclear DNA a preferable target of study, but typical rates of change in nuclear genes are often still too slow to be useful for studies dealing with the short timescales associated with the Holocene. To address this problem, resourceful approaches have been developed which single out nuclear polymorphic areas of the genome. One such approach has recently been used successfully to track the history of Lageneria siceraria (bottle gourd) in the Pacific (Clarke et al., 2005). Microsatellites, with their typically rapid rate of change, have also provided a useful approach to investigate origins, as with Ipomoea batatas (Zhang et al., 2000a; Hu et al., 2003). However, a large number of loci (between 10 and 20) need to be included to build a reliable evolutionary picture (Stumpf and Goldstein, 2001). Chromosomal rearrangements also make robust characters for phylogenetic reconstruction and were used in studies of taro and sugar cane origins (Brandes, 1958; Coates et al., 1988). Quick genomic scanning techniques such as amplified fragment length polymorphism (AFLP; Vos et al., 1995) and random amplified polymorphic DNA (RAPD; Williams et al., 1990) are also attractive ways to identify informative diversity. The characters are anonymous,
meaning that we are not sure from where they come from in the genome. Several important plant species of the Pacific have been examined using genomic scanning techniques, including Colocasia (Irwin et al., 1998; Kreike et al., 2004), Dioscorea (Choi et al., 2002; Ramser et al., 1997), and Ipomoea (Harvey et al., 1997; Huang et al., 2002; Zhang et al., 2000ab, 2004). However, this kind of data is not appropriate for phylogenetic analysis (Allaby and Brown, 2003). When AFLP data are entered into a phylogenetic tree-building algorithm, demic units such as populations or even domesticated species are identified as clades rather than evolutionary events as defined by synapomorphisms. As a result, populations or domesticate species may be falsely identified as monophyletic. The technique cannot effectively resolve introgression events or even population hybridization events. Plant phytochemicals can be used as characters to build an evolutionary history (Judd et al., 1999). For instance, most flowering plants utilize anthocyanin for color pigmentation in flowers, but the members of the Caryophyllales utilize betalain instead, making this a derived character. The relative amounts of chemicals expressed in plant tissues can also be used as characters. Both the presence/absence and relative quantity approaches can have problems. For instance, the relative levels of different kava lactones in Pacific kava may be strongly influenced by environmental factors (Lebot and Siméone, 2004), and flavanoids used as species markers have been spontaneously expressed in hybrids, but not in either parent species (Grivet et al., 2004).
History of Plant Exploitation in the Pacific Near Oceania as a Plant Domestication Center Near Oceania is well situated to be a source of potential domestications. Floristically it is rich because it is at the meeting point of three continental plates, with very different floras originating from two of these. Southern New Guinea represents the northern edge of the Australian plate. It meets the Pacific plate to the north, giving rise to the highlands of New Guinea. To the west these meet the Eurasian plate, giving rise to the Wallace Line in the middle Miocene to Pliocene (Whitmore, 1981). Today’s floral biogeography reflects a dual input, with Australian flora tending to occur in southern New Guinea and Southeast Asian flora in the northern areas (Heads, 2001). In the Pleistocene, New Guinea was joined to Australia forming the Sahul continent, which may have aided floral dispersal northwards. Similarly, lower sea levels meant that floral dispersal from the Sunda shelf was over shorter tracks of water than exist today. However, in the area currently under water between New Guinea and
183
regional studies and conclusion
Australia, savannah occurred which may have acted as a barrier to swamp and rainforest flora for long time periods (Jones and Bowler, 1980). The first suggestion that Near Oceania might be a domestication center came from cytological studies indicating that sugar cane (Saccharum officinarum) originated in New Guinea (Brandes, 1958). A multiple origin for coconut (Cocos nucifera) was also suggested through the identification of multiple biogeographic centers (Sauer, 1971). The archeobotanical records of two archeological sites in particular, the Arawe Islands (Gosden, 1992; Matthews and Gosden, 1997) and Mussau (Kirch, 1989), have provided evidence of plant exploitation by the Lapita peoples and their immediate predecessors. Linguistics has been used to reconstruct various names of species in Proto-Oceanic to give a more complete list of the plants used by the Lapita people (Kirch, 1989: Table 7.2). A central question is how much of this horticulture was adopted by the Lapita people on arrival in Near Oceania from the indigenous peoples, and how much was transferred from Southeast Asia. Evidence for arboriculture before the Lapita social horizon has been found at several sites (Gosden, 1995). The earliest signs of plant exploitation found are of Canarium, based on nut remains from 14,000 years BP in the Sepik-Ramu area (Yen, 1990), and then in Pamwak, followed by a range of sites on mainland New Guinea about 9,000 years ago (Gosden, 1995). A more complete range of arboriculture species was found in the mid Holocene, circa 6,000 years BP, substantially prior to the arrival of the Lapita complex in the Sepik region, including Cocos, Canarium, Pandanus, Pometia, Aleurites, and Areca (Swadling et al., 1991). Consequently, New Guinea and its surrounding region have been considered a domestication center for some time now in terms of arboriculture. The sophistication of plant exploitation in the early Holocene has been questioned (Spriggs, 1996). Undoubtedly, the earliest exploitation involved foraging wild trees for nuts rather than intense cultivation leading to the genotypic modification of species that could be termed “domesticated.” If domestication is the hallmark of sophistication in plant exploitation, then it follows that methods that identify the locality of the domestication process are of paramount importance to address concerns about the level of sophistication of plant exploitation in the early Holocene. There are two possible approaches to establish the location of a crop’s domestication. The first is through archeobotany to observe directly the process of phenotypic change such as the progressive thinning of nut walls, as with Pandanus species (Golson, 1991). The second is through molecular archeobotany, which can be used to identify current centers of genetic diversity of wild and domesticate species, and to infer the location of the most primitive domesticate types by phylogeography relationships between wild and domesticate forms.
184
Evidence of early plant exploitation from archeobotany has been particularly abundant at the Kuk Swamp site in the Wahgi Valley of New Guinea (Denham et al., 2003; Golson, 1977). The Kuk Swamp evidence shows that from the early to mid Holocene, there has been sophisticated cultivation practiced in the New Guinea Highlands that had not previously been attributed to early Near Oceanic inhabitants. Kuk also provides evidence for the extension of early plant exploitation to those species that are not easily visible in the archeobotanical record, such as Colocasia sp. and Musa sp., as well as for cultivation from structural features such as irrigation channels. The site occurs in the Highlands at an elevation of 1560 m above sea level where the climate is lower montane, which is not the most suitable for an agrarian existence. This may represent the development of agriculture in the Highlands through an edge effect (Denham, 2004b; also see taro below), or a movement of agriculture up from the lowlands (Hope and Golson, 1995). The earlier part of the site excavation is divided into three phases radiocarbon dated to 10,220–9,910, 6,950–6,440, and 4,840–3,980 years BP, respectively (Denham et al., 2003). The earliest of these, phase 1, has stake holes, pits, and runnels consistent with a practice of irrigated cultivation, although this is a very cautious interpretation by the workers, and it is clear that these features are not considered definitive of agricultural origin (Denham, 2004b). Phase 2 has more elaborate indications of cultivation including mound structures, and phase 3 again increases in complexity with a series of rectilinear ditches. In the following sections crops that have been identified as primary, domesticated within the Near Oceania region on the basis of archeobotany and molecular archeobotany, are discussed.
Primary Vegeculture Taros (Colocasia sp.). Several closely related genera of the Araceae family are called “taro,” including Colocasia, Alocasia, and Cyrtosperma. Of these Colocasia esculenta is the most important subsistence crop. The biogeographical distribution of this group is strongly suggestive of an origin on the Indian subcontinent after its separation from Gondwana, followed by a movement through Southeast Asia and into the Sahul during the Pliocene (Renner and Zhang, 2004). The Colocasia genus has species sprawling from the Indian subcontinent through to Sahul including C. fallax and C. affinis in northern and northeastern India, C. gigantea spreading from India through to Indonesia, and wild forms of C. esculenta being found from Southeast Asia through to Polynesia (Matthews, 2004). The wide distribution of wild C. esculenta is possibly the result of human dispersal (Matthews, 2004), although avian dispersal may
Origins of Plant Exploitation in Near Oceania: A Review
also be responsible (Lebot, 1999). A chloroplast or similar based phylogeny of all the Colocasia species could help pinpoint where C. esculenta is likely to have arisen. While it is likely that dispersal of the wild form into Polynesia was by human hand, the forms found in New Guinea were present substantially before this spread, having been present for the past 9,000 years based on pollen evidence from the lowlands (Haberle, 1995). The antiquity of taro in Near Oceania is also supported by the existence of the parasite species Tarophagus proserpina which has a range that extends no further west than eastern New Guinea (Matthews, 2003a). Domesticated taro is less acrid and has a larger corm than the wild form. Two types are often recognized, eddoe and dasheen, sometimes classified as subspecies C. esculenta var. antiquorum and var. esculenta although intermediate phenotypes occur. The dasheen type is found in most places, and is grown in rheotrophic (flooded) conditions, while the eddoe type is frequent in Malaysia and grown in ombrotrophic (rain-fed) conditions (Lebot et al., 2004). The two varieties could be the result of independent domestications (Lebot, 1999), however there is no molecular support for this assertion yet. AFLP and RAPD analyses have been applied to C. esculenta from Oceania and Southeast Asia. The highest genetic diversity of cultivated taro is to be found in Indonesia (Lebot and Aradhya, 1991; Irwin et al., 1998; Kreike et al., 2004; Lebot et al., 2004). Superficially this may appear to be the result of a center of origin, but it may be the result of a meeting zone from two or more separate centers, one in Southeast Asia, a second in Papua New Guinea, and possibly a third within Indonesia itself. The Asian and Pacific cultivars can be distinguished by both AFLP and isozyme data (Lebot and Aradhya, 1991; Kreike et al., 2004; Lebot et al., 2004). Kreike et al. (2004) further detected groupings that correspond to demic groupings of islands and countries. Indonesia is a political division that includes several demic groups of taro from different island systems, including West Papua, which in part explains the high genetic diversity observed there. In the light of this, the diversity in New Guinea is actually equal to or greater than that further west in Southeast Asia. Triploid varieties of C. esculenta are rare in the Pacific. It has been argued that they originated in Southeast Asia based on ribosomal DNA evidence (Matthews, 1990, 2004). However, cytological work (Coates et al., 1988; Yen, 1991a) may have been unfairly dismissed. Three lineages of triploid taro were identified that differed by large chromosomal features of metacentricity and acrocentricity (Coates et al., 1988). Two triploid types were observed in New Zealand, typed as 3mmm, 7mAA, 9mmm, and 3AAA, 7AAA, 9AAA, where the number refers to the
chromosome, “m” to a metacentric state and “A” to an acrocentric state. A likely diploid precursor to the first of these, restricted to New Guinea and Australia was identified as 3mm, 7mA, 9mm, while the latter triploid occurred in Southeast Asia. It is likely then that the former triploids formed in situ in the Pacific. Triploid formation in taro is not unusual. About 1 in 170 plants form triploids, regardless of whether they are selfed or crossed, with the seed parent usually donating the double genome (Isshiki et al., 1999). Additionally, taro triploids often display hybrid vigor, which may favor their survival, but they are inherently infertile and so must be vegetatively propagated thereafter (Isshiki et al., 1999). The molecular archeobotanical evidence for C. esculenta favors the possibility of independent domestications in Near Oceania and Southeast Asia. Even within Eastern Asia, evidence based on ribosomal DNA indicates possible dual origins (Matsuda and Nawata, 2002). However, an in-depth phylogenetic study of the genus including the wild forms and closely related genera has yet to be carried out. Archeobotany supports a Near Oceania domestication of taro in the early Holocene. Starch grain analysis indicated the presence of Colocasia esculenta on stone tools from phases 1 and 2 of the Kuk Swamp site. Taro is a lowland crop that almost certainly would have had to be cultivated in such a highland site. In order for granules to be detected they must have been deposited in substantial amounts, again suggesting a domesticated variety. This supports the notion of an edge effect in which the pressure for cultivation is not at the center of the distribution of the wild crop, but rather towards its edges where it is not likely to grow in large stands (Bellwood, 1996). The evidence from taro in this way supports the idea that agriculture may have developed in highland areas rather than in lowlands (Denham, 2004b; Hope and Golson, 1995). In sum, C. esculenta is the clearest example of a crop that was until recently considered a Southeast Asian domesticate, but is now also recognized as an indigenous domesticate of New Guinea. The remaining taros have received much less attention. The cultivated forms of Alocasia macrorrhizus and Cyrtosperma chamissonis do not occur on mainland New Guinea, but do on the Bismarck Archipelago and Solomon Islands. The wild forms, however, do occur on mainland New Guinea (Yen, 1995). On the basis of morphological traits, Hay and Wise (1991) argue that A. macrorrhizus is an introduction from Asia, which appears to conflict with the presence of the wild form in New Guinea (Yen, 1995). C. chamissonis, however is accepted as possibly being derived from within New Guinea (Hay and Wise, 1991; Yen, 1991a, 1995). The early peoples of Near Oceania may have exploited A. macrorrhizus and C. chamissonis if the wild forms were present.
185
regional studies and conclusion
Yams (Dioscorea sp.). Members of the Dioscoraceae can be found across the southern hemisphere, with over 600 species. Five are important in Oceania: D. alata, D. bulbifera, D. esculenta, D. nummularia, and D. pentaphylla. Of these, D. alata, the greater yam, is a major subsistence crop, and occurs from Africa to Hawaii. No wild form of D. alata is known to exist, nor is the cultivated form known to hybridize with any other Dioscorea. Originally thought to be a crop of Southeast Asian origin, the major center of cultivar diversity was identified to be New Guinea and the Solomon Islands (Coursey, 1976). More recently, AFLP analysis has confirmed a center of genetic diversity in New Guinea (Choi et al., 2002), implying a New Guinea origin, although identifying the wild progenitor would help. Two Australian species, D. hastifolia and D. tranversa, are thought to be closely related to D. alata, while the third major Australian yam species, D. sativa, is thought to be closely related to D. bulbifera (Yen, 1991a, 1995). Again, this distribution implies an early spread of the genus from New Guinea. The bitter yam D. bulbifera also has a widespread distribution. A restriction fragment length polymorphism (RFLP) analysis of D. bulbifera chloroplast DNA revealed that plants from Africa, Asia, and the Pacific formed three groups (Matthews and Terauchi, 1994). The Pacific group included plants from Australian and Tonga with identical haplotypes implying an intriguing connection between the two areas. Such a distribution could have been the result of an emanation from New Guinea to both areas. Despite the small size of this study (15 plant accessions and 9 restriction enzymes), it indicates further research is likely to be fruitful. Yams have been directly observed in the archeobotanical record through starch grain analysis on Dauar Island in the Torres Strait, tentatively identified as D. pentaphylla indicative of subsistence prior to 1,200 years BP (Parr and Carter, 2003). The large size of yam starch grains implies that more are likely to be found in the archeobotanical record. In summary, the molecular archeobotanical data imply that yams were present in Near Oceania in the early Holocene, but this has yet to be confirmed by the archeobotanical record. Kava (Piper methysticum). Kava (Piper methysticum) is a mild narcotic used throughout Oceania. In Near Oceania, the plant is of more limited significance in New Guinea and the Solomon Islands where a few coastal communities use it for ceremonial purposes. In Fiji and Vanuatu it is commonly used. In Polynesia modern kava usage is restricted to Tonga and Samoa and more recently Hawaii, but in the past usage was widespread (Lebot and Siméoni, 2004). It is a domesticated species derived from P. wichmanii. P. methysticum is a sterile plant, and so
186
represents a domesticated species entirely dependent on human intervention for propagation. P. wichmanii has a distribution that stretches from New Guinea to Vanuatu, with a center of genetic diversity in New Guinea (Lebot, 1991; Lebot et al., 1991, Lebot et al., 1999). There is no doubt that kava was domesticated in Melanesia, but exactly where in the region is debatable. One possibility that it was domesticated in New Guinea and spread eastwards. The alternative is that it was domesticated in northern Vanuatu. New Guinea is favored by being the center of the wild distribution, but only four morphotypes of P. methysticum occur there. Northern Vanuatu is on the edge of the wild distribution, but has 80 P. methysticum morphotypes, and so clearly represents the center of diversity for the domesticated form (Lebot et al., 1992). Linguistic evidence had originally suggested that New Guinea was the more likely source of domestication, to account for the basal proto-Oceanic term *kawa, taken to mean P. methysticum, occurring in New Guinea (Pawley and Green, 1973). Lynch (2002) argues that a linguistic hypothesis can be constructed which does not conflict with the current pattern of diversity. In Lynch’s hypothesis *kawa is actually a derived term, and another term *kawaRi is basal. The Proto-Oceanic term *kawaRi was generally applied to psychoactive roots, including a wild ginger (Zingiber zerumbet) and P. wichmanii. This term then became applied to kava after its domestication. The term then spread to Fiji and Polynesia as *kawaR, whereupon several derivative transformations occurred. It is then suggested that the Polynesians introduced the term to New Guinea in western voyages as *kawa. The genetic diversity of P. methysticum is very low, based on AFLP (Lebot et al., 1999) and isozyme data (Lebot et al., 1991b). Isozyme analysis identified three patterns of isozyme variation (zymotypes) in P. methysticum: Papua New Guinea and Vanuatu each had two of the zymotypes, one of which they shared, while Vanuatu also shared a zymotype with the rest of the Pacific, which only had one. The AFLP analysis, although carried out on a limited accession number (only 22), also indicated that the Papua New Guinea form was most distant to Polynesia. Similar relationships have also been established using chemotypical data (Lebot and Lévesque, 1989). In this case the relative quantities of six kavalactones, the chemicals responsible for the psychotropic effects of kava, were measured in different cultivars. One chemotype was found to occur only in Papua New Guinea and Vanuatu, and accounts for all P. methysticum in Papua New Guinea. However, three other chemotypes are present in Vanuatu, Fiji, and Polynesia. The chemotypical data also support Vanuatu as an origin of domestication, because the chemotypes of P. wichmanii found on New Guinea are not the most closely related to the types found in P. methysticum (Lebot and Siméoni, 2004).
Origins of Plant Exploitation in Near Oceania: A Review
In summary, the genetic and chemical data now support a Vanuatu origin for domestication, and an import of kava to New Guinea from Vanuatu, not Polynesia or Micronesia. The conclusion that Vanuatu is the domestication center for kava has the corollary that domestication was relatively recent, probably 3,000 years ago, to coincide with early human occupation of the island. Hence, kava is an interesting example of a crop that is indigenous to Island Melanesia, but in all likelihood domesticated by the peoples of the Austronesian expansion. It may be the case that the indigenous peoples of New Guinea may have exploited the wild root, but they did not domesticate it. Other Root Crops. The site at Kuk Swamp also provides some surprising evidence of other root crops. In particular, Denham et al. (2004) report the occurrence of Zingiber phytoliths at pre-phase 1 levels (10,220–9,910 years BP). Although wild ginger (Zingiber zerumbet) grows in the forests of the Pacific, it is thought to be a naturalized introduction (Whistler, 1991). Possibly, on confirmation of the phytolith evidence this theory might need to be revised, unless the introduction was by the early peoples of Near Oceania.
Primary Arboriculture Sweet Almond (Canarium sp.). The Canarium group of tree species is exploited for nuts. The Burseraceae probably originated on the North American continent, not Gondwanaland (Weeks et al., 2005). Consequently, the Canarium species would have reached Near Oceania possibly as early as the Pliocene from the Laurasian continental plate rather than the Indian or Australian continents. Seven domesticated species of Canarium occur throughout Near Oceania, with a biogeography suggestive of a New Guinea center of origin (Yen, 1991a: figure 5). The distribution of the different species is disjunct. In the “Vulgare” group C. ovatum occurs in the Phillipines, C. vulgare occurs in the islands of Indonesia, while C. indicum has the broadest distribution stretching from eastern Indonesia through New Guinea to Vanuatu. The “Maluense” group generally has a more eastern distribution, with C. lamii occurring in New Guinea, C. salomonense occurring from eastern New Guinea to the Solomon Islands, and C. harveyii occurring from the Solomon Islands through to western Polynesia. Additionally, the seventh species C. decamanum, a giant fruit-producing tree which fits into neither of these two groupings, occurs from Borneo through to the western Bismarck Archipelago. C. indicum is the major economic species supplying most subsistence. Although this distribution is suggestive of an expansion out of New Guinea, there are other Canarium species from adjacent regions of Australia and southern and Southeastern Asia, which may help elucidate the
overall picture. To the south, C. baileyanum is exploited in northern Australia by the Aboriginies. To the west, C. littorale occurs in Sumatra (Maloney, 1996), and C. zeylanicum has been exploited as a Mesolithic crop in Sri Lanka (Kajale, 1989). Interestingly, C. littorale, takes a pleisiomorphic position in the Canarium clade of the Burseraceae relative to C. zeylanicum, C. indicum and C. vulgare (Weeks et al., 2005), which might not be expected from the general eastward movement of the genus. Some of the oldest archeobotanical finds in the Pacific are of Canarium. Nut husks from the Sepik-Ramu area dated to 14,000 years BP, identified as C. salomonese, give the earliest indication of Near Oceanic arboriculture (Gorecki cited in Yen, 1991b), followed by Manus Island at 11,000 years BP (Frederickson et al., 1993). However, C. indicum dominates the archeobotanical record of Lapita sites, and also occurs in the Sepik-Ramu area (Swadling et al., 1991). These old recoveries of Canarium not only demonstrate exploitation, but also suggest the process of domestication already in action in the late Pleistocene. Coconut (Cocos nucifera). The coconut (Cocos nucifera) was the first clear example of a species that was domesticated multiple times across the Pacific (Sauer, 1971). The global distinctiveness of cultivars is clearly visible from RFLP data, with groups occurring for Africa and the Indian Ocean, and the Pacific (Lebrun et al., 1998). There is no clear distinction with this technique between the cultivars of Malaysia and Near Oceania. The wild form is well adapted for oceanic dispersal, with its large well-protected fruit and long germination time. Domesticates differ from the wild form in many features including fruit morphology, germination time, and plant habit (Harries et al., 2004). The wild form is a strand plant, exclusively found in coastal environments. The ability of the coconut to disperse so freely has led to a fairly recalcitrant evolutionary history. Several areas in Southeast Asia have been identified as domestication centers (Harries et al., 2004: figure 1). Harries et al. (2004) make the distinction that early domesticate forms of the plant would become reproductively isolated from the wild strand plants as a result of their being moved inland. Wild forms of the plant occur throughout the Pacific in strand environments. These wild forms can be distinguished from cultivars using the dimensions and features of the fruit in a “fruit component analysis” (Harries, 1978). Harries et al. (2004) predict that domesticated forms, spread from Southeast Asia through the Pacific by the Austronesian-speaking peoples, will generally occur inland on Pacific islands but will also have introgressed with the local wild coconut forms. As a result one should expect to find two or three forms on most Pacific Islands. In support of this prediction, the two forms present in Samoa are explained in this way (Harries, 1978). This pattern has also been observed in the Mussau Islands
187
regional studies and conclusion
of the Bismarck Archipelago (Lepofsky, 1992). Considerable molecular techniques have been applied to studying Pacific coconut, usefully summarized in Harries et al. (2004). The molecular data are as yet inconclusive, although markers that distinguish Tongan and Fijian varieties from the rest of the Pacific have been identified, as well as markers specific to Papua New Guinea using AFLP and simple sequence repeats (SSRs) (Teulat et al., 2000). A dual grouping of Papua New Guinea coconuts was also discernible from this data in concurrence with the predictions of Harries et al. (2004). However, the techniques applied, such as AFLP (Teulat et al., 2000), are poor for resolving introgression events. The current genetic evidence generally favors a Southeast Asian origin for the domesticated form of the coconut, but this picture is likely to become more complex as more phylogenetic-based molecular techniques are applied. The archeobotanical evidence shows that the Lapita peoples exploited coconut (Kirch, 1989), and that it has also been found in the Sepik-Ramu area dated to 5,500–6,000 years BP (Swadling et al., 1991). The early peoples of Near Oceania were certainly exploiting coconut that was present in its wild form as a strand plant. It is a possibility that they domesticated coconut, but there is no direct evidence for it. Breadfruit (Artocarpus altilis). Breadfruit (Artocarpus altilis) is an important source of starch throughout the Pacific. There is a clinal trend west to east across Oceania toward parthenocarpy. The breadfruit of New Guinea is seeded, and may be considered closer to the wild type, while the fruits of Polynesia and Micronesia are sparsely seeded to seedless, and consequently require human intervention for vegetative propagation. At one time it was thought that breadfruit was introduced from Southeast Asia, but the range of the wild type is now accepted to include western Near Oceania. The genetic diversity of breadfruit was found to be highest in New Guinea based on isozyme analysis, supporting its probable origin there (Ragone, 1997). This evidence has recently been strengthened by the identification of A. camansi, an indigenous species of New Guinea, as being the closest wild relative to A. altilis based on internal transcribed spacer (ITS) and trnL-F gene sequence phylogenies (Zerega et al., 2004). The clinal change to the seedless form across the Pacific indicates human selection, the seedless forms having a higher value as a source of starch. The seedless form is the result of triploid plants, but also some of the low-seeded varieties are diploid (Ragone, 2001). The low-seeded varieties are thought to be the result of somatic mutations accumulated through vegetative propagation, and also hybridization between different species. In Micronesia diploid seedless varieties occur resulting from the hybridization of A. altilis and A. mariannenis, a local wild
188
form of Artocarpus growing in the forests of Micronesia. The distribution of AFLP markers between the camansi, mariannensis, and altilis species of Artocarpus demonstrate the complex history of the genus in the Pacific (Zerega et al., 2004). The A. altilis of Melanesia and Polynesia have mostly A. camensis markers, whereas the A. altilis of Micronesia have a mixture of A. camensis and A. mariannenis markers, and a few Micronesian A. altilis have only A. mariannenis markers. The occurrence of some A. mariannenis markers in eastern Melanesia (Solomons, Vanuatu, and Fiji) but not in New Guinea, is taken as evidence of reciprocal voyages between this region and Micronesia (Zerega et al., 2004). Breadfruit is another example of a crop that was domesticated in Near Oceania, but there appears to be little direct evidence from the archeobotanical record that it was exploited before the Lapita peoples arrived, and the evidence for Lapita use is linguistic rather than archeobotanical. The early peoples of Near Oceania may have either exploited A. camensis, or indeed produced from it the domesticated A. atilis. Possibly starch grain or phytolith analysis may provide some answers in the future. Screw Pine (Pandanus sp.). The screw pine (Pandanus sp.) is similar to coconut in that it appears it to have been domesticated on numerous occasions largely on the basis of archeobotanical evidence. Yen (1995) notes that wild Pandanus species occur in a wide range of habitats, highland and lowland, with domesticate forms around most human settlements in the Pacific. Consequently, there is no obvious domestication center. In New Guinea three primary species of interest occur, P. antaresensis, P. brosimos, and P. julianettii. There are also numerous other species of Pandanus in Polynesia, such as P. tectorius, P. spurious, and P. whitmeanus. Some of the Polynesian species are indigenous, and some were probably introduced by the Austronesian settlers from Near Oceania or Southeast Asia. Of the New Guinea species, P. julianettii is the domesticated form, although all three produce large subglobose heads of fruit. It is thought that P. julianettii was domesticated from P. brosimos. P. antaresensis has a tougher fruit and so is less exploited today, but it may have been more important in the past due to its aseasonal fruit production (Harberle, 1995). The archeobotanical record indicates that Pandanus brosimos was domesticated early. Haberle (1998) argues that the domestication process may have been ongoing at 8500 years BP based on pollen data showing P. brosimos to be present in lower altitudes than expected indicating cultivation. Nut remains from 10,000-year-old contexts have been tentatively identified as the domesticated species P. julianettii, indicating Pleistocene cultivation (Yen, 1995). At the Kuk Swamp site, pollen evidence of both P. brosimos and P. antaresensis has been found at pre-Phase 1
Origins of Plant Exploitation in Near Oceania: A Review
levels (10,220–9,910 years BP). However, this in itself proves little, since one would expect these trees to have been naturally growing in that area at that time. However, it has been suggested that the dominance of the Pandanus pollen in the swamp from 10,000–7,400 years BP is indicative of the higher crop stands associated with cultivation (Denham et al., 2004: figures 5 and 6). Other Tree Species. Contemporary arboriculture of Near Oceania has many additional component species (Yen, 1974). Some of these are widely believed to be indigenous to Near Oceania on biogeographical evidence. The sea almond (Terminalia catappa) falls into this category (Yen, 1995). T. kaernbachii and T. copelandii are also cultivated and endemic to New Guinea. Barringtonia procera and B. edule are cultivated trees of New Guinea, probably domesticated in the Solomons (Yen, 1995). The Burckella genus is restricted to a range from eastern Indonesia to Samoa and Tonga, and so probably represents another Near Oceania domestication. The sago palm (Metroxylon sagu) is cited by Yen (1995) as a likely candidate for early domestication in New Guinea. M. sagu is considered a domesticated species, possibly derived from M. rumphii, which has a distribution stretching from New Guinea to the Malay peninsula. In this case, the likelihood of a Near Oceania domestication is increased by the phyletic argument that the genus originated from the Australian continent, and has spread upward toward Indonesia (Dransfield, 1981). Archeobotanical evidence for this genus is lacking prior to the horizons immediately preceding the Lapita contexts (Kirch, 1997), which casts some doubt on its antiquity. Phylogenetic evidence has recently been particularly helpful in the case of Metrosideros (Wright et al., 2000, 2001). Phylogenetic reconstruction based on internal transcribed spacer (ITS) regions of ribosomal DNA show that the Metrosideros species of the Pacific fall into three major clades, with New Zealand at the base. The first clade leads to several endemic species in New Caledonia, and related species in Fiji and also Bonin Island. The second and third clades lead to the M. collina complex, which is shown to have several derived species. Clade 2 again relates to Near Oceania biogeography, including M. collina species in Vanuatu, Samoa, and Fiji. The third clade contains New Zealand species and all the Hawaiian species of Metrosideros. The data support a spread of Metrosideros to Polynesia from New Zealand, probably in the late Tertiary, and a slightly earlier spread to New Caledonia from New Zealand. In contrast to Metrosideros, the Corynocarpus genus shows a center of origin around the Bismarck Archipelago area, based on phylogenetic reconstruction of ITS sequence data (Wagstaff and Dawson, 2000). The expansion, tentatively dated using the molecular clock to
2.5 MYA, followed two independent radiations from this area. The first extended through New Guinea to Australia, terminating in species such as C. cribbianus. The second went through New Caledonia (C. dissimilis) and stretched through to New Zealand (C. laevigatus). Although the karaka (C. laevigatus) has been thought to be an introduction by the Maori, it is likely that it was spread by these people from the North Island to the rest of New Zealand. The Kuk Swamp site yielded wood samples from many species, but of particular interest are Garcinia sp. and Syzgium sp. There are several indigenous species of Garcinia, the genus that boasts the mangosteen in Southeast Asia (Richards, 1990), which produce fragrant fruit. Similarly, some species of Syzygium are indigenous to the Pacific, while some, such as the Malay Apple (Syzygium malaccense syn. Eugenia), were introduced from Southeast Asia.
Primary Herbiculture Banana (Musa sp.). There are two sections of the Musa genus that yield banana crops, Australimusa and the Eumusa. The Australimusa genus domestication involved the development of parthenocarpy, although some of the resulting Fe’i banana cultivars can still produce seeds. Fe’i bananas have a very restricted distribution. Their domestication probably occurred in the Solomon Islands where the endemic form Musa maclayi ssp. maclayi var. erecta occurs today, rather than in New Guinea (De Langhe and De Maret, 1999). The Eumusa section is somewhat more complex. Originally, the Eumusa banana was thought to have originated on the Southeast Asian mainland (Simmonds, 1962). Two species of Eumusa banana are involved, M. acuminata and M. balbisiana, endowed with AA and BB type genomes respectively. The distribution of M. acuminata is from Near Oceania to the Southeast Asian mainland, while M. balbisiana occurs further west, around eastern India and southern China. The two distributions are adjacent around Burma and southern China. Various diploids and triploids of banana exist including the genotypes AA, AAA, AAB, ABB. It is likely the triploids with both the A and B genomes formed in Southeast Asia. The triploids include two frequently confused types, plantain and the “Maia Maoli” bananas of Polynesia, both of which have the AAB genotype. The species of M. acuminata ssp. banksii range from edible seminiferous wild forms through to seedless parthenocarpic varieties, a large number of which occur in New Guinea. The AA genome of M. acuminata ssp. banksii of New Guinea is closer to the donor of the AA genomes to the AAB triploids than is the Asian M. acuminata on the basis of isozyme variation (Lebot et al., 1993) and DNA sequence data (Carreel, 1994). Additionally, the AA genome of the New Guinea M. acuminata ssp. banksii was
189
regional studies and conclusion
found to be identical to the A genome of the AAB triploids of Polynesia in isozyme characters (Lebot et al., 1994). The molecular evidence therefore supports a New Guinea origin for the AA genome in domesticated banana rather than Southeast Asia. The establishment of New Guinea, or perhaps even eastern Indonesia, as the area of origin of the AA edible bananas, leads to the argument that there was a westward movement of the AA types to Southeast Asia, where they hybridized with BB types (De Langhe and De Maret, 1999: figure 20.3). The resulting triploids could then have been transported back across Oceania with the Austronesian-speaking peoples. De Langhe and De Maret (1999) favor this model over an eastward movement of BB types from Southeast Asia because of the paucity of AAB triploids in New Guinea. De Langhe and De Maret (1999) go on to argue that the AAB plantains and AAB “Maia Maoli” plantain-like bananas of Polynesia had origins in different areas of island Southeast Asia to account for their differing distributions. In this model the “Maia Maoli” zone stretches from Borneo across the northern side of Near Oceania to New Britain, from whence it was spread into Polynesia. Plantain formed in southern Indonesia, and spread from there to the Southeast Asian mainland and Indian subcontinent. The evidence from molecular archeobotany supports the hypothesis that bananas were domesticated in New Guinea and the wild progenitors of these domesticates were available to the peoples of Near Oceania in the early Holocene. Recently, evidence has been unearthed that suggests the banana was exploited in the early Holocene in Near Oceania (Denham et al., 2003). Eumusa type phytoliths have been found in the earliest phase of the Kuk Swamp site, dating to 9,000 years BP. The early Musa phytoliths might have been produced in the forest margins around the swamp during Phase 1. However, the very high quantities of phytoliths in the second phase, dating to 6,950–6,440 years BP relative to grass phytoliths indicate active cultivation. Compared to members of the grass family (Poaceae), Musa produces relatively few phytoliths. The equable quantities of phytoliths from grasses and bananas recovered from phase 2 therefore indicate a managed landscape of banana cultivation. Also recently, the first formal identification of Eumusa type phytoliths has been reported from a Lapita site dating to 2,400–1,350 years BP on Watom Island (Lentfer and Green, 2004). Sugar Cane (Saccharum officianarum). Early cytogenetic work identified New Guinea as a world center of origin for Saccharum officianarum (sugar cane), where it was domesticated from S. robustum (Brandes, 1958). The crop is of such global significance that considerable effort has been directed at its genomic characterization (Grivet and
190
Arruda, 2001; Grivet et al., 2004). Various wild species of sugarcane grow in the Pacific including S. spontaneum and S. robustum. The biogeographic range of S. robustum is restricted to western Indonesia and Near Oceania, while S. spontaneum occurs from Africa and the Middle East to tropical Asia to the Pacific islands. Grivet et al. (2004) suggest that S. spontaneum evolved on the Sunda shelf and S. robustum in the Sahul during the Quaternary. S. robustum has a karyotype that matches that of S. officinarum (2n = 80). An origin of S. officinarium from S. robustum is complicated by hybridization with other species. Although S. robustum and S. officinarum have been shown to be closely related to each other on DNA evidence (Besse et al., 1997; D’Hont et al., 1998), S. officinarum includes some markers not found in S. robustum but rather in S. spontaneum and even in the sister genera Erianthus and Miscanthus (Lu et al., 1994). Daniels and Daniels (1993) suggested that a combination of these other species came from Southeast Asia with the Austronesian peoples, and then hybridized with S. officinarium in New Guinea. Molecular evidence suggests that at most there was limited introgression and that the Saccharum genus is highly differentiated to Erianthus (Besse et al., 1997). Hybridization between S. spontaneum and S. robustum or S. officinarium in the New Guinea area was supported by recent work that showed New Guinea accessions of S. spontaneum were closely related enough to S. robustum to hybridize fully and produce fertile offspring of the same ploidy (D’Hont et al., 1998). The sugarcane story is complicated by the fact that the Austronesians had sugarcane 6,000 years BP (Bellwood, 1985). Daniels and Daniels (1993) argue that this was probably another species S. sinense (Asiatic sugar). It is not thought, on the basis of molecular evidence, that Asiatic sugar was involved in the domestication of S. officinarium. Unfortunately, there is no archeobotanical evidence for sugarcane, and its use by Proto-Oceanic peoples is largely inferred from linguistic evidence (Kirch, 1997). This is another crop that was probably available in the wild form to the early peoples of Near Oceania, whether they had a hand in its domestication or not. Other Primary Herbaceous Crops. Another herbaceous crop cited as a cultivar endemic to New Guinea is Cordyline fruticosa (Yen, 1991a). However, there is no direct archeobotanical, linguistic, or molecular archeobotanical evidence. The natural biogeography suggests that it might be so. This plant occurs naturally from the Himalayas to Polynesia, and is considered an important emergency food, along with the sago palm (Yen, 1974).
Secondary Prehistoric Crop Introductions A large number of plants that are indigenous to Southeast Asia were introduced to Near Oceania and the Pacific by
Origins of Plant Exploitation in Near Oceania: A Review
human agency. A small number of these species, discussed in the previous section, may have been introduced by peoples of the early Holocene and late Pleistocene. A large number of these species are attributed to the later Austronesian human expansion. This attribution comes partly from archeobotanical evidence and partly from linguistic evidence. However, as has been seen in the example of kava discussed above, the attribution of a Proto-Oceanic linguistic root for a plant species does not necessarily equate with its origin with those people, an implicit assumption that terms will be borrowed from local languages if they exist already is not a secure one. The roster of plants exploited by the peoples of the Austronesian expansion includes all those in the previous section. Additionally, the following species are attributed to imports of the Austronesian culture. The tree species include the Malay Apple (Syzygium malaccense syn. Eugenia sp.), Tahitian chestnut (Inocarpus fagiferus), and Wi Apple (Spondias dulcis), all of which are of Indo-Malaysian origin. Archeological remains of the last two have been found at Mussau (Kirch, 1989). Indian Mulberry (Morinda citrifolia) has a wide distribution ranging from India to Hawaii to tropical America, and has an effective sea dispersal mechanism (Morton, 1992). It is possible that this tree, which is attributed to the Lapita complex on the basis of linguistic evidence, may have reached Near Oceania under its own dispersal mechanism. Abelmoschus manihot is attributed to the Austronesian expansion on linguistic evidence as well (Kirch, 1989). The Austronesians are likely to have brought various crops that do not preserve well in the archeological record, such as root crops. For instance, tumeric (Curcuma longa) is associated with the Austronesians through linguistics. Lynch (2002) provides evidence for a ProtoOceanic term for wild ginger (Zinger zerumbet), which is thought to be naturalized in the Pacific islands. However, as discussed above, this species has been attributed to the early phases of Kuk Swamp (Denham et al., 2004), so is currently of less certain status. A range of species that may ultimately have origins in Southeast Asia have been shown in the archeobotanical record to have arrived earlier than the Austronesianspeaking peoples. The Sepik-Ramu site on the northern side of Papua New Guinea dates to 5,800 years BP in the mid Holocene (Swadling, 1997). This site has yielded archeobotanical evidence of Aleurites moluccana (candle nut), Areca catechu (betel nut), Cordia sp., Pangium sp., Parinarium sp., and Sterculia sp.. Aleurites moluccana is indigenous to Indo-Malaysia while Cordia, Sterculia sp., and Areca catechu have distributions that include Near Oceania. The peoples of the early Holocene may have introduced these species from Southeast Asia. Finally, there is a rather large category of species that are not yet specifically claimed by any particular
anthropological framework, but nevertheless are species of importance for exploitation, and derive some time in the past from Southeast Asia. These species may have been brought by the early peoples of the Holocene, or the later Austronesians, or perhaps spread with the aid of natural dispersal mechanisms. Tree species include Broussenetia papyrifera, Hibiscus rosa-sinensis, and Bischofia javanica (Whistler, 1991). In the case of the paper mulberry (B. papyrifera), for instance, the tree is indigenous to China and Japan but in Polynesia it does not seed and is vegetatively propagated. There are various grass species that are some cause of interest. One example is Job’s tears (Coix lacryma-jobi), a panicoid grass of Asia closely related to sorghum, and also curiously closely related to maize (Takahashi et al., 1999). Whistler (1991) attributes Coix to an aboriginal introduction. Wild rice O. rufipogen is found both in swampy habitats in New Guinea and northern Australia (Yen, 1995). Although not used in New Guinea, the Aboriginal populations of Australia do forage for grains. If wild rice spread to Australia from Southeast Asia naturally, then it is likely to have been at a time when there was no Sahul savannah. An aboriginal spread of the species makes an interesting and contentious alternative hypothesis. Curiously, the Austronesian languages have terms for rice and rice cultivation indicating some connection to the crop in the past (Blust, 1985). It has been argued that Oryza sativa ssp. indica brought as part of the Austronesian crop roster may not have been suited to growing conditions in the non-seasonal humid tropics (Bellwood, 1985). Recent genomic evidence has supported a dual origin for rice, placing a common ancestry between the indica and japonica cultivars at 440,000 years BP (Ma and Bennetzen, 2004). The wild species O. nivara and O. rufipogen are progenitors to O. indica and O. japonica respectively based on recent phylogenetic evidence (Cheng et al., 2003). Bellwood’s assertion may have been directed at the wrong branch of the rice genus, and molecular archeobotanical research of wild rice in this area may prove useful.
Tertiary Prehistoric Crop Introductions There is a small group of crops that are derived not from Southeast Asia or Australia but from the Americas. There are some strand species that have managed to spread from the American continents into the Polynesian islands. One such case is cotton (Gossypium tomentosum) that successfully became established in Hawaii, probably by floating across from America in the Pleistocene (Fryxell, 1979). The Austronesian diaspora finally made contact with America at some point in the last two millennia before European contact, also resulting in the introduction of a small number of New World crops into Oceania of special importance.
191
regional studies and conclusion
The Tripartite Hypothesis Yen (1973) put forward a tripartite origin hypothesis to explain the cultivar diversity in morphology in sweet potato (Ipomoea batatas) in Oceania. In the first origin, sweet potato was introduced into Polynesia from South America, most likely Peru, somewhere between 1,600 and 1,300 years BP. The second and third origins he suggested to be due to European contact during the sixteenth century. Recent archeological evidence suggests that East Polynesia was colonized no earlier than 1,000 years BP (Anderson et al., 2002), so Yen’s original estimates should be updated accordingly. In the second origin the Spanish introduced the sweet potato to the Philippines fairly directly from Mexico and Peru on a prototype of the Manilla Galleon. The third origin was via Portuguese shipping by Africa and India to Indonesia. The hypothesis predicted three sweet potato lines, “kumara,” “camote,” and “batata” for the three origins respectively. The genetic predictions of the hypothesis include a distinction between crops of the eastern and western Pacific. Also, the eastern crops should have undergone a population bottleneck caused by the voyage from the New World to Polynesia in comparatively small outriggers. If Yen were correct, the western Pacific varieties should be closer in phylogeny to those of Mesoamerica while the eastern Pacific varieties should be closer to those of Peru. This hypothesis also raised the possibility that other crops were also brought to Polynesia from the New World, such as bottle gourd (Lagenaria siceraria) and the Polynesian tomato (Solanum repandum). The latter plant is now increasingly rare in Polynesia (Whistler, 1991), so it is too late to study it effectively.
Sweet Potato (Ipomoea batatas) The tripartite hypothesis predicts the presence of sweet potato in prehistoric eastern Oceania. Since tuberous plant material does not survive well in the archeobotanical record, it is surprising that any prehistoric evidence has been unearthed at all. The first sweet potato remains were found in a charred deposit in Hawaii (Rosendahl and Yen, 1971), radiocarbon dated to AD 1425–1605, which meant they could not unequivocally be attributed to a prehistoric origin. Firm evidence for prehistoric Polynesian sweet potatoes was found on Mangaia Island, using parenchymous tissue. These samples have a maximum possible age of 1,000 years BP, clearly in the prehistoric (Hather and Kirch, 1991). More recently, prehistoric remains of sweet potato have been detected through starch grain analysis from a site at Pouerua, New Zealand (Horrocks et al., 2004). In this case the remains are dated to 467 years BP. The evidence from molecular archeobotany is equivocal in regard
192
to Yen’s tripartite hypothesis. The New Zealand varieties fall into two groups based on RAPD data, corresponding to nineteenth century and “ancient” (pre-European) varieties, implying two separate antipodean introductions as would be expected if the tripartite hypothesis were correct (Harvey et al., 1997). The genetic diversity of sweet potato is higher in Mesoamerica than Peru on the basis of both AFLP and microsatellite data, possibly indicating a Mesoamerican origin of the species (Zhang et al., 2000ab). These authors also suggested that since the genetic diversity of Oceanic varieties is also high, they were derived from Mesoamerica rather than Peru, as hypothesized by Yen (Zhang et al., 2004). However, the Mesoamerican and Peruvian AFLP character combinations are discernible, and the Oceanic sweet potato lines have a diversity that spans both these regions (Zhang et al., 2004: figure 1). Therefore, the Oceanic diversity may reflect a movement of crops across Oceania west to east in historic times. Since the AFLP-based approach is of limited use when dealing with a model of introgressing populations, phylogenetic data would be more useful in this instance. Another hypothesis of sweet potato movement in the Pacific was suggested on the basis of linguistic evidence (Rensch, 1991). Rensch forwarded a dual pre-Columbian route of sweet potato into Polynesia in which the linguistic terms *kuara/*kuala are derived from a northern route via Hawaii, and the term *kumara is derived along a southern route via Easter Island. The linguistic basis of this argument is contested (Green, 2005), but opens up the possibility of multiple voyages between Polynesia and New World. Alternatively, the *kuumala may be the proto-form of the Polynesian for sweet potato from which all other terms are derived (Green 2005). The term *kuumala originates in eastern Polynesia, where sweet potato first arrived. Under Green’s hypothesis the pre-Columbian sweet potato input into Oceania was by Polynesians originating from Mangarevan-Pitcairn-Rapa Nui area into an eastern Polynesian region spanning Mangaia to the Marquesas Islands. The crop was then introduced into Rapa Nui in the twelfth and thirteenth centuries indirectly from the Marquesas via the Tuamotu Archipelago. Movements out of the central eastern region of Polynesia can also variously explain the distribution of sweet potato in the Oceanic peripheries such as Hawaii and New Zealand. Phylogenetic data may prove useful in investigating these hypotheses. The speciation events that gave rise to I. batatas and its wild relatives are too recent to obtain phylogenetic resolution using ITS sequences of the different species. The species are so similar that both AFLP and RAPD have been used as phylogenetic characters between species (Dhillon and Ishiki, 1999; Huang et al., 2002), which is problematic because of the very fast mutation rates involved.
Origins of Plant Exploitation in Near Oceania: A Review
Both approaches establish I. trifida as the closest living relative to I. batatas. Alternatively, microsatellite diversity could provide phylogenetic quality data to track the evolutionary history and relationships of sweet potato in the Pacific (Hu et al., 2003). In summary, elements of Yen’s original tripartite hypothesis have been shown to be true. It has been established with certainty that sweet potato has a dual origin in the Pacific, firstly from prehistoric imports from the New World and secondly from European contact. However, further elucidation of the evolutionary history of sweet potato in the Pacific will require phylogenetic analyses to establish the likely area of origin in the New World, the existence of Yen’s camote and batata lines, and to test the linguistic hypotheses of Rensch and Green respectively.
Bottle Gourd (Lagenaria siceraria) Bottle gourd (Lagenaria siceraria) is another crop that originally was thought to be introduced to Polynesia from the Indo-Malaysian region (Heiser, 1979). However, there is now a reasonable amount of linguistic, archeobotanical, and phylogeographical evidence to argue that bottle gourd also has a dual origin from both sides of the Pacific (Green, 2000). The bottle gourd has a long history in Southeast Asia, having been present there for as long as 7,000 years (Chang, 1986), and its long history in island Southeast Asia is supported by linguistics (Green, 2000). The gourd was probably not brought to Oceania during the Austronesian expansion since it cannot be successfully identified in Proto-Oceanic (Ross, 1996). It appears to have been introduced somewhat later, since the oldest Oceanic sample occurring in the New Guinea Highlands is about 2,000 years BP (Golson et al., 1967). The archeobotany of the bottle gourd has been subject to some confusion because of its similarity to wax gourd (Benincasa hispida), another gourd commonly found in the Pacific. Wax gourd has now been identified at the Kana site in the New Guinea Highlands dating to 2,450 years BP, which casts some doubt on the original oldest identification of bottle gourds in the New Guinea Highlands (Matthews, 2003b; Muke and Nadui, 2003). Significantly, samples originally identified as bottle gourd in western Polynesia and Fiji were re-identified as wax gourd (Whistler, 1990). This re-identification established that the distribution of bottle gourd is disjunct across the Pacific, with a gap occurring in western Polynesia in which there has been no history of the gourd. The bottle gourd gap supports a dual input of gourd into the Pacific, with a New World origin for bottle gourd in eastern Polynesia. It is probable that the bottle gourd would have been introduced from the New World with the sweet potato (Whistler, 1991). As with sweet potato, a New World origin of bottle gourd in eastern Polynesia
predicts the occurrence of prehistoric samples. Archeobotanical evidence for prehistoric bottle gourd is still limited, but pollen grains have been identified from New Zealand dated to 467 years BP (Horrocks et al., 2004). Recent molecular evidence has also supported a dual origin for the gourd (Clarke et al., 2005). This study found that gourd samples from New Zealand have alleles that are associated with the New World as well as ones associated with Southeast Asia. Finally, linguistic support has also been found for a dual origin. Green (1998) argues that the bottle gourd is associated with the earliest cultures that arrived on Easter Island (Rapa Nui) in the twelfth and thirteenth centuries, where the term *fue is applied specifically to bottle gourd rather than to a more general use in western Oceania. In summary, it is likely on the basis of archeobotany, molecular archeobotany, and also linguistics that the bottle gourd in eastern Oceania originated from the New World while that in the west came from Southeast Asia.
A New Paradigm for Pacific Crop Origins New Guinea was recently described as a “Cradle of Agriculture,” echoing Sauer’s (1952) description of Southeast Asia (Neumann, 2003). The accumulating objective evidence has certainly countered the supposition that subsistence techniques in Near and Remote Oceania were wholly imported from Southeast Asia. This is in part because the range of techniques available to archeobotany and molecular archeobotany has increased significantly in the past decade, and there has been a concurrent decrease in dependency upon anthropological frameworks for the interpretation of crop origins. Denham (2004a) makes the point that to view the plant exploitation origins through an artificial dichotomy of pre- and post-Austronesian expansion may be distorting the truth of the matter, when the reality may be closer to a continuum of indigenous domestications and introductions over time. Of course, as Green (2000) points out, one cannot separate the human processes from the equation. This review comes at a time when the botanical and anthropological sources of evidence have become more balanced leading to a new view of crop origins in Near Oceania. The evidence has shown that many species previously thought imported from Southeast Asia were actually domesticated within the New Guinea/Near Oceania area. Some of these Near Oceania crops may have been domesticated in the early Holocene while others may have been domesticated later. The form of this evidence has frequently been through the identification of genetic diversity using anonymous genomic markers. The power of such approaches is adequate for identifying centers of diversity, but less useful in providing timescales of events
193
regional studies and conclusion
and a reconstruction of the evolutionary steps that gave rise to the observed diversity. More phylogenetic-based approaches will be useful in these regards. In particular an interspecific phylogeny of the Dioscorea and Canarium genera would be helpful in establishing where the great yam came from and the nature of the identity of the basal Canarium species from which the other species expanded. Similar approaches to intraspecific phylogeny would also be useful in the cases of Metroxylon and Musa to identify the geographical regions that have basal or derived plants to test the emergent ideas about the movements of these crops. Microsatellites could prove useful as tools for calibrating the amount of time that has passed since the founding domestication events of plant species in an area, and track how plants have moved from island to island in the Pacific through the identification of apomorphic characters. Such microsatellite loci are now available for the recalcitrant coconut (Rivera et al., 1999, Perera et al., 2000), as well as quantitative trait loci (Heran et al., 2000). This kind of objective botanical evidence may help establish whether indigenous species domesticated within Near Oceania were domesticated simultaneously or in a staggered fashion. Similarly, it may help establish whether introduced species came over together from Southeast Asia as a coherent unit, or in a staggered fashion in the late Pleistocene and early Holocene.
References Allaby RG, Brown TA. 2003. AFLP analysis and the origins of agriculture. Genome 46: 448–53. Allen J, Gosden C, White JP. 1989. Human pleistocene adaptations in the tropical island Pacific: Recent evidence from New Ireland, a Greater Australian outlier. Antiquity 63: 548–61. Anderson AJ, Sinoto YH. 2002. New radiocarbon ages of colonization sites in East Polynesia. Asian Perspectives 41: 242–57. Bellwood P. 1985. Prehistory of the Indo-Malaysian archipelago. Sydney, Australia: Academic Press. Bellwood P. 1996. The origins and spread of agriculture in the Indo-Pacific region: Gradualism and diffusion or revolution and colonization? In: Harris DR, editor. The origins and spread of agriculture and pastoralism in Eurasia. London: UCL Press. pp 465–98. Besse P, McIntyre C, Berding N. 1997. Characterization of Erianthus sect. Ripidium and Saccharaum germplasm Andropogoneae Saccharinae using RFLP markers. Euphytica 93: 283–92. Blust R. 1985. The Austronesian homeland: A linguistic perspective. Asian Perspectives 26: 45–67. Brandes EW. 1958. Origin, classification and characteristics. In: Artshwager E, Brandes EW, editors. Sugarcane (Saccharum officanurum L.). U.S.D.A. Handbook 122. pp 1–35.
194
Carreel F. 1994. Etude de la diversité génétique des bananiers genre Musaàl’aide des marquers RFLP. Thèse, Institut National Agronomique, Paris-Grignon. Chang KC. 1986. The archaeology of ancient China. New Haven, CT: Yale University Press. Cheng C, Motohashi R, Tchuchimoto S, Fukuta Y, Ohtsubo H, Ohtsubo E. 2003. Polyphyletic origin of cultivated rice: Based on the interspersion patterns of SINEs. Molecular Biology and Evolution 20: 67–75. Choi IY, Lee JK, Goo HJ, Park JH, Kim NS, Park CH, Chang KJ. 2002. Genetic diversity and relationships among Dioscorea alata L., and related Dioscorea species revealed by AFLP analysis. Korean Journal of Genetics 24: 305–12. Clarke A, Burtenshaw MK, McLenachan PA, Erickson D, Penny D. 2006 Reconstructing the origins and dispersal of the Polynesian bottle gourd (Lagenaria siceraria). Molecular Biology and Evolution 23: 893–900. Coates DJ, Yen DE, Gaffrey PM. 1988. Chromosomal variation in taro, Colocasia esculenta: Implications for origin in the Pacific. Cytologia 55: 551–60. Coursey DG. 1976. Yams. In: Simmonds NW, editor. Evolution of crop plants. London: Longman. pp 70–4. Daniels J, Daniels C. 1993. Sugarcane in prehistory. Archaeology in Oceania 28: 1–7. De Langhe E, De Maret P. 1999. Tracking the banana: Its significance in early agriculture. In: Gosden C, Hather J, editors. The prehistory of food: Appetites for change. London: Routledge. pp 377–96. Denham T. 2004a. The roots of agriculture and arboriculture in New Guinea: Looking beyond the Austronesian expansion, Neolithic packages and indigenous origins. World Archaeology 36: 610–20. Denham T. 2004b. Early agriculture in the Highlands of New Guinea: An assessment of Phase 1 at Kuk Swamp. Records of the Australian Museum, Supplement 29: 47–57. Denham T, Haberle S, Lentfer C. 2004. New evidence and revised interpretations of early agriculture in Highland New Guinea. Antiquity 78: 839–57. Denham T, Haberle SG, Lentfer C, Fullager R, Field J, Therin M, Porch N, Winsborough B. 2003. Origins of agriculture at Kuk Swamp in the Highlands of New Guinea. Science 301: 189–93. Dhillon NPS, Ishiki K. 1999. Genomic variation and genetic relationships in Ipomoea spp. Plant Breeding 118: 161–5. D’Hont A, Ison D, Alix K, Roux C, Glaszmann JC. 1998. Determination of basic chromosome numbers in the genus Saccharum by physical mapping of ribosomal RNA genes. Genome 41: 221–5. Dransfield J. 1981. Palms and Wallace’s line. In: Whitmore TC, editor: Wallace’s line and plate tectonics. Oxford, UK: Clarendon Press. pp 43–56. Frederickson C, Spriggs M, Ambrose W. 1993. Pamwak rockshelter: A Pleistocene site on Manus Island, Papua New Guinea. In: Smith MA, Spriggs M, Fankhauser B, editors. Sahul in review: Pleistocene archaeology in Australia, New Guinea and Island Melanesia. Occasional papers in prehistory, No. 24. Canberra, Australia: Department of Prehistory, Australian National University. pp 144–54.
Origins of Plant Exploitation in Near Oceania: A Review
Fryxell PA.1979. The natural history of the cotton tribe. College Station and London: Texas A&M University Press. Golson J. 1977. No room at the top: Agricultural intensification in the New Guinea Highlands. In: Allen J, Golson J, Jones R, editors. Sunda and Sahul: Prehistoric studies in southeast Asia, Melanesia and Australia. London: Academic Press. pp 601–38. Golson J. 1991. The New Guinea Highlands on the eve of agriculture. Indo-Pacific Prehistory Association Bulletin 11: 82–91. Golson J, Lampert RJ, Wheeler JM, Ambrose WR. 1967. A note on carbon dates for horticulture in the New Guinea Highlands. Journal of the Polynesian Society 76: 369–71. Gosden C. 1992. Production systems and the colonization of the Western Pacific. World Archaeology 24: 55–69. Gosden C. 1995. Aboriculture and agriculture in coastal Papua New Guinea. Antiquity 265: 807–17. Green RC. 1998. Rapanui origins prior to European contact— the view from Eastern Polynesia. In: Vargas Casanova P, editor. Easter Island and East Polynesian prehistory. Santiago, Chile: University of Chile. pp 87–110. Green RC. 2000. A range of disciplines support a dual origin for the bottle gourd in the Pacific. Journal of the Polynesian Society 109: 191–7. Green RC. 2005. Sweet potato transfers in Polynesian prehistory. In: Ballard C, Brown P, Bourke M, Harwood T, editors. The sweet potato in Oceania: A reappraisal. Sydney, Australia: University of Sydney. pp 43–62. Grivet L, Arruda P. 2001. Sugarcane genomics: Depicting the complex genome of an important tropical crop. Current Opinions in Plant Biology 5: 122–7. Grivet L, Daniels C, Glaszmann JC, D’Hont A. 2004. A review of recent molecular genetics evidence for sugarcane evolution and domestication. Ethnobotany Research and Applications 2: 9–17. Haberle S. 1994. Anthropogenic indicators in pollen diagrams: Problems and prospects for late Quaternary palynology in New Guinea. In: Hather JG, editor. Tropical archaeobotany: Applications and new developments. London: Routledge. pp 172–201. Haberle S. 1995. Identification of cultivated Pandanus and Colocasia in pollen records and the implications for study of early agriculture in New Guinea. Vegetation History and Archaeobotany 4: 195–210. Haberle SG. 1998. Late quaternary vegetational change in the Tari Basin, Papua New Guinea. Palaeogeography, Palaeoclimatology and Palaeoecology 137: 1–24. Harries H, Baudouin L, Cardeña R. 2004. Floating, boating and introgression: Molecular techniques and the ancestry of coconut palm populations on Pacific Islands. Ethnobotany Research and Applications 2: 37–53. Harries HC. 1978. The evolution dissemination and classification of Cocos nucifera. Botanical Review 44: 265–320. Harris DR. 1969. Agricultural systems, ecosystems and the origins of agriculture. In: Ucko PJ, Dimbleby GW,
editors. The domestication and exploitation of plants and animals. London: Duckworth. pp 3–16. Harvey CF, Gill GP, Crossman C, Fraser LG. 1997. Assessing relationships of kumara cultivars by RAPD analysis. New Zealand Journal of Botany 35: 479–85. Haslam M. 2004. The decomposition of starch grains in soils: Implications for archaeological residue analysis. Journal of Archaeological Science 31: 1715–34. Hather JG. 1991. The identification of charred archaeological remains of vegetative parenchymous tissue. Journal of Archaeological Science 18: 661–75. Hather JG. 1994a. The identification of charred root and tuber crops from archaeological sites in the Pacific. In: Hather JG, editor. Tropical archaeobotany: Applications and new developments. London: Routledge. pp 51–64. Hather JG. 1994b. A morphological classification of roots and tubers and its bearing on the origins of agriculture in Southwest Asia and Europe. Journal of Archaeological Science 21: 719–24. Hather JG. 1996. The origins of tropical vegeculture: Zingiberaceae, Araceae and Dioscoraceae in Southeast Asia. In: Harris DR, editor. The origins and spread of agriculture and pastoralism in Eurasia. London: UCL Press. pp 538–50. Hather J, Kirch PV. 1991. Prehistoric sweet potato (Ipomoea batatas) from Mangaia Island, Central Polynesia. Antiquity 65: 887–93. Hawkes JG. 1969. The ecological background of plant domestication. In: Ucko PJ, Dimbleby GW, editors: The domestication and exploitation of plants and animals. London: Duckworth. pp 17–30. Hay AR, Wise R. 1991. The genus Alocasia (Araceae) in Australasia. Blumea 35: 499–545. Heads M. 2001. Regional patterns of biodiversity in New Guinea plants. Botanical Journal of the Linnean Society 136: 67–73. Heiser CB. 1979. The gourd book. Norman, OK: University of Oklahomoa Press. Herran A, Estioko L, Becker D, Rodriguez MJ, Rohde W, Ritter E. 2000. Linkage mapping and QTL analysis in coconut (Cocos nucifera L.). Theoretical and Applied Genetics 101: 292–300. Hope G, Golson J. 1995. Late Quaternary change in the mountains of New Guinea. Antiquity 69: 818–30. Horrocks M, Irwin G, Jones M, Sutton D. 2004. Starch grains and xylem cells of sweet potato (Ipomoea batatas) and bracken (Pteridium esculentum) in archaeological deposits from northern New Zealand. Journal of Archaeological Science 31: 251–8. Hu J, Nakatani M, Lalusin AG, Kuranouchi T, Fujimura T. 2003. Genetic analysis of sweet potato and wild relatives using inter-simple sequence repeats (ISSRs). Plant Breeding 53: 297–304. Huang J, Corke H, Sun M. 2002. Highly polymorphic AFLP markers as a complementary tool to ITS sequences in assessing genetic diversity and phylogenetic relationships of sweet potato (Ipomeoa batatas (L.) Lam.) and its wild relatives. Genetic Resources and Crop Evolution 49: 541–50.
195
regional studies and conclusion
Irwin SV, Kaufusi P, Banks K, de la Peña, R, Cho JJ. 1998. Molecular characterization of taro (Colocasia esculenta) using RAPD markers. Euphytica 99: 183–9. Isshiki S, Otsuka KI, Tahiro Y, Miyazaki S. 1999. A probable origin of triploids in taro [Colocasia esculenta (L.) Schott]. Journal of the Japanese Society of Horticulture Science 68: 774–9. Jones R, Bowler J. 1980. Struggle for the savannah: Northern Australia in ecological and prehistoric perspective. In: Jones R, editor. Northern Australia: Options and implications. Canberra, Australia: Australian National University. pp 3–31. Judd WS, Campbell CS, Kellogg EA, Stevens PF. 1999. Plant systematics: A phylogenetic approach. Sunderland, MA: Sinauer Associates Inc. Kajale MD. 1989. Mesolithic exploitation of wild plants in Sri Lanka: Archaeological study of the cave site of BeliLena. In: Harris DR, Hillman G, editors. Foraging and farming: The evolution of plant domestication. London: Unwin Hyman. pp 269–81. Kirch PV. 1997. The Lapita peoples: Ancestors of the Oceanic World. Oxford, UK: Blackwell. Kirch PV. 1989. Second millennium B.C. arboriculture in Melanesia: Archaeological evidence from the Mussau Islands. Economic Botany 43: 225–40. Kreike CM, van Eck HJ, Lebot V. 2004. Genetic diversity of taro, Colocasia esculenta (L.) Schott, in Southeast Asia and the Pacific. Theoretical and Applied Genetics 109: 761–8. Lebot V. 1991. Kava (Piper methysticum Forst. F.): The Polynesian dispersal of an Oceanic plant. In: Cox PA, Banack SA, editors. Islands, plants and polynesians. Portland, OR: Dioscorea Press. pp 169–202. Lebot V. 1999. Biomolecular evidence for plant domestication in the Sahul. Genetic Resources and Crop Evolution 46: 619–28. Lebot V, Aradhya KM. 1991. Isozyme variation in taro (Colocasia esculenta (L.) Schott) from Asia and Oceania. Euphytica 56: 55–66. Lebot V, Aradhya KM, Manshardt RM. 1991b. Geographical survey of genetic variation in kava (Piper methysticum). Pacific Science 45: 169–85. Lebot V, Aradhya KM, Manshardt R, Meilluer B. 1993. Genetic relationships among cultivated bananas and plantains from Asia and the Pacific. Euphytica 67: 163–75. Lebot V, Johnston E, Zheng QY, McKern D, McKenna DJ. 1999. Morphological, phytochemical, and genetic variation in Hawaiian cultivars of ‘awa (kava, Piper methysticum, Piperaceae). Economic Botany 53: 407–18. Lebot V, Lévesque J. 1989. The origin and distribution of kava (Piper methysticum Forst. F. Piperaceae): A phytochemical approach. Allertonia 5: 223–80. Lebot V, Meilleur BA, Manshardt RM. 1994. Genetic diversity in eastern Polynesian Eumusa Bananas. Pacific Science 48: 16–31. Lebot V, Merlin M, Lindstrom L. 1992. Kava: The Pacific drug. New Haven, CT: Yale University Press. Lebot V, Prana MS, Kreike N, van Heck H, Pardales J, Okpul T, Gendua T, Thongjiem M, Hue H, Viet N,
196
Yap TC. 2004. Characterisation of taro (Colocasia esculenta (L.) Schott) genetic resources in Southeast Asia and Oceania. Genetic Resources and Crop Evolution 51: 381–92. Lebot V, Siméoni P. 2004. Is the quality of Kava (Piper methysticum Forst. f.) responsible for different geographical patterns? Ethnobotany Research and Applications 2: 19–28. Lebrun P, N’Cho YP, Seguin L, Grivet, L, Baudouin L. 1998. Genetic diversity in coconut (Cocos nucifera L.) revealed by restriction fragment length polymorphism (RFLP) markers. Euphytica 101: 103–8. Lentfer CJ, Green RC. 2004. Phytoliths and the evidence for banana cultivation at the Lapita Reber-Rakival site on Watom Island, Papua New Guinea. Records of the Australian Museum Supplement 29: 75–88. Lentfer C, Therin M, Torrence R. 2002. Starch grains and environmental reconstruction: A modern test case from West New Britain, Papua New Guinea. Journal of Archaeological Science 29: 687–98. Lepofksy D. 1992. Aboriculture in the Mussau Islands, Bismark archipelago. Economic Botany 46: 192–211. Loy T, Spriggs M, Wickler S. 1992. Direct evidence for human use of plants 28,000 years ago: Starch residues on stone artifacts from northern Solomon Islands. Antiquity 66: 898–912. Lu YH, D’Hont A, Waler DIT, Rao PS, Feldmann P, Glaszmann JC. 1994. Relationships among ancestral species of sugarcane revealed with RFLP using single copy maize nuclear probes. Euphytica 78: 7–18. Lynch J. 2002. Potent roots and the origin of kava. Oceanic Linguistics 41: 493–513. Ma J, Bennetzen JL. 2004. Rapid and recent growth and divergence of rice nuclear genomes. Proceedings of the National Academy of Sciences USA 101: 12404–10. Maloney BK. 1996. Canarium in the Southeast Asian and Oceanic archaeobotanical and pollen records. Antiquity 70: 926–33. Matsuda M, Nawata E. 2002. Geographical distribution of ribosomal DNA variation in taro, Colocasia esculenta (L.) Schott, in eastern Asia. Euphytica 128: 165–72. Matthews PJ. 1990. The origins, dispersal and domestication of taro. PhD thesis. Canberra, Australia: The Australian National University. Matthews PJ. 2003a. Taro planthoppers (Tarophagus spp.) in Australia and the origins of taro (Colocasia esculenta) in Oceania. Archaeology in Oceania 38: 192–202. Matthews PJ. 2003b. Identification of Benincasa hispida (wax gourd) from the Kana archaeological site, Western Highlands Province, Papua New Guinea. Archaeology in Oceania 38: 186–91. Matthews PJ. 2004. Genetic diversity in taro and the preservation of culinary knowledge. Ethnobotany Research and Applications 2: 55–71. Matthews PJ, Gosden C. 1997. Plant remains from waterlogged sites in the Arawe Islands, West New Britain province, Papua New Guinea: Implications for the history of plant use and domestication. Economic Botany 51: 121–33.
Origins of Plant Exploitation in Near Oceania: A Review
Matthews PJ, Terauchi R. 1994. The genetics of agriculture: DNA variation in taro and yam. In: Hather JG, editor. Tropical archaeobotany: Applications and new developments. London: Routledge. pp 251–62. Morton JF. 1992. The ocean-going Noni, or Indian Mulberry (Morinda citrifolia, Rubiaceae) and some of its ‘colourful’ relatives. Economic Botany 46: 241–56. Muke J, Mandui H. 2003. In the shadows of Kuk: Evidence for prehistoric agriculture at Kana, Waghi Valley, Papua New Guinea. Archaeology in Oceania 38: 177–85. Neumann K. 2003. New Guinea: A cradle of agriculture. Science 301: 180–81. Palmer JD, Herbon LJ. 1988. Plant mitochondrial DNA evolves rapidly in structure, but slowly in sequence. Journal of Molecular Evolution 28: 87–97. Parr JF, Carter M. 2003. Phytolith and starch grain analysis of sediment samples from two archaeological sites on Dauar Island, Torres Strait, northeastern Australia. Vegetation History and Archaeobotany 12: 131–41. Pawley A, Green R. 1973. Dating the dispersal of the Oceanic languages. Oceanic Linguistics 12: 1–67. Perera L, Russell JR, Provan J, Powell W. 2000. Use of microsatellite DNA markers to investigate the level of genetic diversity and population genetic structure of coconut (Cocos nucifera L.). Genome 43: 15–21. Piperno DR. 1991. The status of phytolith analysis in the American tropics. Journal of World Prehistory 5: 155-91. Piperno DR, Ranere AJ, Holst I, Hansell P. 2000. Starch grains reveal early root crop horticulture in the Panamanian tropical forest. Nature 407: 894–97. Ragone D. 1997. Breadfruit, Artocarpus altilis (Parkinson) Fosberg. Promoting the conservation and use of underutilized and neglected crops series vol. 10. Rome, Italy: International Plant Genetic Resources Institute. Ramser J, Weising K, Lopez-Peralta C, Terhalle W, Terauchi R, Kahl G. 1997. Molecular marker based taxonomy and phylogeny of Guinea yam (Dioscorea rotundata– D. cayenensis). Genome 40: 903–15. Ragone D. 2001. Chromosome numbers and pollen stainability of three species of breadfruit (Artocarpus, Moraceae). American Journal of Botany 88: 693–6. Renner SS, Zhang LB. 2004. Biogeography of the Pistia Clade (Araceae): Based on chloroplast and mitochondrial DNA sequences and Bayesian divergence time inference. Systematic Botany 53: 422–32. Rensch KH. 1991. Polynesian plant names: Linguistic analysis and ethnobotany, expectations and limitations. In: Cox PA, Banack SA, editors. Islands, plants and polynesians. Portland, OR: Dioscorea Press. pp 97–111. Richards AJ. 1990. Studies in Garcinia, dioecious tropical forest trees: The origin of the mangosteen (G. mangostana L.). Botanical Journal Linnean Society 103: 301–8. Rivera R, Edwards KJ, Barker JHA, Arnold GM, Ayad G, Hodgkin T, Karp A. 1999. Isolation and characterization of polymorphic microsatellites in Cocos nucifera L. Genome 42: 668–75. Rosendahl P, Yen DE. 1971. Fossil sweet potato remains from Hawaii. Journal of the Polynesian Society 80: 379–85.
Ross M. 1996. Reconstructing food plant terms and associated terminologies in Proto-Oceanic. In: Lynch J, Faafo P. editors. Oceanic studies: Proceedings of the First International Conference on Oceanic Linguistics. Pacific Linguistics Series C–133. Canberra, Australia: Australian National University. pp 163–221. Sauer CO. 1952. Agricultural origins and dispersals. New York: American Geographical Society. Sauer CO. 1971. A reevaluation of the coconut as an indicator of human dispersal. In: Riley CL, Kelley JC, Pennington CW, Rands RL, editors. Man across the sea. Austin, TX: University of Texas Press. pp 309–19. Simmonds NW. 1962. The evolution of bananas. London: Longman. Spriggs M. 1996. Early agriculture and what went before in Island Melanesia: Continuity or intrusion. In: Harris DR, editor. The origins and spread of agriculture and pastoralism in Eurasia. London: UCL Press. pp 524–37. Stumpf MPH, Goldstein DB. 2001. Genealogical and evolutionary inference with the human Y chromosome. Science 291: 1738–42. Swadling P. 1997. Changing shorelines and cultural orientations in the Sepik-Ramu, Papua New Guinea: Implications for Pacific prehistory. World Archaeology. 29: 1–14. Swadling P, Araho N, Ivuyo B. 1991. Settlements associated with the inland Sepik-Ramu sea. Bulletin of the Indo Pacific Prehistory Association 11: 92–112. Takahashi C, Marshall JA, Bennett MD, Leitch IJ. 1999. Genomic relationships between maize and its wild relatives. Genome 42: 1201–7. Teulat B, Aldam C, Trehin R, Lebrun P, Barker JH, Arnold GM, Karp A, Baudouin L, Rognon F. 2000. An analysis of genetic diversity in coconut (Cocos nucifera L.) populations from across the geographic range using sequence tagged microsatellites (SSRs) and AFLPs. Theoretical and Applied Genetics 100: 764–71. Therin M, Fullager R, Terrence R. 1998. Starch in sediments: A new approach to the study of subsistence and land use in Papua New Guinea. In: Gosden C, Hather JG, editors. The prehistory of food. London: Routledge. pp 438–62. Vos P, Hogers R, Bleeker M, Reijans M, van der Lee T, Hornes M, Frijters A, Pot J, Peleman J, Kuiper M, Zabeau M. 1995. AFLP—A new technique for DNA fingerprinting. Nucleic Acids Research 23: 4407–14. Wagstaff SJ, Dawson MI. 2000. Classification, origin, and patterns of diversification of Corynocarpus (Corynocarpaceae) inferred from DNA sequences. Systematic Botany 125: 34–49. Weeks A, Daly DC, Simpson BB. 2005. The phylogenetic history and biogeography of the frankincense and myrrh family (Burseraceae) based on nuclear and chloroplast sequence data. Molecular Phylogenetics and Evolution 35: 85–101. Whistler WA. 1990. The other Polynesian Gourd. Pacific Science 44: 115–22.
197
regional studies and conclusion
Whistler WA. 1991. Polynesian plant introductions. In: Cox PA, Banack SA, editors. Islands, plants and polynesians. Portland, OR: Dioscorea Press. pp 41–66. Whitmore TC. 1981. Wallace’s line and plate tectonics. Oxford, UK: Clarendon Press. Williams JGK, Kubelik AR, Livac KJ, Rafalski JA, Tingey SV. 1990. DNA polymorphisms amplified by arbitrary primers are useful as genetic markers. Nucleic Acids Research 18: 6531–5. Wolfe KH., Li W-H., Sharp PM. (1987). Rates of nucleotide substitution vary greatly among plant mitochondrial, chloroplast and nuclear DNAs. Proc. Natl. Acad. Sci. USA 84: 9054–8. Wright SD, Yong CG, Dawson JW, Whittaker DJ, Gardner RC. 2000. Riding the Ice Age El Niño? Pacific biogeography and evolution of Meterosideros (Myrtaceae) inferred from nuclear ribosomal DNA. Proceedings of the National Acadademy of Sciences USA 97: 4118–23. Wright SD, Yong CG, Wichman SR, Dawson JW, Gardner RC. 2001 Stepping stones to Hawaii: A trans-equatorial dispersal pathway for Meterosideros (Myrtaceae) inferred from nrDNA (ITS + ETS). Journal of Biogeography 28: 769–74. Yen DE. 1973. The sweet potato in Oceania. Bishop Museum Bulletin 236. Honolulu, Hawaii: Bishop Museum Press. Yen DE. 1974. Arboriculture in the subsistence of Santa Cruz, Solomon Islands. Economic Botany 28: 247–84. Yen DE. 1990. Environment, agriculture and the colonization of the Pacific. In: Yen DE, Mummery JMJ, editors. Pacific Production Systems. Occasional Papers in Prehistory 18. Canberra, Australia: Department of
198
Prehistory, RSPacS, Australian National University. pp 258–77. Yen DE. 1991a. Polynesian cultigens and cultivars: The question of origin. In: Cox PA, Banack SA, editors. Islands, plants and polynesians. Portland, OR: Dioscorea Press. pp 67–95. Yen DE. 1991b. Domestication: The lesson from New Guinea. In: Pawley A, editor. Man and a half: Essays in Pacific anthropology and ethnobiology in honour of Ralph Bulmer. Auckland, New Zealand: Polynesian Society. pp 558–69. Yen DE. 1995. The development of Sahul agriculture with Australia as a bystander. Antiquity 69: 831–47. Zerega NJC, Ragone D, Motley T. 2004. Complex origins of breadfruit (Artocarpus altilis, Moraceae): Implications for human migrations in Oceania. American Journal of Botany 91: 760–6. Zhang DP, Carbajulca D, Ojeda L, Rossel G, Milla S, Herrera C, Ghislain M. 2000a. Microsatellite analysis of genetic diversity in sweet potato varieties from Latin America. Lima, Peru: CIP Program Report 1999–2000, CIP, pp 295–301. Zhang DP, Cervantes J, Huamán Z, Carey E, Ghislain M. 2000b. Assessing genetic diversity of sweet potato (Ipomoea batatas (L.) Lam.) cultivars from tropical America using AFLP. Genetics Resources and Crop Evolution 41: 659–65. Zhang DP, Rossel G, Kriegner A, Hijmans R. 2004. AFLP assessment of diversity in sweet potato from Latin America and the Pacific region: Its implications on the dispersal of the crop. Genetics Resources and Crop Evolution 51: 115–20.
13 Extraordinary Population Structure among the Baining of New Britain Jason A. Wilder and Michael F. Hammer
Introduction The manner in which human genetic diversity is partitioned across geographic space remains a relatively poorly understood aspect of human biology. Many studies of human population structure have focused on quantifying the overall magnitude of genetic subdivision, or the fraction of total genetic diversity that is shared among populations versus that fraction which is population specific. However, a second facet of population structure involves the geographic and/or socio-cultural scales at which subdivision becomes apparent between human groups (e.g. Destro-Bisol et al., 2004). For instance, groups sharing a similar cultural identity and history while also living in close geographic proximity may be expected to be relatively undifferentiated genetically. Here we describe a surprising contrast to this expectation in two samples of Baining-speaking groups from the island of New Britain that are included in the broader population analyses of Part II of this book. Despite sharing cultural affinity as Baining, having no known tendency toward strong endogamy, and living within less than 100 km from one another with no major geographic barriers, we find exceptionally high levels of genetic subdivision between these two Baining groups (e.g., similar to that observed among populations residing on different continents). The Baining live on the Gazelle Peninsula in the easternmost portion of New Britain (figure 13.1). They have historically lived in the mountainous interior of New Britain, with relatively few living in coastal environments. They are linguistically heterogeneous, speaking several closely related Papuan dialects or languages
(Fajans, 1997; Whitehouse, 1995). Despite this linguistic variability, Baining tend to self-identify as a unified cultural group. Baining society is characterized by an exceptional lack of social complexity. Individual nuclear families operate as relatively autonomous units and higher order social arrangements tend to be transient in nature, involving small numbers of inter-family relationships (Whitehouse, 1995). No importance is ascribed to one’s lineage or clan, and there are no major individual figures (chiefs, religious leaders, etc.) around whom Baining culture is organized (Fajans, 1997). Moreover, there is no historic concept of individual or collective land ownership, with control over a land parcel enduring only so long as it is actively cultivated by a family group. Families tend to be highly transient, moving as agricultural needs demand. The lack of material ownership has generally precluded a formal mechanism of inheritance between generations. Marriage customs of the Baining reflect the lack of regimented social structure that characterizes their culture. Marriages tend to be arranged by parents, but there is no transfer of material goods between families of the bride and groom at marriage. Furthermore, the act of marriage does not carry with it long-term social obligations between families. Marriage and sexual intercourse are strongly linked in Baining culture, and extra-marital sex is not condoned (Fajans, 1997; Hesse et al., 1982). Couples tend to be ambilocal after marriage, with an arrangement usually agreed upon by families prior to marriage. Monogamy is the most common marriage arrangement among the Baining, although polygyny is also widespread (Fajans, 1997). Wives in polygynous marriages normally live together in a single household,
199
regional studies and conclusion
Figure 13.1 The location of the two sampled Baining populations on the island of New Britain. Individuals from Marambu speak the Mali dialect; those from Rangulit speak Kaket.
although separation in multiple locations over a wide area is not uncommon. Baining relations are also influenced by an extraordinarily permissive concept of adoption between unrelated families. Estimates in some areas place the adoption rate at approximately 34% (Fajans, 1997), with the practice involving mutual exchanges of children at birth and unidirectional adoption of young children. Adoption among the Baining is typically done freely and is not a consequence of material necessity or duress. Few aspects of Baining culture would lead one to hypothesize that local groups are highly genetically differentiated from one another. In fact, although our goal entering this study was to describe the population genetics of the Baining (as representatives of a relatively small island population), we had no a priori expectation of uncovering significant differentiation among groups within our Baining sample. Our result highlights the degree to which extraordinary local-scale population structure will influence the patterns and distribution of genetic variation, and underscores the general lack of understanding of the determinants of structure on both local and broad scales.
200
Molecular Methods and Sampling Approach We take a multi-locus approach to examining the population genetics and patterns of subdivision among the Baining. Our study describes genetic variation at four unlinked loci that should be affected by demographic and reproductive dynamics in different ways. Two of these are on the X chromosome, APXL (apical protein-like Xenopus laevis) and DMD44 (dystrophin). The remaining two loci are the uniparentally inherited mtDNA and the non-recombining portion of the Y chromosome (NRY), which trace maternal and paternal lineages, respectively. The mtDNA and NRY should be affected by sex-specific demographic processes. In the past their comparative analysis has led to interesting insights into the relative effective population sizes of males and females (which may be an indicator of differences in reproductive parameters between the sexes) as well as sex-specific differences in migration rates among populations (Destro-Bisol et al., 2004; Hamilton et al., 2005; Oota et al., 2001; e.g., Salem et al., 1996; Seielstad et al., 1998; Wilder et al., 2004a, 2004b). Our coupling of X-linked and uniparentally inherited loci
Extraordinary Population Structure among the Baining of New Britain
also allows us to examine markers that will be differentially influenced by genetic drift because of their unequal effective population sizes (Ne). Under the assumption that all loci are evolving neutrally and that the effective breeding sex ratio is one, the X-linked loci are expected to have Ne values that are three times that of uniparentally inherited loci. Because of this, the X-linked loci will be less influenced by recent genetic drift (taking longer to exhibit differentiation between populations, for instance). For each individual DNA sample all loci were re-sequenced in their entirety, allowing us to estimate the level of polymorphism and the frequency spectra of mutations in an unbiased manner. At each locus (except the mtDNA) we examined only non-coding DNA. Specifically, we surveyed 5,441 base pairs (bp) of the 4th intron of APXL (corresponding to the region reported in Hammer et al. (2004)), 3,044 bp of the 44th intron of DMD (corresponding to the region reported in Nachman and Crowell (2000)), and 6,650 bp of non-coding sequence encompassing 13 separate Alu elements on the Y chromosome. For the mtDNA we surveyed 782 bp of sequence encompassing nearly all of the cytochrome oxidase 3 gene. Data from these latter loci have been reported in Wilder et al. (2004a). In the calculation of summary statistics, we considered only single nucleotide replacements as segregating sites (i.e., no indels or other changes in sequence length were included). Our sample of Baining individuals originates in two localities: 24 individuals sampled from the region of Marambu (where the Mali Baining dialect is spoken), and 24 sampled from Rangulit (speakers of the Kaket Baining dialect) (figure 13.1). The linguistic terms are used in other chapters of this book. Samples were collected with informed consent, and provided for this study by Jonathan Friedlaender. All individuals included in this study are male, allowing us to sequence a single haplotype in each sample at all four genetic loci.
Levels and Patterns of Polymorphism among the Baining The relative levels of diversity at the four loci examined here (table 13.1) follow the same rank order as seen in worldwide surveys of genetic diversity (Garrigan et al., in preparation): mtDNA has the highest level of variability, followed by DMD44, APXL, and, finally, the Y chromosome. To a large extent, these differences are driven by variation in the mutation rate across loci. A full comparison of levels of diversity within the Baining compared with those of other world populations at these four loci will be made elsewhere; however, preliminary results indicate that the Baining exhibit low levels of variability relative to other population groups (which were sampled from Asia, Europe, and Africa).
Table 13.1 Nucleotide Diversity at Four Loci in Baining Populations π (%)
θ (%)
Y-Alu All Baining Marambu Rangulit
0.019 0.014 0.017
0.010 0.008 0.011
3 2 3
4 3 3
1.993* 1.713 1.228
mtDNA All Baining Marambu Rangulit
0.143 0.184 0.077
0.116 0.103 0.069
4 3 2
4 3 3
0.552 1.977* 0.270
APXL All Baining Marambu Rangulit
0.040 0.047 0.024
0.045 0.038 0.038
10 7 7
7 4 5
DMD44 All Baining Marambu Rangulit
0.112 0.111 0.117
0.067 0.071 0.080
9 8 9
17 14 13
Baining mean Marambu mean Rangulit mean
0.079 0.089 0.059
0.060 0.055 0.050
6.50 5.00 5.25
S
h
8.00 6.00 6.00
TD
−0.302 0.771 −1.108 1.883* 1.805 1.527 1.032 1.567 0.479
*p<0.05.
Considering the two Baining samples separately, the mean level of diversity across the four genetic loci is slightly higher in the Marambu than the Rangulit population (table 13.1). This is true for two different measures of diversity, π, or the per nucleotide heterozygosity (Nei and Li, 1979), and Watterson’s (1975) estimator of θ. Under neutral equilibrium conditions these measures are estimates of the quantity 2Nefµ for the mtDNA, 2Nemµ for the Y chromosome and 3Neµ for the X-linked loci, where Nef and Nem represent the female and male effective population sizes, respectively (in contrast to the total effective population size, Ne), and µ represents the mutation rate (Hedrick, 2005). Despite the average pattern of slightly higher levels of diversity in Marambu than Rangulit, not all loci follow this trend. Marambu harbors more diversity than Rangulit at mtDNA and APXL, while the opposite is true at the Y chromosome and DMD44. Thus, the overall picture that emerges from the DNA diversity data is that the two populations do not differ greatly from one another in terms of overall level of variability. This is most simply shown by the number of SNPs and haplotypes observed in each population (table 13.1). At no locus does the observed number of SNPs or haplotypes differ by more than one between Marambu and Rangulit. On the other hand, the pattern of polymorphism in the two Baining populations appears to differ. One way to characterize the pattern of polymorphism is to estimate the frequency spectrum of mutations, which can be
201
regional studies and conclusion
summarized by Tajima’s D (Tajima, 1989) statistic (shown in table 13.1). Tajima’s D is a measure of the standardized difference between the estimators π and θ. Under neutral equilibrium conditions the expected value of Tajima’s D is close to zero, with negative values indicating an excess of rare polymorphisms, and positive indicating an excess of intermediate frequency variants. In our present study the value of Tajima’s D is higher at every locus in Marambu than Rangulit. In Marambu, Tajima’s D values at individual loci range from 0.771 to 1.993, with the highest value (observed for the mtDNA) differing significantly from the neutral equilibrium expectation (p = 0.041, determined via coalescent simulations). Rangulit, in contrast, exhibits considerable heterogeneity with respect to Tajima’s D across loci. Observed values range from –1.108 at APXL to 1.527 at DMD44, and none differs significantly from equilibrium expectations. When observed across multiple unlinked loci, as is the case here, the uniform skew in the Marambu frequency spectrum likely indicates that the population is at a non-equilibrium state with respect to demographic processes (as opposed to a skew in the frequency spectrum caused by natural selection, which would be unlikely to affect more than a single locus). Such a positive skew can result from a contraction in population size, which tends to eliminate rare genetic variants (Tajima, 1989). The patterns at all four genetic loci are consistent with the hypothesis that Marambu has experienced a recent reduction in population size. In contrast, we see no such evidence for a change in size of the Rangulit population.
Population Differentiation within the Baining The overall similarity in the level of diversity between Marambu and Rangulit belies significant differences in the composition of genetic variants within these two populations. This genetic differentiation is shown by two statistics, Fst and the proportion of shared sites (PSS), for this pair of populations. Fst measures the partitioning of diversity within and between a set of populations. The statistic was developed by Wright (1931) to describe the balance between migration and genetic drift under an island model of population structure. For the mtDNA this value approximates 1/(1+2Nefm), where m is the sexspecific migration rate (the proportion of migrants per generation). The same equation defines Fst for the Y chromosome, although Nef is replaced with Nem. For X-linked loci the value approximates 1/(1+3Nem). In table 13.2 we report a modified version of Fst (Hudson et al., 1992) which accounts for molecular distances between haplotypes in addition to differences in their observed frequency. The proportion of shared sites is simply an accounting of the fraction of variable sites that segregate
202
Table 13.2 Fst and Proportion of Shared Sites (PSS) between Marambu and Rangulit Locus Y-Alu mtDNA APXL DMD44
Fst
PSS
0.311 0.162 0.212 −0.037
0.666 0.250 0.400 0.889
in both populations. Values for both of these statistics range widely across the four loci. At one extreme is DMD44, which has a slightly negative value of Fst and a high PSS value 0.889, both of which are indicative of little differentiation between populations. In contrast, all other loci indicate substantial subdivision between populations. Observed values of Fst range from 0.162 to 0.311, and PSS varies from 0.250 to 0.666. How can we gauge the relative level of subdivision observed between the two Baining populations? We recently examined 10 worldwide populations using the same regions of Y chromosome and mtDNA (Wilder et al., 2004a) and current work is also examining patterns of divergence at DMD44 and APXL in these populations (Garrigan et al., in preparation). For three loci (the Y chromosome, mtDNA, and APXL) values of Fst between Marambu and Rangulit are similar to, if not greater than, the worldwide mean for all pair-wise combinations of populations (excluding comparisons to the Baining). Figure 13.2 presents Fst-based trees for all four loci. Of the four loci examined here, only DMD44 has a relatively low degree of divergence between Baining populations (Fst for this locus is, in fact negative, at –0.037, table 13.2). Despite this, the observation of three loci with a degree of genetic differentiation similar to that observed between widely spaced populations from different continents is extraordinary, given that residents of Rangulit and Marambu live less than 100 km from one another on the same island.
How Important Are Sex-Specific Factors in Shaping Observed Genetic Patterns? Because of their sex-specific inheritance, the Y chromosome and mtDNA can be used to estimate parameters that may vary between females and males, such as effective population size and migration rate. Early work on this subject suggested a global-scale pattern whereby the rate of migration between populations is much higher for females than males, reflected by higher values of Fst for the Y chromosome than mtDNA (Seielstad et al., 1998). We have recently revisited this issue, using more powerful
Extraordinary Population Structure among the Baining of New Britain
Figure 13.2 Fst trees describing patterns of differentiation among the two Baining populations (Marambu and Rangulit) along with nine other globally distributed populations. Individual trees are shown for each locus, and branch lengths are scaled uniformly across trees. In the case of DMD44 the apparent differentiation between Marambu and Rangulit (labeled with an asterisk on the figure) is a spurious result of a difference in genetic distances between these two populations and Papua New Guinea.
molecular methodology and a more even sampling strategy, and find no such bias in migration rates among 10 widely dispersed populations (Wilder et al., 2004a). At smaller geographic scales, however, differences in sexspecific migration rates do sometimes leave a detectable imprint on mtDNA and Y chromosome variability, yielding reciprocal patterns of structure in matrilocal versus patrilocal populations (Hamilton et al., 2005; Oota et al., 2001; Salem et al., 1996). It may be that broad-scale
geographic patterns are primarily influenced by ancient demographic processes, while local-scale patterns are more strongly influenced by recent or ongoing processes. In addition to differences in migration rate between populations, the sexes may vary with respect to patterns of reproductive success. Based on deeper estimated times to the most recent common ancestor (TMRCAs) for the mtDNA than the Y chromosome in many populations, we recently suggested (Wilder et al., 2004b) that the effective
203
regional studies and conclusion
population sizes of the sexes are unequal, with a skew toward more females than males. This may be caused by a lower long-term variance in female reproductive success, a mechanism that is compatible with many wellknown behavioral attributes of humans, including the widespread practice of polygyny (Murdock, 1981). A cursory examination of mtDNA and Y chromosome Fst values in the Baining (table 13.2) suggests a higher rate of migration between populations for females than males, based on the lower Fst values for the mtDNA than Y chromosome (0.162 versus 0.311). However, this interpretation relies on the assumption that male and female effective population sizes are equal (i.e., Nef = Nem), and differences in Fst only reflect variation in the migration parameter. Otherwise, differences between Nef and Nem must be accounted for when estimating sex-specific migration rates. In the two Baining populations, we can examine simultaneously both sex-biased gene flow and variation in sex-specific effective population size using a single analytical framework. Because the mtDNA and Y chromosome are non-recombining genetic units, we can create non-reticulating gene trees that describe genealogical relationships among segregating alleles in the two Baining populations (figure 13.3). These gene trees can then be used in a coalescent-based procedure to estimate population parameters using the maximum likelihood framework implemented by the Genetree software package (Bahlo and Griffiths, 2000). Using this analytical approach, we can evaluate parameters that describe the mtDNA and Y chromosome in the Marambu and Rangulit populations both independently and jointly. In this analysis alleles were defined using all observed polymorphisms at each locus, which in the case of the Y chromosome included a single-base indel that segregates in both populations. In the simplest implementation of the Genetree program, we treat each population separately to estimate the maximum likelihood value of the population-mutation
Figure 13.3 Gene trees showing relationships among Y chromosome and mtDNA haplotypes. Numbers of individuals with each haplotype are indicated for the Rangluit and Marambu populations.
204
Table 13.3 Sex-specific Effective Population Size (Nem or Nef) and TMRCA of Haploid Loci in Marambu and Rangulit Populations under a Constant-Sized Population Model with no Migration among Demes θml
Nem or Nef
Y-Alu Marambu Rangulit
0.72 0.96
512 683
29.2 (12.2) 37.4 (14.2)
mtDNA Marambu Rangulit
0.72 0.54
1198 899
66.3 (28.6) 41.7 (18.7)
TMRCA (kya, s.d.)
parameter (θml) for each locus. Here, we consider only a constant-size population model, as models incorporating population growth are not a significantly better fit to the data (results not shown). Resulting estimates of θml are shown in table 13.3 (note that the θml values shown in table 13.3 are per locus, not per bp). Concordant with the diversity values shown in table 13.2, we observe a higher value of θml in Marambu than Rangulit at the mtDNA, and the opposite pattern at the Y chromosome. With knowledge of the mutation rate at the Y chromosome and mtDNA we can extend our study to a direct comparison of estimates of Nef, Nem, and the TMRCA for each locus. In this analysis we use the same mtDNA mutation rate estimated in our 2004 study (1.58×10−8 mutations per site per year; Wilder et al., 2004b); and a rate of 4.19×10−9 mutations per site per year for the Y chromosome (which we infer proportionally from the increased SNP density over the regions from which the mutation rate was estimated in Wilder et al., 2004b). Incorporation of these mutation rates into estimates of effective population size and TMRCA are shown in table 13.3. The data are consistent with a higher value of Nef than Nem, along with deeper mtDNA than Y chromosome TMRCAs, in both populations. In Marambu there is a more than 2-fold increase in the female versus male effective population size, while in Rangulit the disparity is somewhat less, at 1.3-fold. The preceding analysis assumes that both Marambu and Rangulit are closed systems, with no migration into either from outside groups. However, we can also use the Genetree framework to examine the two populations as demes following Wright’s island model of subdivision, thus treating the two populations together. In this analysis we jointly estimate not only the value of θml, but also the migration parameter (m) between the two demes. Our methodology for joint-parameter estimation follows the procedure outlined in Wilder et al. (2004b). In that study, we jointly estimated θml and the population growth parameter (β). In this work, we assume a constant population size (see above) and instead jointly estimate m and θml
Extraordinary Population Structure among the Baining of New Britain
(moreover, we only consider symmetrical models of migration). In theory, it is possible to estimate jointly more than two parameters, although it is computationally untenable for the present work. In addition to parameter estimation, Genetree evaluates the likelihood of the resulting data, given a particular demographic model. In this case, we explicitly test whether the data better fit a panmictic or structured population model by comparing estimates of the likelihood of the data in each case. The results of our coalescent-based population structure analysis are shown in table 13.4. Using the modeltesting procedure set forth by Bahlo and Griffiths (2000), which is a modified likelihood ratio test (comparing the likelihood values (L) listed in table 13.4 for structured versus panmictic models), we find the structure model to be a significantly better fit to the data than panmixia for both loci; the associated p value is 0.0039 in the case of the mtDNA and 0.0001 for the Y chromosome. Thus, in the following discussion, we will focus on parameters estimated under the model of population structure. Incorporation of subdivision into our analysis of sexspecific migration and effective population size yields results that contrast sharply with a simple examination of Fst. Table 13.4 shows the jointly estimated values of Nef and Nem, together with the quantity Nm for both sexes. Estimates of both parameters are higher for females than males (Nef: 1098; Nem: 320; Nefm: 1.34; Nemm: 0.78). However, when we focus on m directly, by accounting for Nef and Nem, we observe a migration parameter that is twice as high for males than females (0.0024 versus 0.0012). In other words, our data support a model where the proportion of males that are migrating between populations each generation is higher for males than females, despite a higher Fst for the Y chromosome than mtDNA. The cause of this apparent discrepancy is a much smaller male than female effective population size. The implication of this finding with respect to the use of Fst to infer sex-specific patterns of gene flow may be quite widespread. For instance, a recent study by Hamilton et al. (2005) also jointly estimated both sex-specific effective population size and Nm in a sample of patrilocal and
matrilocal hill tribes from northern Thailand (using a Bayesian statistical approach). With respect to Nm their results are consistent with an earlier study by Oota et al. (2001), in that females have a much higher value than males (7.4 versus 0.5) in patrilocal groups, while the numbers are reversed where matrilocality is prevalent (3.1 versus 3.9). However, Hamilton et al. (2005) also estimate an extraordinary difference in Nef versus Nem (with a female to male ratio of 21:1 in patrilocal populations, and 2.7:1 in matrilocal populations). The implication, left unexplored in the Hamilton et al. (2005) study, is that a higher proportion of males than females are migrating into populations in patrilocal groups and vice versa in matrilocal groups. One way to reconcile this finding is to consider that emigrants may experience greater reproductive success than non-migrants, boosting the apparent proportion of migrant individuals. Regardless, incorporation of variation in male and female effective population size appears essential to understanding the extent and nature of sex-specific demographic processes.
Conclusions With respect to the Baining, the composite picture that emerges from our study is that of a group divided into extremely differentiated sub-populations. This differentiation rivals that observed between populations sampled from different continents (e.g. Wilder et al., 2004a). Ours is not the first study to find exceptional levels of population structure over very small geographic scales. For example, Zhivotovsky et al. (2001) describe surprisingly high differentiation at both autosomal and Y chromosome markers among geographically close endogamous communities from Pakistan. A similar pattern has been observed at the scale of hundred of kilometers among populations in highland Papua New Guinea (Stoneking et al., 1990). In the case of the Pakistani communities, they are described as being indistinguishable with respect to cultural, historical, or geographic factors, and the extent of their differentiation was unexpected. In that case, a
Table 13.4 Joint Maximum Likelihood Estimate of Sex-Specific Gene Flow and Population Size (a model of panmixia is included for comparison) θml
Nem or Nef
Nm
m 0.0024 0.0012
Structured Y-Alu mtDNA
0.45 0.66
320 1098
0.78 1.34
Panmixia Y-Alu mtDNA
0.83 0.88
590 1464
-
TMRCA (kya, s.d.)
L
33.3 (13.5) 74.4 (31.8)
3.08×10−10 1.33×10−8
31.7 (12.7) 71.6 (30.4)
8.70×10−8 1.86×10−6
205
regional studies and conclusion
high level of endogamy may have played a role in causing inter-population divergence. In Papua New Guinea impassable geographic barriers have been suggested to cause isolation among populations. Based on anthropological descriptions, high levels of endogamy are not prevalent among the Baining and the island of New Britain is not characterized by the same mountainous terrain as highland Papua New Guinea. Thus, among the Baining there is no obvious, social, cultural, or geographical feature that would explain such a high level of differentiation, although it is possible that language variability may act to reinforce genetic isolation among groups. There are a number of hypotheses that are compatible with the observed data. For instance, the two Baining groups sampled here may represent fractioned remnants of a larger ancestral population. Under this scenario, the high level of observed subdivision may be a consequence of differential lineage sorting within demes, coupled with low levels of migration between demes. As mentioned, the population genetic data are compatible with a recent population contraction, at least for the Marambu, which may have accompanied the break-up of a large ancestral deme. Alternatively, the apparent cultural affinity between the two groups studied here may not reflect recent common ancestry. Instead, it may be that these two groups arose from different ancestral populations and that Baining cultural identity was later acquired by residents of Marambu and/or Rangulit without genetic homogenization. Several such cases of cultural adoption without gene flow have been documented with respect to the transition between hunter-gatherer and agricultural cultures (see examples in Diamond and Bellwood, 2003, for instance). Regardless of the precise historical scenario producing extant patterns of differentiation, the maintenance of this subdivision must be enforced by relatively low levels of gene flow between the Rangulit and Marambu samples. Furthermore, the small size of these groups, as indicated by the low estimates of effective population size and TMRCA in tables 13.3 and 13.4, suggest that they are influenced strongly by genetic drift, allowing them to differentiate quite rapidly. One factor unique to the Baining that may underlie their low effective population size and exacerbate the rate of genetic drift is their high rate of adoption between families (Fajans, 1997). Although the degree to which this practice influences the average inbreeding coefficient among the Baining is not understood, widespread adoption should tend to make inbreeding-avoidance behaviors (such a consanguineous marital taboos) more difficult to enact. The pattern of strong subdivision between the Rangulit and Marambu samples was seen at only three of the four loci surveyed in this study (the Y chromosome, mtDNA, and APXL). In contrast, X-linked DMD44 actually had a negative value of Fst, indicative of no genetic
206
differentiation between populations (table 13.2). Therefore, it is difficult to reconcile patterns of population subdivision as revealed by X-linked versus uniparentally inherited loci. Expected rates of drift and lineage sorting vary between these loci, so that differentiation should accumulate more slowly for X-linked loci due to their larger effective population size. Based on the loci included in this study, we cannot distinguish between a recent origin of the population structure observed between Rangulit and Marambu (as suggested by the contrast between DMD44 and the uniparentally inherited loci) versus relatively more ancient subdivision (as suggested by the high levels of differentiation shared by APXL and the uniparentally inherited loci). Examination of additional genetic loci from the X chromosome and autosomes will be necessary to distinguish between these two cases. Our inferences regarding sex-specific demographic patterns among the Baining uncovered a strong skew in effective breeding ratio of the sexes, with a 3:1 ratio of females to males in the total sample (table 13.4). Moreover, while we find that the absolute number of migrants (i.e., the parameter Nm) between the populations is higher for females than males, the proportion of migrants (i.e., m) is actually higher for males than females. Based on marriage residency patterns, which are ambilocal among the Baining, we have no a priori expectation of higher rates of gene flow for either sex. However, our observations here hint that sex-specific accounting of the total number of migrants between demes, which is the approach of all previous Fst-based studies, may not be the most revealing way to consider the impact of sex-specific demography on population structure. The finding that the male breeding population is composed of a higher proportion of migrants than is the female population (i.e., the parameter m is higher for males than females) suggests that male migrants among the Baining may contribute more than migrant females with respect to reproductive success. Whether this is a result that is peculiar to the Baining, or a common feature of male migration, remains an open question. The results of Hamilton et al. (2005), where males have a higher value of m in patrilocal groups and females a higher m in matrilocal groups, suggest that the relationship between marriage residency pattern and gene flow due to migration may be more complex than is currently appreciated.
References Bahlo M and Griffiths RC. 2000. Inference from gene trees in a subdivided population. Theoretical Population Biology 57: 79–95. Destro-Bisol G, Donati F, Coia V, Boschi I, Verginelli F, Caglia A, Tofanelli S, Spedini G, and Capelli C. 2004.
Extraordinary Population Structure among the Baining of New Britain
Variation of female and male lineages in sub-Saharan populations: the importance of sociocultural factors. Molecular Biology and Evolution 21: 1673–82. Diamond J and Bellwood P. 2003. Farmers and their languages: the first expansions. Science 300: 597–603. Fajans J. 1997. They make themselves: work and play among the Baining of Papua New Guinea. Chicago, IL: University of Chicago Press. Hamilton G, Stoneking M, and Excoffier L. 2005. Molecular analysis reveals tighter social regulation of immigration in patrilocal populations than in matrilocal populations. Proceedings of the National Academy of Sciences of the USA 102: 7476–80. Hammer MF, Garrigan D, Wood E, Wilder JA, Mobasher Z, Bigham A, Krenz JG, and Nachman MW. 2004. Heterogeneous patterns of variation among multiple human X-linked loci: the possible role of diversityreducing selection in non-Africans. Genetics 167: 1841–53. Hedrick PW. 2005. Genetics of populations. Boston, MA: Jones and Bartlett Publishers. Hesse KD, Aerts T, Hesse KD, and Hesse KD. 1982. Baining life and lore. Port Moresby, Papua New Guinea: Institute of Papua New Guinea Studies. Hudson R, Slatkin M, and Maddison W. 1992. Estimation of levels of gene flow from DNA sequence data. Genetics 132: 583–9. Murdock GP. 1981. Atlas of world cultures. Pittsburgh, PA: University of Pittsburgh Press. Nachman MW and Crowell SL. 2000. Contrasting evolutionary histories of two introns of the Duchenne muscular dystrophy gene, Dmd, in humans. Genetics 155: 1855–64. Nei M and Li WH. 1979. Mathematical model for studying genetic variation in terms of restriction endonucleases. Proceedings of the National Academy of Sciences of the USA 76: 5269–73.
Oota H, Settheetham-Ishida W, Tiwawech D, Ishida T, and Stoneking M. 2001. Human mtDNA and Y-chromosome variation is correlated with matrilocal versus patrilocal residence. Nature Genetics 29: 20–1. Salem AH, Badr FM, Gaballah MF, and Pääbo S. 1996. The genetics of traditional living: Y-chromosomal and mitochondrial lineages in the Sinai Peninsula. American Journal of Human Genetics 59: 741–3. Seielstad MT, Minch E, and Cavalli-Sforza LL. 1998. Genetic evidence for a higher female migration rate in humans. Nature Genetics 20: 278–80. Stoneking M, Jorde LB, Bhatia K, and Wilson AC. 1990. Geographic variation in human mitochondrial DNA from Papua New Guinea. Genetics 124: 717–33. Tajima F. 1989. The effect of change in population size on DNA polymorphism. Genetics 123: 597–601. Watterson GA. 1975. On the number of segregating sites in genetical models without recombination. Theoretical Population Biology 7: 256–76. Whitehouse H. 1995. Inside the cult: religious innovation and transmission in Papua New Guinea. Oxford, UK: Oxford University Press. Wilder JA, Kingan SB, Mobasher Z, Pilkington MM, and Hammer MF. 2004a. Global patterns of human mitochondrial DNA and Y-chromosome structure are not influenced by higher migration rates of females versus males. Nature Genetics 36: 1122–5. Wilder JA, Mobasher Z, and Hammer MF. 2004b. Genetic evidence for unequal effective population sizes of human females and males. Molecular Biology and Evolution 21: 2047–57. Wright S. 1931. Evolution in Mendelian populations. Genetics 16: 97–159. Zhivotovsky LA, Ahmed S, Wang W, and Bittles AH. 2001. The forensic DNA implications of genetic differentiation between endogamous communities. Forensic Science International 119: 269–72.
207
14 Immunoglobulin Allotypes as a Marker of Population History in the Southwest Pacific Moses S. Schanfield, Frank B. Austin, Peter B. Booth, D. Carlton Gajdusek, Richard W. Hornabrook, Keith P. W. McAdams, Jan J. Saave, Susan W. Serjeantson, and Graeme W. Woodfield
Introduction In Part II of this book, various genes, but particularly mtDNA and Y DNA markers, were used to investigate questions involving the peopling of New Guinea and Island Melanesia. Prior to the discovery and use of DNAbased testing, the only other genetic markers that had the variation among human populations seen by mtDNA and Y DNA were those markers found on the heavy chains of human immunoglobulins. The markers on the immunoglobulin genes are called allotypes, and were originally detected serologically. That is, each marker was detected by a specific antibody test. The data for many of the allotypes are reviewed in Schanfield and van Loghem (1986) and Dard et al. (2001). Though serologically detected initially, immunoglobulin allotypes represent DNA substitutions. As with the mtDNA and Y chromosome markers, the heavy chain immunoglobulin allotypes are inherited in blocks referred to as haplotypes. Each haplotype spans the loci IGHG3, IGHG1, through IGHG2 and IGHA2 near the terminal end of chromosome 14 (Schanfield and van Loghem, 1986). The anthropologically useful markers are: G1M A,F,X,Z; G1/3M G5; G2M N; G3M B0,1,3,4,5,C3,5,G,S and T; and A2M 1,2. The nomenclature indicates that they are located on different heavy chain loci (IGH) with markers on IGHG1 (G1M), IGHG2
208
(G2M), IGHG3 (G3M), and IGHA2 (A2M). The exception to the rule is the G1/3M G5 allotype which occurs on IGHG3, with G3M G in all non-Sub-Saharan African populations and IGHG1 with G1M A,Z in Sub-Saharan populations. The nomenclature of human immunoglobulin allotypes was standardized by a WHO meeting in 1974 (Ropartz et al., 1976) and modified by the creation of the International System of Genetic Nomenclature in 1987 (Shows et al., 1987). The alphanumeric nomenclature will be used through out this chapter following those guidelines. Table 14.1 contains a list of informative heavy chain haplotypes and approximate frequencies in several populations from around the world showing the characteristic variation (from Schanfield and Fudenberg, 1975). The much simpler light chain markers show regional variation but are less useful and have not been included in this table. Based on an analysis of immunoglobulin heavy chain haplotypes with one exception, major human populations can be separated either by the presence of specific haplotypes, or by markedly different frequencies of haplotypes. The only two populations that cannot be so separated are the central Australian Aborigines, and the Paleo-Indians of South America (Schanfield, 1980). To simplify the presentation of data, the IGH haplotypes presented in table 14.1 will be reduced as follows: the
Immunoglobulin Allotypes as a Marker of Population History in the Southwest Pacific
Table 14.1 Anthropologically Useful Immunoglobulin Heavy Chain Haplotypes and Approximate Frequencies from Several Worldwide Populationsa New Guinea Population N samples Haplotype F N B0,1,3,4,5 1 F N′ B0,1,3,4,5 1 A,Z N′ G,G5 1 A,Z N′ G,G5 2 A,X,Z N′ G,G5 1 A,X,Z N′ G,G5 2b A,Z N B0,1,3,4,5 1 A,Z N B0,1,3,4,5 2 A,Z N′ B0,1,3,4,5 1* A,Z N′ B0,1,3,4,5 2* A,Z N′ B0,3,5,S,T 1 A,Z N′ B0,3,5,S,T 2 A,F N B0,1,3,4,5 1 A,F N B0,1,3,4,5 2 A,F N′ B0,1,3,4,5 1 A,F N′ B0,1,3,4,5 2 A,Z N′ B0,1,C3,5 1* A,Z N′ B0,1,C3,5 2* A,Z N′ B0,1,4,5C3 1* A,Z N′ B0,3,5,S 1* A,Z N′ B0,3,5,S 2* OTHERS
Surinama
Europe
China
Japan
PAP
AN
612
58
98
587
207
62
0.450 0.292 0.175 0.005 0.045 0.005 0.000 0.000 0.002 0.003 0.003 0.008 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.012
0.000 0.000 0.121 0.000 0.069 0.000 0.000 0.000 0.000 0.000 0.000 0.026 0.112 0.672 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
0.000 0.000 0.349 0.057 0.156 0.026 0.000 0.000 0.000 0.000 0.047 0.214 0.010 0.141 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
0.000 0.000 0.411 0.031 0.027 0.000 0.384 0.003 0.025 0.000 0.000 0.000 0.020 0.084 0.009 0.005 0.000 0.000 0.000 0.000 0.000 0.001
0.000 0.000 0.171 0.027 0.010 0.000 0.126 0.010 0.027 0.000 0.000 0.000 0.162 0.403 0.048 0.017 0.000 0.000 0.000 0.000 0.000 0.000
0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.137 0.476 0.000 0.000 0.000 0.000 0.000 0.000 0.040 0.234 0.024 0.024 0.057 0.008
a
The samples from Surinam serve as a surrogate for Sub-Saharan Africans. The allotype G2M N’ indicates that the haplotype was tested for G2M N but was found to be negative. *Indicates Sub-Saharan haplotypes.
b
Eurasian haplotypes 3 and 4 are equivalent to GM*A G, and 5 and 6 are equivalent to GM*A,X G; the AustraloPapuan marker haplotypes 7, 8, 9, and 10 are equivalent to GM*A B; the North Asian marker haplotypes 11 and 12 to GM*A T; and the Southeast Asian marker haplotypes 13, 14, 15, and 16 to GM*A,F B. The distribution of GM haplotypes in East Asia, Southeast Asia, and Oceania data from the literature are included in Table 14.2. Note the very extensive coverage in Oceania, as well as the degree of variation. Given this battery of haplotypes, how can the immunoglobulin allotypes help us look at the origins and relationship of the populations of Australia, New Guinea, and Near and Remote Oceania? There are several questions that recur through this volume. Archeologically, Australia and New Guinea, as part of Sahul were populated first at approximately 60,000 years ago with the Bismarck Archipelago and Solomon Islands populated by 30–40,000 years ago (see chapter 2). The languages of Australia and the Papuan languages of New Guinea are significantly different from each other, indicative of a long period of isolation (Bowern and Koch, 2004). Are there genetic markers shared among these populations
that link them together? Is there any genetic evidence to suggest more than one human migration to Sahul during the Pleistocene? During the Holocene migrations from Southeast Asia brought different groups of Austronesianspeaking populations to Near and Remote Oceania. What can the immunoglobulin markers tell us about these migrations? To judge from tables 14.1 and 14.2, there are several haplotypes that are informative concerning these issues. The haplotype GM*A B only occurs in New Guinea and the northern coast of Australia. Since GM*A B occurs in Austronesian speakers, this may reflect post-contact gene flow. Is it uniformly distributed among all Austronesian (AN) speakers, or only those in close contact with Papuan (PAP) speakers? The AN speakers are postulated to have arisen in Southeast Asia. Is there a population marker for Southeast Asia? The haplotype GM*A,F B occurs in highest frequency in Thailand and Taiwan, decreases in Japanese, is not found in Europeans or Sub-Saharan Africans, and appears to be much higher in AN speakers than in PAP speakers. What is the distribution of this haplotype in PAP populations? Does it only occur among AN speakers, or is it distributed throughout New Guinea
209
regional studies and conclusion
Table 14.2
Summary Allotype Frequencies from East Asia and the Pacific
Language
Location
Altaic Altaic Sino-Tibetan Sino-Tibetan Sino-Tibetan Tai/Kadai) Tai/Kadai) Tai/Kadai) ANa AN AN AUSTb AUST AUST AUST AUST Non-ENGHc PAP/AN? Non-ENGH Non-ENGH Non-TNGPe AN Non-TNGP ENGH Non-ENGH ENGHf ENGH ENGH ENGH Non-ENGH ENGHg Non-ENGH Non-TNGP AN Non-ENGH Non-ENGH ENGH AN AN/PAP? Non-ENGH Non-ENGH AN Non-ENGH AN AN AN PAP AN AN/PAP? PAP PAP AN AN AN PAP Polynesian
Japan Korea China—Northern China—Taiwan China—Southern Thailand—Northern Thailand—Central Thailand—South Taiwan—Lowland Taiwan—Highland Indonesia—Surabaya Java Aust—Cent/West Desert Aust—Kimberleys Aust—Arnhem Land Aust—Gulf of Carpentaria Aust—Cape York West Papua—SE Coast West Papua—NE Coast PNG—Westernd PNG—West Sepik PNG—West Sepik PNG—West Sepik PNG—East Sepik PNG—South Highlands PNG—South Highlands PNG—Enga PNG—Western Highlands PNG—Chimbu PNG—Eastern Highlands PNG—Eastern Highlands PNG—Madang—Karkar PNG—Madang PNG—Madang PNG—Madang PNG—Gulf PNG—Morobe PNG—Morobe PNG—Morobe PNG—Morobe—Lae PNG—Northern PNG—Milne Bay PNG—Milne Bay PNG—Central PNG—Central New Britain—Western New Britain—Eastern New Britain—Eastern New Ireland—Tench Island New Ireland—Kavieng Bougainville—Central Bougainville—Southern Bougainville—Southern Solomons—Malaita Solomons non-Malaita Solomons Solomons—Rennel Bellona
210
N
*A G
*A,X G
*A B
*A,F B
others
*A T
KM*1
950 195 167 208 325 344 127 153 151 241 99 1189 342 96 556 424 677 254 1151 11 53 10 423 205 428 196 279 29 470 235 492 601 217 377 46 1314 138 617 176 165 19 119 430 417 200 386 97 38 226 1102 1158 351 774 138 5 152
0.470 0.507 0.404 0.220 0.152 0.014 0.090 0.073 0.036 0.166 0.158 0.743 0.573 0.638 0.546 0.751 0.152 0.104 0.301 0.909 0.693 0.600 0.805 0.594 0.228 0.786 0.642 0.741 0.710 0.408 0.415 0.136 0.278 0.109 0.435 0.490 0.569 0.154 0.364 0.257 0.763 0.139 0.359 0.016 0.141 0.202 0.236 0.324 0.095 0.264 0.047 0.066 0.124 0.065 0.100 0.185
0.175 0.213 0.153 0.086 0.060 0.023 0.056 0.071 0.010 0.041 0.089 0.257 0.238 0.264 0.163 0.232 0.013 0.004 0.006 0 0 0 0.001 0.109 0.008 0.138 0.187 0.155 0.130 0.022 0.045 0.009 0.026 0.008 0.022 0.038 0.159 0 0.020 0.022 0.079 0.017 0.018 0 0.005 0.234 0.312 0.053 0.022 0.040 0.044 0.029 0.016 0.045 0.100 0.171
0 0 0 0 0 0 0 0 0 0 0.005 0 0.189 0.099 0.291 0.018 0.834 0.016 0.692 0.091 0.207 0.150 0.180 0.297 0.754 0.076 0.167 0.103 0.120 0.403 0.168 0.110 0.071 0.127 0.337 0.263 0.212 0.489 0.122 0.194 0.026 0.109 0.336 0.384 0.045 0.010 0.023 0 0 0 0 0 0 0 0 0
0.078 0.141 0.356 0.649 0.737 0.951 0.842 0.830 0.950 0.790 0.717 0 n.t. (0) 0 0 0 0.001 0.877 0.001 0 0.101 0.250 0.013 0 0.003 0 0.004 0 0.040 0.169 0.289 0.746 0.623 0.756 0.207 0.175 0.060 0.357 0.494 0.528 0.132 0.735 0.287 0.600 0.810 0.554 0.428 0.623 0.883 0.696 0.908 0.905 0.852 0.890 0.800 0.644
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0.082 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0.008 0 0 0
0.276 0.138 0.087 0.046 0.051 0.012 0.012 0.026 0.003 0.002 0.030 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0.306 0.305 0.361 0.303 0.303 0.163 0.081 0.158 0.226 0.178 0.141 0.230 0.225 0.144 0.232 0.267 0.106 0.079 0.041 0.147 0.012 0.051 0.035 0.052 0.030 0.034 0.113 0.034 0.045 n.t. 0.037 0.070 0.185 0.046 0.033 0.043 0.015 0.060 0.097 0.066 0 0.103 0.031 0.321 0.108 0.161 0.135 n.t. 0.193 0.718 0.380 0.352 0.247 0.348 0.368 0.684
Immunoglobulin Allotypes as a Marker of Population History in the Southwest Pacific
Table 14.2
Summary Allotype Frequencies from East Asia and the Pacific—cont’d
Language
Location
PAP PAP AN AN AN AN Polynesian Micronesian
Solomons—Reef Islands Solomons—Santa Cruz Vanuatu New Caladonia—Loyalty Isls Fiji—Rotuman Fiji Niue West Truk
N
*A G
*A,X G
*A B
*A,F B
others
12 33 276 379 24 861 215 48
0.542 0.378 0.403 0.642 0.250 0.405 0.478 0.239
0 0.031 0.177 0.114 0 0.153 0.043 0.125
0 0 0.030 0.047 0 0.005 0 0
0.458 0.591 0.390 0.237 0.750 0.437 0.479 0.635
0 0 0 0 0 0 0 0
*A T 0 0 0 0 0 0 0 0
KM*1 0.236 0.478 0.116 0.248 0.323 0.255 0.121 0.196
a
Austronesian Autralian Aboriginal c Trans-New Guinea Language Phylum, but not from the Eastern Highlands d All PNG locations are province names e Papuan language group, but non-Trans New Guinea Language Phylum f Trans-New Guinea Language Phylum from the Eastern Highlands g Wurm (1975) classified these ENGH, but Pawley in chapter 3 classified them as Madang (non-ENGH), and the genetic fit is better with Madang, non-ENGH. b
and among AN speakers? GM*A G and GM*A,X G are also noteworthy because they occur in variable frequencies in all human populations with the exception of those in Sub-Sahara Africa. Gm allotypes first came to significant attention because they suggested a major genetic difference between PAP and AN speakers. Over 1,600 inhabitants of the Morobe Province of Papua New Guinea and a series from Bougainville were tested for G1M A and X and G3M B and KM 1, divided between AN speakers and PAP speakers (Giles et al., 1965). They concluded that “the allele Gmax is almost absent from MN (Austronesian) speakers. …” “It is suggestive that the Gm gene frequencies of the MN-speaking New Guineans are closer to those found in Southeast Asia than are those to the NAN (Papuan) speakers, particularly the Eastern Highland villages. …” “The Gm frequencies imply the existence of two populations with origins separate in time or space. One of the populations, the MN speakers, appears closely related to modern Southeast Asians and not an autochthonous Melanesian differentiation (emphasis ours) ”(Giles et al., 1965: 1160). In the Markham valley and adjacent highlands, highly significant differences separated the Austronesian (AN) and Papuan speakers (PAP), with Gmax virtually absent in the AN speakers (0.001 vs. 0.159) and Gmab was much more prevalent among the AN than the PAP speakers (0.893 vs. 0.447). Subsequent retesting of G3M B positive samples by Steinberg for G1M F and G3M B3 and B4 (Steinberg, 1967) indicated that 63% of the AN speakers, but only 39% of the PAP speakers, were GM A,F B1,3,4 and the rest were GM A B1,3,4 suggesting that AN and PAP speakers could not only be differentiated based on the frequency of G1M X, but also on the distributions of the “Gma,f b1,3,4” (GM*A.F B) and “Gma b1,3,4” (GM*A B) haplotypes. Subsequent studies also demonstrated that “Gma,f b1,3,4”
(GM*A,F B) was the haplotype found in SE Asia while the “Gma b1,3,4” (GM*A B) found in Australia and New Guinea were different from the African haplotypes (see table 14.1). By 1967, there appeared to be a clear genetic basis for linking the AN speakers found in New Guinea and the SW Pacific to extant populations in Southeast Asia. Also, the AN speakers were genetically differentiated from the older PAP-speaking populations of New Guinea and the nearby Solomon Islands. However, these differences did not hold up when PAP and AN speakers on the Island of Bougainville were studied (Friedlaender and Steinberg, 1970), in which no similar differentiation of PAP and AN speakers could be found (GM*A,F B averaged 0.865 among PAP vs. 0.905 among AN populations, and GM*A,X G averaged 0.032 among PAP and 0.028 among AN groups). Further, the PAP haplotype GM*A B was not detected on Bougainville (Friedlaender and Steinberg, 1970). This led to a controversy regarding the distribution of immunoglobulin haplotypes in PAP and AN populations and particularly how general was the distinction that had been found in the Markham Valley. This and other questions concerning the distribution of GM allotypes can now be addressed with more clarity because of the accumulation of data during the past 20 years.
The Southwest Pacific Sample The current state of immunoglobulin allotyping in the Southwest Pacific is summarized here (table 14.2) in an attempt to clarify the differentiation of languages and populations within New Guinea, Island Melanesia, and Remote Oceania. Samples tested by MSS came from a variety of collections provided by S.W. Serjeantson,
211
regional studies and conclusion
R.W. Hornabrook, K.P.W. McAdams, P.B. Booth, J.J. Saave, G.W. Woodfield, D.C. Gajdusek, H. Cleve, F.B. Austin, G.G. Crane, J.D. Stavely, P. Fauran, H. Indrayana, and B. Boetcher. These were augmented with data from the literature including Steinberg (1962), Ropartz et al. (1962), Vos et al. (1963), Flory (1964), Nicholls et al. (1965), Ropartz et al. (1966), Steinberg and Kirk (1970), Friedlaender and Steinberg (1970), Curtain et al. (1971), Schanfield (1971), Curtain et al. (1972), Steinberg et al. (1972a), Steinberg et al. (1972b), Schanfield et al. (1975), Schanfield et al. (1979), Steinberg and Larrick (1981), and Long et al. (1986). Immunoglobulin allotyping was carried out as described in Schanfield and Fudenberg (1975). All samples were tested for G1M A,X, and F, G3M B, G, and KM 1. In addition, many samples were typed for G1M Z, G2M N, G1/3M G5, G3M C3, C5, S, as well as T and A2M 1 and 2. The A2M1 and 2 data are not included in this summary. GM haplotype frequencies were calculated using gene counting or allocation if ambiguous phenotypes are present (Anderson, 1985). KM frequencies were estimated using either gene counting or square root extraction of the KM 1 frequency. To reduce the data down to a manageable number of categories, the allele frequencies for multiple samples were arithmetically averaged as suggested by Nei and Kumar (2000). The data are reported by location and language group. The linguistic data are presented as follows for New Guinea; languages are classified as Papuan (PAP) or Austronesian (AN), with PAP subdivided into the Trans New Guinea Phylum (TNGP) or non-TNGP, with the TNGP further subdivided as East New Guinea Highland (ENGH) or non-ENGH. The language classifications came from Wurm (1975a, 1975b) and the Summer Institute of Linguistics classifications of languages through “Ethnologue” (http://www.ethnologue.com). These may in some cases be at variance with the classification presented in chapter 3. In addition, table 14.2 contains representative data on the immunoglobulin haplotypes in East Asia, Southeast Asia, Insular Southeast Asia, Micronesia, and Polynesia.
Allotype Distributions GM*A G As shown in table 14.2, GM*A G is so broadly distributed throughout the entire region that it is of no use as a distinguishing character.
GM*A,X G GM*A,X G has a very limited distribution in aboriginal Australia, New Guinea, and Island Melanesia. In New
212
Guinea, its distribution is sporadic except among the Eastern New Guinea Highlands languages (ENGH), where its mean is 0.149 with a range of 0.109–0.187. Among other Trans New Guinea phylum languages, its mean is only 0.022 with a range of 0.000–0.079. Among non-TNGP PAP speakers, the mean is also vanishingly low: 0.009, with a range of 0.000–0.026. In Northern Island Melanesian PAP groups, GM*A,X G has an uneven distribution, with a mean of 0.124 and a range of 0.040–0.312. It is particularly high in East New Britain among both PAP and AN speakers. This is an intriguing distribution that connects certain Northern Island Melanesian populations, New Guinea Eastern Highlands, and Australian Aboriginal groups.
GM*A B This allotype is most common in certain areas of New Guinea, but also is found in northern areas of aboriginal Australia (the Kimberley Mountains to Cape York). It is rare elsewhere in the Pacific, as shown in table 14.2, and is therefore of special importance. The northern coastal aboriginal Australian series from the Kimberley Mountains to Cape York all have the GM*A B haplotype in varying frequencies. This makes the northern coast Australian Aborigines markedly different from the central Australian Aborigines. The highest frequencies of GM*A B are associated in general with the non-ENGH TNGP populations, which occupy West Papua, and more coastal areas surrounding the highland populations. Interestingly, although Northern Island Melanesia was occupied very early, with the exception of New Britain, there is little documentation of the presence of GM*A B there. It has also been found in very low frequencies in other sections of Island Melanesia as far as Fiji.
GM*A,F B In tables 14.1 and 14.2 it is clear that the GM*A,F B haplotype in East and Southeast Asia has its highest concentration in Thailand. Further, it is present in all AN-speaking populations from the aboriginal populations of Taiwan to Madagascar (Weber et al., 2000) and into Remote Oceania. As mentioned above, the association of this haplotype with AN speakers, and its differentiation from PAP speakers, was initially suggested by Giles et al. (1965) and confirmed by Steinberg, (1967) when the previously tested G3M B positive samples were tested for G1M F. There can be no question that GM*A,F B evolved in southeast Asia and spread both northward in Asia and moved out of southeast Asia into insular southeast Asia and into the Pacific. The distribution of AN languages and GM haplotypes in New Guinea and Northern Melanesia suggests that populations that originated in
Immunoglobulin Allotypes as a Marker of Population History in the Southwest Pacific
Southeast Asia came along the northern coast of New Guinea, settling in sporadic locations until reaching Madang and Morobe provinces. In Madang, there is a considerable presence of GM*A,F B in both PAP and AN groups, while in Morobe, there is the distinction following linguistic lines found by Giles et al. (1965). The high frequency of this allotype continues around the coast of New Guinea as far as the Central Province AN speakers, and in Northern Island Melanesia is very high in both AN and PAP groups. This is particularly striking in Bougainville. This allotype continues in high frequency out into Remote Oceania.
GM*AT and Rare Variants in the Region As shown in tables 14.1 and 14.2 GM*A T only occurs in Northern Asian populations or in populations that have had contact with them. The appearance of GM*A T in Thailand and Indonesia is associated with the presence of Chinese traders. Undefined variant haplotypes were reported in the ENGH samples from Madang Province (Long et al., 1986).
KM*1 This allotype variant on the light chain is generally not distinctive, but it does show a particularly high frequency in central/north Bougainville (the Aita and Rotokas), and has a steep cline through southern Bougainville (Friedlaender and Steinberg, 1970).
Relevance for Population Relationships and Past Migrations It appears based on the data in table 14.2 that the initial migrants from Southeast Asia to Sahul may have only had GM*A G and GM*A,X G, like the central desert Australian Aborigines of today. If they had GM*A B, they had to have lost it through drift in central Australia. The discontinuous distribution of these allotypes suggests that the highland New Guinea and central Australian aboriginal populations are refuge or relic populations of the earlier inhabitants of New Guinea and Australia. Are there other markers that support this linkage between central Australia and highlands New Guinea? Roberts-Thompson et al. (1996) demonstrated extensive sharing of α-globin haplotypes between central Australian Aborigines and highland populations of New Guinea. Additional support for a common origin of Australian Aborigines and the people of New Guinea is the widespread distribution of GC*Aborigine, or more accurately GC*1A1, in New Guinea and Australia (See Kirk et al., 1963; Kitchin et al., 1972; Schanfield et al., 1975).
Interestingly, GC*1A1 is also found in Sub-Saharan Africans, though usually at lower frequencies than found in New Guinea. Studies by Kofler et al. (1995) indicate that the GC*1A1 allele found in Australian Aborigines and that found in South Africa are the same mutation, suggesting a very old origin of this particular mutation in Africa that was carried to Oceania. In any case, the data from α-globin and GC*1A1 also support a common origin of ancient Australian Aborigines and Papuans. As noted originally by Giles et al. (1965), there is a marked difference in the frequency of GM*A,X G between PAP and AN speakers in the Markham Valley in Morobe Province. This held up for all AN populations in New Guinea (mean 0.005, range 0.000–0.017). However, the majority of Papuan-speaking populations that AN speakers interacted with (non-ENGH TNGP or non-TNGP) also had little or no GM*A,X G (0.024, range 0.000–0.079). In contrast, the GM*A,X G haplotype is found sporadically in Northern Melanesia in both AN speakers (mean.058, range 0.005–0.234) and PAP speakers (mean 0.124, range 0.040–0.312). The GM*A B haplotype appears to have a reverse distribution. It is found in all ANspeaking populations on New Guinea (mean 0.252, range 0.109–0.489), but with the exception of New Britain, it is extremely rare elsewhere in Northern Melanesia in AN (mean 0.008, range 0.000–0.045) or PAP (mean 0.006, range 0.000–0.023) speakers. This suggests that the AN populations in New Guinea acquired GM*A B from the PAP inhabitants, and some PAP speakers in Northern Island Melanesia acquired GM*A,X G early on, and that the original AN speakers had neither of these haplotypes, very much like some populations from Southeast Asia and aboriginal Taiwan. The data comparing AN populations to adjacent PAP populations are summarized in Table 14.3. In New Guinea, on the average, the frequency of GM*A,F B is higher in AN populations (mean 0.540, range 0.250–0.735) than in the adjacent PAP populations (mean 0.214, range 0.000–0.528), with the exception of the Madang Province, where frequencies of GM haplotypes in PAP TNGP nonENGH and AN speakers are virtually identical. In Northern Melanesia there is also little differentiation. These findings suggest that there was little back and forth movement of AN or PAP populations between New Guinea and Northern Melanesia after the initial migration. The distribution of immunoglobulin haplotypes in the Central and Milne Bay districts is not consistent with back migration of AN speakers from Northern Melanesia, as virtually all of the populations have detectable GM*A B (table 14.2). Further, it suggests a prolonged period of contact and gene flow between local AN and PAP populations there. This is most marked in Madang Province which appears to have the highest concentration of GM*A,F B and little differentiation between the AN and PAP populations,
213
regional studies and conclusion
Table 14.3 Summary of Differences in Immunoglobulin Allotype Distributions between PAP and AN Speakers in Areas of New Guinea (the mean allele frequencies and the range [Low vs High] are presented) GM haplotypes Location
Language
N
Non-AN NG
Non-ENGH
Highland NG
ENGH
Madang Province Madang Province
Non-ENGHa Non-ENGH Non-TNGP AN Non-ENGH Non-TNGP
2256 low high 1317 low high 492 679
Madang Province AN NG—Madang
AN NG—Madang
AN
Northern Melanesia
PAP
Northern Melanesia
AN
a
*A G
*A,X G
*A B
*A,F B
OTHER
KM*1
0.227 0.152 0.301 0.674 0.415 0.786 0.415 0.169
0.009 0.006 0.013 0.146 0.109 0.187 0.045 0.011
0.760 0.692 0.834 0.163 0.076 0.212 0.168 0.063
0.001 0.001 0.003 0.017 0.000 0.060 0.289 0.756
0.000
0.082 0.000
0.059 0.030 0.106 0.049 0.015 0.113 0.037 0.114
377 2696
0.109 0.540
0.008 0.023
0.127 0.211
0.756 0.223
0.000 0.000
0.046 0.061
low high 1163 low high 2362 low high 2113 low high
0.257 0.805 0.227 0.016 0.600 0.162 0.047 0.264 0.145 0.066 0.324
0.000 0.079 0.004 0.000 0.017 0.124 0.040 0.312 0.058 0.005 0.234
0.026 0.403 0.283 0.109 0.489 0.006 0.000 0.023 0.008 0.000 0.045
0.013 0.623 0.486 0.250 0.735 0.708 0.428 0.908 0.788 0.554 0.905
0.000
0.000
0.000
0.001
0.000 0.185 0.134 0.051 0.321 0.400 0.135 0.718 0.235 0.193 0.352
Wurm (1975) classified these as ENGH, but Pawley in chapter 3 has classified them as Madang (non-ENGH).
although GM*A,F B appears to be overlaid on the basic patterns seen in the non-TNGP, and non-ENGH TNGP. Using the GM haplotypes, it is not clear whether the AN speakers in this region are Austronesianized Papuans, or Austronesians with Papuan gene flow. The East and West Sepik Provinces, home to the highest concentration of non-TNGP PAP languages have a markedly different distribution of GM haplotypes than the TNGP languages in most of New Guinea (table 14.2) as they are characterized by high frequencies of GM*A G, little or no GM*A,X G and moderate GM*A B, and variable GM*A,F. To summarize the GM relationships, (1) because of GM*A B, there is a connection between northern Australian Aborigines and southern Papua New Guinea PAP speakers; (2) there is a clear relationship between Australian Aborigines and scattered New Guinea Papuanspeaking groups, particularly in the Eastern Highlands, because of GM*A,X G frequencies, and this extends as far as East New Britain; (3) the distribution of GM*A,F B appears to be a clear signal of recent Southeast Asian/ Austronesian influence. These GM data are consistent with the mtDNA results in particular presented in chapter 4 and elsewhere
214
(Friedlaender et al., 2005; Merriwether et al., 2005). To simplify the mtDNA data presentation, mtDNA haplogroups P and Q appear to have originated in New Guinea, and that only a few old branches of P are shared with Australian Aborigines, indicative of a long isolation. Further the Q haplogroups and M27, M28, and M29 are derived from the M haplogroups which also have different forms within Melanesia but show a lack of, or very distant, connections to aboriginal Australia. The final player of interest is the B haplogroup. The presence of the B haplogroup in New Guinea is surely a marker of the relatively recent AN incursion into New Guinea (see chapter 2), similar to GM*A,F B. Coastal New Guinea populations have a mean frequency of the Southeast Asian B haplogroup of 0.390 (range 0.286–0.556) and it was not found in the highlands of New Guinea except for the admixed highland population from the Morobe Province (0.350) which are close to the mean for coastal populations. In Northern Island Melanesia, the mtDNA data suggest considerable variation in the frequency of the Southeast Asian marker haplogroup B, with AN speakers having a mean haplogroup B frequency of 0.570 (range 0.155–0.800) with PAP speakers having a slightly lower mean (range 0.000–0.952). The distribution
Immunoglobulin Allotypes as a Marker of Population History in the Southwest Pacific
of P, Q, and M are also of considerable interest in these populations, with P and Q serving as markers of Papuan maternal lineages, and M representing a localized differentiation in Northern Melanesia. The presence of P and Q in both PAP and AN populations along with B suggests that virtually all Northern Island Melanesian populations are a mixture of Papuan and Austronesian maternal lineages. Similarly, the presence of GM*A,F B in all Northern Melanesians regardless of linguistic affiliation further supports the presence of AN gene flow, with a very few exceptions. There is the suggestion that Eastern New Britain is a focus of the early Pleistocene immigrants to Northern Melanesia from Sahul. The highest frequency of the locally evolved haplogroup M is also in the PAP and AN speakers in eastern New Britain, suggesting that these haplogroups evolved there (Friedlaender et al., 2005; chapter 4). Thus, the explanation for the apparent contradiction seen between the original observation of Giles et al. (1965) and Friedlaender and Steinberg (1970) rests on the nature of the PAP populations in Northern Island Melanesia, which appear to have been largely first migration protoPapuans who had GM*A G, GM*A,X G with little or no GM*A B and mtDNA haplogroups P, Q, and M, who were relatively isolated from New Guinea. They later mixed with AN speakers bringing GM*A G, GM*A,F B, and haplogroup B. Further, there appears to have been a long period of interaction in Northern Island Melanesia between the earlier PAP individuals and the more recent AN individuals since the settlers who went out of Northern Island Melanesia to populate remote Oceania took P, Q, and M, along with B and GM*A,F B with them to remote Oceania. A data-mining tool used to try to reduce large amounts of information to a smaller number of uncorrelated parameters is Principal Components Analysis (PCA). PCA was performed on a restricted set of populations (31) listed in Table 14.4, using Statistica (7th version, Statsoft) reducing the four GM haplotypes and KM*1 to two significant factor scores. Factor 1, which accounted for 37.2% of the variation, was negatively loaded on GM*A G and positively loaded on GM*A,F B, while Factor 2 accounted for 25.4% of the variation and negatively loaded on GM*A,X G and positively loaded on GM*A B. KM*1 was not heavily loaded in either factor but has a noticeable effect in the two-dimensional plot. Figure 14.1 is the graphic representation of the plot of the factor scores of the 31 samples. Are there any consistent findings? On the left side of figure 14.1 from top to bottom are a series of populations that reflect the peopling of New Guinea before the arrival of the AN populations. The upper left quadrant is characterized by high frequencies of GM*A G and GM*A,X G, while the bottom right quadrant is characterized by GM*A B. From top to
Table 14.4 The List of Population Samples Used in the Principal Components Analysis Shown in Figure 14.1 Population Language 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31
AN AUST AUST PAP/AN? Non-ENGH Non-TNGP AN AN Non-TNGP Non-ENGH ENGH ENGH Non-ENGH Non-ENGH AN AN AN AN PAP AN AN PAP AN AN PAP AN AN AN AN Polynesian Micronesian
Geographic region Indonesia Australia—Central and Western Desert Australia—North Coast Irian Jaya—Northeast Coast PNG—Sepik Province PNG—Sepik Province PNG—West Sepik Province PNG—Madang Province PNG—Madang Province PNG—Madang Province PNG—Madang Province All but Madang AN areas, except Madang Non-AN areas PNG—Morobe Province PNG—Milne Bay Province PNG—Central Province New Britain—Western New Britain—Eastern New Britain—Eastern New Ireland Bougainville Bougainville—Southern Solomons Solomons Santa Cruz and Reef Islands Vanuatu New Caledonia—Loyalty Islands Fiji Niue West Truk
bottom, group 2 is the central Australian aborigines, group 3 is the northern coast Australian Aborigines, group 12 is the ENGH speakers, excluding Madang, group 5 is the non-ENGH speakers from Sepik, group 6 is the nonTNGP speakers from the Sepik and 14 is the non-ENGH speakers from western New Guinea. This transition goes from the earliest extant representatives of the earliest inhabitants, the central Australian Aborigines, through various mixtures to the second migration to Sahul with the western New Guinea non-ENGH speakers. The latter is characterized by high GM*A B frequencies and low GM*A G and GM*A,X G. The far left side of figure 14.1 represents the populations with the lowest contact with AN speakers. In contrast on the far right in the upper left, populations 22, 23, 24, and 25 form a cluster. These are the PAP and AN populations from Bougainville and the Solomon Islands. These populations are all characterized by high frequencies of GM*A,F B and KM*1. The middle cluster represents the intermediate populations divided
215
regional studies and conclusion
Figure 14.1 A two-dimensional plot of factor 1 (37.2% of variance) and factor 2 (25.4%) obtained by principal components analysis of 31 samples tested for four GM haplotypes and KM*1. The identification of the samples is found in table 14.4. Squares are Australian; triangle is Indonesian; diamonds are New Guinean (empty are AN); filled circles are Near Oceanic other than New Guinean; empty circles are Remote Oceanic.
based on the frequency of GM*A,F B. The southeast quadrant is largely characterized by samples with high GM*A,F B frequencies, as the samples move to the upper left quadrant the frequencies of GM*A G and GM*A,X G increase. Thus, the PCA analysis reflects less geography and language than the overall distribution of immunoglobulin allotypes, though there is some clustering of remote Oceania and the upper left quadrant. The immunoglobulin allotypes, particularly the GM/IGH haplotypes, are therefore clearly good markers of the migrations of people out of Southeast Asia over the late Pleistocene and Holocene, and are consistent with many unusual aspects of the mtDNA pattern of variation in the region. The KM marker appears to have regional variation such that Australia, New Guinea, Northern Melanesia, and remote Oceania have slightly different distributions of KM*1, with a maxima in Bougainville (table 14.2).
216
Acknowledgments I would like to thank Robert Baylarian, Milwaukee Blood Center, Milwaukee, WI, Susan Schoeppner and Rebecca Brown, American Red Cross Blood Services, Washington DC for technical support. Besides my co-authors, those who provided samples and data on populations include Gregg Crane, J. D. Stavely, Pierre Fauran, Haroon H Indrayana (deceased), Richard H Ward (deceased) and Barry Boetcher.
References Anderson L. 1985. The estimation of blood group gene frequencies: a note on the allocation methods. Animal Blood Groups and Biochemical Genetics 16: 1–7. Bowern C, Koch H. 2004. Australian languages: classification and comparative method. Amsterdam: John Benjamins.
Immunoglobulin Allotypes as a Marker of Population History in the Southwest Pacific
Curtain CC, van Loghem E, Baumgarten A, Golab T, Gorman J, Rutgers CF, Kidson C. 1971. The ethnological significance of the Gamma-globulin (Gm) factors in Melanesia. American Journal of Physical Anthropology 34:257–72. Curtain CC, van Loghem E, Fudenberg HH, Tindale NB, Simmons RT, Doherty RL, Vos G. 1972. Distribution of the immunoglobulin markers at the IgG1, IgG2, IgG3, IgA2 and κ- chain loci in Australian Aborigines: Comparison with New Guinea populations. American Journal of Human Genetics 24:145–55 Dard P, Lefranc M-P, Osipova L, Sanchez-Mazes A. 2001. DNA sequence variability of IGHG3 alleles associated to the main G3m haplotypes in human populations. European Journal of Human Genetics 9: 765–72. Flory L. 1964. Serum factors of Australian Aborigines from North Queensland. Nature 201: 508–9. Friedlaender JS, Steinberg AG. 1970. Anthropological significance of gamma globulin (Gm and Inv) antigens in Bougainville Island, Melanesia. Nature 28: 59–61. Friedlaender JS, Schurr T, Gentz, F, Koki G, Friedlaender FR, Horvat G, Babb P, Cerchio S, Kaestle F, Schanfield M, Deka R, Yanagihara R, Merriwether DA. 2005. Expanding Southwest Pacific mitochondrial haplotypes P and Q. Molecular Biology and Evolution 22: 1506–17. Giles E, Ogan E, Steinberg AG. 1965. Gamma-globulin factors (Gm and Inv) in New Guinea: Anthropological significance. Science 150: 1158–60. Kirk RL, Cleve H, Bearn AG. 1963. The distribution of the Gc-Types in sera from Australian aborigines. American Journal of Physical Anthropology 21: 215–23. Kitchin DF, Bearn AG, Alpers M, Gajdusek DC.1972. Genetic studies in relation to Kuru. III. distribution of the inherited serum Group-specific protein (GC) phenotypes in New Guineans: An association of Kuru and the Gc Ab phenotype. American Journal of Human Genetics 24: S72–S85. Kofler A, Braun A, Jenkins T, Serjeantson SW, Cleve H. 1995. Characterization of mutants of the vitamin-Dbinding protein/group specific component: GC aborigine (1A1) from Australian aborigines and South African blacks and 2A9 from south Germany. Vox Sanguines 68: 50–4. Long JC, Naidu JM, Mohrenweiser HW, Gershowitz, H, Johnson PL, Wood, J, Smouse PE. 1986. Genetic characterization of Gainj- and Kalam-speaking peoples of Papua New Guinea. American Journal of Physical Anthropology 70: 75–96. Merriwether DA, Hodgson JA, Friedlaender FR, Allaby R, Cerchio S, Koki G, Friedlaender JS. 2005. Ancient mitochondrial M haplogroups identified in the Southwest Pacific. Proceedings of the National Academy of Sciences, USA 102: 13034–9. Nei M, Kumar S. 2000. Molecular Evolution and Phylogenetics. Oxford, UK: Oxford University Press, pp 333.
Nicholls EM, Lewis HBM, Cooper DW, Bennett JH. 1965. Blood groups and serum protein differences in some Central Australian Aborigines. American Journal of Human Genetics 17: 293–307. Roberts-Thomson JM, Martinson JJ, Norwich JT, Harding RM, Clegg JB, Boettcher B. 1996. An ancient common origin of Aboriginal Australians and New Guinea Highlanders is supported by α-Globin haplotype analysis. American Journal of Human Genetics 58: 1017–24. Ropartz C, Rousseau P-Y, Rivat L, Kirk RL.1962. Frequence du facteur Inv(a) chez les Aborigines de l’Ouest Australien. Nouvelle Revue Francaise d’Hemotologie 2: 86–90. Ropartz C, Gold ER, Rivat L, Rousseau P-Y.1966. Frequence du facteur Gm(4) parmi quelques population blanche, noire et jaunes. Transfusion (Paris) 9: 293–301. Ropartz C, Schanfield MS, Steinberg AG. 1976. Review of the notation for the allotypic determinants and related markers of human immunoglobulins, WHO meeting on human immunoglobulin allotypic markers held 16-19 July 1974, Rouen, France. Report amended June 1976. Journal of Immunogenetics 3: 357–62. Schanfield MS. 1971. Population Studies on the Gm and Inv Antigens in Asia and Oceania. Ann Arbor: University Microfilms. Schanfield MS. 1977. Population affinities of the Australian aborigines as reflected by the immunoglobulin allotypes. Journal of Human Evolution 6: 341–52. Schanfield MS. 1980. The anthropological usefulness of highly polymorphic systems: HLA and immunoglobulin allotypes. In: Crawford MH, Mielke J, editors. Current Developments in Anthropological Genetics. New York: Plenum, pp 65–86. Schanfield MS, Fudenberg HH. 1975. The anthropological usefulness of the IgA allotypic markers. In: Watts EW, Johnston FE, Lasker GW, editors. Biosocial Interrelations in Population Adaption. The Hague, The Netherlands: Mouton, pp 105–14. Schanfield MS, van Loghem E. 1986. Human immunoglobulin allotypes. In: Weir DM, Herzenberg LA, Blackwell CC, Herzenberg LA, editors. Handbook of Experimental Immunology, 4th edn. Edinburgh, UK: Blackwell Scientific, ch 94, 94.1–94.18. Schanfield MS, Giles E, Gershowitz H. 1975. Genetic studies in the Markham Valley, Northeastern Papua New Guinea: Gamma globulin (Gm and Inv), Group Specific Component (Gc) and Ceruloplasmin (Cp) typing. American Journal of Physical Anthropology 42: 1–8. Schanfield MS, Wells JV, Fudenberg HH. 1979. Immunoglobulin allotypes and response to tetanus toxoid in Papua New Guinea. Journal of Immunogenetics 6: 311–15. Schanfield MS, Atmodirono H, Indrayana NS, Soekry EK, Kartika P. 1995. Identification of a torso found in Bali using RFLP, PCR and non-DNA markers. 5th IndoPacific Congress on Legal Medicine and Forensic
217
regional studies and conclusion
Science. 16–20 July, 1995, Bali, Indonesia. Program, p 50 (abst). Shows TB, McAlpine PJ, Bouchix C, et al. 1987. Guidelines for human gene nomenclature. An international system for human gene nomenclature (ISGN, 1987). Cytogenetics and Cell Genetics 46: 11–28. Steinberg AG. 1962. Progress in the study of genetically determined human gamma globulin types (Gm and Inv groups). Progress in Medical Genetics 2: 1–33. Steinberg AG. 1967. Genetic variations in human immunoglobulins: the Gm and Inv types. In: Greenwalt TJ (ed) Advances in Immunogenetics. Philadelphia: JB Lippincott, pp 75–98. Steinberg AG, Kirk R. 1970. Gm and Inv types of Aborigines in the Northern Territory of Australia. Archeology & Physical Anthropology in Oceania 5: 163–72. Steinberg AG, Larrick JW. 1981. Gm and Inv (Km) Studies of Melanesian people on the Huon Penninsula in Northeast Papua New Guinea: Polymorphism for a Gm1,5,10,11,13,14,17,21,26 haplotype. American Journal of Physical Anthropology 55: 89–94. Steinberg AG, Damon A, Bloom J. 1972a. Gammaglobulin allotypes in Malanesians from Malaita nd Bougainville,
218
Solomon Islands. American Journal of Physical Anthropology 36: 77–4. Steinberg AG, Gajdusek DC, Alpers M. 1972b.Genetic studies in relation to Kuru. V. Distribution of human Gamma Globulin allotypes in New Guinea populations. American Journal of Human Genetics 24, supplement: S95–S110. Vos GH, Kirk RL, Steinberg AG. 1963. The distribution of gamma globulin types Gm(a), Gm(b) , Gm(x) and Gmlike in South and Southeast Asia and Australia. American Journal of Human Genetics 15: 44–52. Weber W, Nash DJ, Giblett E, Motulsky AG, Henneberg M, Crawford MH, Emejuaiwe SO, Martin SK, Goldsmid JM, Spedini G, Glidewell S, Schanfield MS. 2000. Phylogenetic relationships of human populations in sub-Saharan Africa. Human Biology 72: 753–772 Wurm SA. 1975a. New Guinea area Languages and Language Study Vol. 1. Papuan Languages and the New Guinea Linguistic Scene. Pacific Linguistics Series C-No 38. Australian National University. Wurm SA. 1975b. New Guinea area Languages and Language Study Vol. 2. Austronesian Languages. Pacific Linguistics Series C-No 39. Australian National University.
15 Contributions of Population Origins and Gene Flow to the Diversity of Neutral and Malaria Selected Autosomal Genetic Loci of Pacific Island Populations J. Koji Lum
Introduction The ability of a population to adapt to its environment and respond effectively to its pathogens depends on its genetic diversity. The amount of genetic diversity of a population reflects its origin, extent of gene flow with neighboring populations, population size over time, and exposure to selective forces. Here I illustrate the contribution of these factors to the maintenance of genetic diversity by summarizing three of our recent studies of neutral short tandem repeat (STR) loci collected from populations throughout the Pacific and genetic markers that confer resistance and susceptibility to malaria from Papua New Guinea (PNG) and Vanuatu. These studies illustrate the contribution of evolutionary studies to the understanding of infectious disease epidemiology within Melanesia. The initial settlement of the Pacific began in the Pleistocene and resulted in the peopling of New Guinea and the large, intervisible islands of the western Pacific as far east as the central Solomon Islands by 35 thousand years before present (kybp)(Groube et al. 1986; Wickler and Spriggs, 1988). This region of early settlement is coincident with the current limit of the Papuan languages and is known as Near Oceania (Pawley and Green, 1973; Green, 1991, 1999). The second major human range expansion into the Pacific began approximately 3.5 kybp and resulted
in the rapid settlement beyond the Solomon Islands, through Vanuatu, New Caledonia, Fiji, out to Tonga and Samoa by 3.1 kybp, and eventually all of Polynesia and Micronesia by 600 ybp (Kirch, 2000). All of these recently settled archipelagos are currently inhabited exclusively by Austronesian-speakers and are collectively known as Remote Oceania (Green, 1999). The region known as Melanesia contains all of Near Oceania and parts of Remote Oceania (Vanuatu, New Caledonia, and Fiji) and thus, populations derived from both major linguistic groups and temporally distinct migrations into the Pacific. The extent and pattern of genetic exchange among Papuanand Austronesian-speaking populations of Melanesia during the past 3,500 years remains contentious (Terrell et al., 2001). The coincident distribution of Pleistocene settlement sites and Papuan-speakers on the one hand and the rapid Holocene expansion into Remote Oceania on the other led Diamond (1988) to propose the “Express Train to Polynesia.” This model implied that the push through Near Oceania during the settlement of Remote Oceania was so rapid that there was minimal genetic exchange between the long-term Papuan-speaking residents of Near Oceania and the intrusive Austronesian speakers arriving from Island Southeast Asia. Current Austronesian-speaking populations of Island Melanesia and New Guinea generally show biological affinities to Papuan speakers to the
219
regional studies and conclusion
exclusion of Austronesian speakers from Polynesia and Micronesia (Pietrusewsky, 1990; Hanihara, 1993; RobertsThomson et al., 1996; Lum et al., 2002), indicating that Pacific prehistory has been more complex than proposed by Diamond (1988). More recently, Green (1991) proposed the Triple “I” model, arguing that the intrusive Austronesian speakers from Island Southeast Asia that entered the Pacific 3.5 kybp encountered the long-resident Papuan speakers, integrated aspects of their culture, in particular tree crops and root cultivars to create an innovative cultural complex now known as Lapita that formed the basal strata of Remote Oceania. In this model, present Pacific Island societies are to varying degrees composites of the two linguistically, culturally, and genetically distinct ancestral populations. Furthermore, during the past 3,500 years Pacific Island populations have continued to exchange genes and ideas. The extent to which this post-settlement gene flow has occurred determines the correlation between linguistic and genetic patterns among populations and the amount of genetic diversity each population has accumulated. In the following three sections I will summarize studies of both neutral and malaria selected autosomal loci from Pacific Island populations in order to gain insight into the extent of gene flow among populations and further, the consequence of this gene flow on genetic diversity and malaria resistance and susceptibility.
Neutral Genetic Patterns In 2002 we (Lum et al.) published a study examining the patterns of relationships among 27 Pacific Island and
Asian populations based on analyses of 13 unlinked autosomal STR loci from 965 individuals (figure 15.1). Like mitochondrial DNA (mtDNA) and Y chromosome polymorphisms discussed in previous chapters, STR variation is generally thought to be neutral and so unaffected by selection (but see Costa Lima and Pimentel, 2004). Unlike mtDNA and Y chromosome polymorphisms, the STR loci discussed here are inherited from both parents. STR loci are regions of the genome where each individual inherits two copies of simple, defined sequences repeated a variable numbers of times. For example, at a GAT trinucleotide STR locus an individual might have inherited three copies of the repeat (GATGATGAT) from one parent and two copies (GATGAT) from the other. Thus, for each STR locus each person may have up to two length polymorphisms. By examining allele length polymorphisms at 13 STR loci, frequency distributions characterizing each population were compiled from which genetic distances among populations were estimated. These genetic distances were further compared to linguistic and geographic distances among populations to infer the effects of settlement bottlenecks and post-settlement gene flow on the genetic diversity of present populations. We examined the amount of STR variation within each population as a function of its distance from Island Southeast Asia, the ultimate origin of all Pacific populations. In general, the highest levels of genetic diversity were found in the Western Pacific and the least in the peripheries of Remote Oceania. This resulted in a significant correlation between the increasing losses of genetic diversity and increasing straight-line distances from Island Southeast Asia (r = 0.48, p < 0.013) (figure 15.2a). To further explore the relationship between maintained
Figure 15.1 Map showing the geographic origins of the populations examined for STR diversity. The 27 populations are grouped by region: triangles, Islands of Southeast Asia; filled circles, Western Micronesia; open circles, Central–Eastern Micronesia; filled squares, Melanesia; open squares, Polynesia.
220
Contributions of Population Origins and Gene Flow
Figure 15.2 Correlations between loss of genetic diversity and distance from Southeast Asia under three scenarios: (a) straight line distances, (b) migration distances incorporating linguistic and archeological information, (c) migration distances as in (b), but with Melanesian populations excluded.
genetic diversity and putative distance traveled from the ancestral homeland, we also estimated migration distances incorporating information from settlement patterns and linguistic relationships. For example, linguistic, ethnographic, and archeological information indicate that Kapingamarangi, a Polynesian outlier within Micronesia was settled from Western Polynesia (Pawley and Green, 1973; Pawley and Ross, 1993; Spriggs and Anderson, 1993; Intoh, 1997). Likewise, linguistic comparisons suggest that Central–Eastern Micronesia was settled from east to west (Bender, 1971; Jackson, 1986) from a source in the southern Solomon Islands or Northern Vanuatu. By using these migration distances based on more realistic models of settlement rather than simple straight line distances, the correlation with loss of genetic diversity
increased (r = 0.78, p < 0.001) (figure 15.2b), indicating that incomplete sampling of genetic diversity by successive colonizing populations island hopping through the Pacific has likely resulted in the lack of diversity we observe in peripheral Remote Oceanic populations. One exception to the general pattern of loss of genetic diversity during migration into the Pacific is found in the Island Melanesian archipelago of Vanuatu. This archipelago was settled approximately 3.2 kybp during the initial rapid expansion of the Lapita cultural complex. Although all of the languages of Vanuatu are Austronesian, as noted above, Ni Vanuatu (the indigenous peoples of Vanuatu) cluster with Papuan-speakers of Near Oceania to the exclusion of Austronesian-speaking populations from Polynesia and Micronesia in analyses of morphological traits (Pietrusewsky, 1990; Hanihara, 1993), suggesting extensive post-settlement gene flow throughout the region now known as Island Melanesia. The four populations from Vanuatu we examined (Santo, Maewo, Tanna, and Banks) harbored relatively high levels of genetic diversity, consistent with substantial contributions from both of the ancestral gene pools of the Pacific. By excluding New Guinea and Island Melanesian populations from the comparisons the correlation between migration distance and loss of genetic diversity increased (r = 0.83, p < 0.001) (figure 15.2c). Furthermore, these populations clustered in multidimensional scaling analysis with both Austronesian (Trobriands) and Papuan speakers (Asaro Valley) from PNG to the exclusion of Fijians and other Austronesian speakers from Polynesia and Micronesia (figure 15.3). We also estimated pairwise FST values among populations. FST is a genetic statistic that ranges from 0 to 1 with “0” indicating two populations have identical allele frequencies and “1” indicating that they do not share any alleles. Thus, pairwise FST values are inversely correlated with the inferred level of gene flow between populations. After estimating FST values between populations we employed the permutation procedure of Arlequin version 2.000 (Schneider et al., 2000) to generate approximately 10,000 data sets by randomly exchanging individuals between populations and calculating FST values for each permutation. Comparisons of the observed FST values to those of the permuted data sets allowed us to determine which pairs of populations were significantly distinct and which were not. Figure 15.4 summarizes the results of the simulated gene flow analyses. One interesting feature of figure 15.4 is that Central–Eastern Micronesian populations are not significantly distinct from the Polynesian population of Samoa. This is in contrast to comparative linguistic studies that predict that Central–Eastern Micronesians should link with populations in Vanuatu that speak relatively similar languages (Jackson, 1986). Rather than indicate settlement of Micronesia from
221
regional studies and conclusion
Figure 15.3 Non-metric multidimensional scaled plot of the FST genetics distances between populations. Symbols for regions are as in figure 15.1. The language spoken by Melanesian populations is indicated by: A, Austronesian and P, Papuan. The proportion of the total variance summarized in the first two dimensions shown is 0.75.
Polynesia, the high genetic diversities in Vanuatu relative to the low genetic diversities of Micronesia suggested to us that the current patterns of relationships reflect extensive post-settlement gene flow within Island Melanesia. In this model Micronesian populations show primary
relationships with Polynesians rather than Ni Vanuatu because the latter currently contains substantial frequencies of alleles acquired from Near Oceania since protoMicronesians and proto-Polynesians branched off and migrated north and east, respectively.
Figure 15.4 Map showing the 15 pairs of populations that are not significantly different at the 0.05 level. Levels of gene flow between populations inferred from the proportion of permuted data sets yielding higher FST values than those actually observed divided into two classes: moderate (0.05 ≤ dashed lines < 0.10) and high (solid lines ≥ 0.10).
222
Contributions of Population Origins and Gene Flow
Our analyses of both genetic diversity and patterns of inferred gene flow support substantial genetic exchange among populations within the area now known as Island Melanesia leading to a break down in correlations among genetic and linguistic patterns. We (Cann and Lum, 2004) have argued that this post-settlement gene flow within Melanesia has resulted in a series of boundaries extending east through the region (figure 15.5). The first boundary at the southern Solomon Islands is the limit of both Pleistocene settlement and Papuan languages that defines Near Oceania; the next at Buxton’s Line in Southern Vanuatu is the limit of anopheles mosquitoes, the vector of malaria in the Pacific; and the third at Fiji is the limit of the Island Melanesian morphology. In this model, post-settlement gene flow from Near Oceania at any given point in time was not of sufficient scale to introduce Papuan languages into Remote Oceania, but apparently enough canoes from Near Oceania reached Vanuatu to establish anopheles mosquitoes and malaria, and accumulated gene flow extended the Island Melanesian phenotype as far as Fiji. The overall pattern of anopheles species diversity and Plasmodium falciparum
genetic diversity within the Pacific and the correlation between patterns of human and malaria parasite genetic affinities within Vanuatu is consistent with this model (Lum et al., 2004). Post-settlement gene flow within the Pacific over the past 3,000 years may have thus had substantial impacts on the pattern of genetic diversity of not only human populations, but also their vectors and pathogens. The next two sections will discuss the effects of selection on the distribution of genes that confer resistance and/or susceptibility to malaria with inferred geographic origins from populations of East Sepik Province, Papua New Guinea (PNG), and within the Vanuatu archipelago.
Malaria Selection within PNG (Near Oceania) Since 2001 we have been conducting a multidisciplinary study examining various aspects of malaria epidemiology, drug efficacy, parasite drug resistance, and population genetics in East Sepik Province, PNG (Masta et al., 2003; Hombhanje et al., 2005; Tsukahara et al., 2006).
Figure 15.5 Three boundaries of Melanesia reflecting the limit of Pleistocene settlement and Papuan languages (thick line), the limit of anopheles mosquitoes and malaria (dashed line), and the Melanesian morphology (thin line). The latter two may result from a decreasing gradient of cumulative gene flow from Near Oceania.
223
regional studies and conclusion
One goal of these projects is to evaluate the exchange of genes among Papuan- and Austronesian-speaking populations at a local scale approximating the restricted range of movement and interactions that characterized preEuropean contact populations of PNG (Leahy, 1991). Each of our 13 study populations is within about 15–20 km of another population and they collectively extend from offshore islands, across the coastal plain, over the Prince Alexander Ridge, and into the Sepik Plain. These populations speak four distinct languages; three Papuan (Boiken, Ambalam, and Arapesh) and one Austronesian (Kairiru). Of these, the Boiken-speakers are thought, based on ethnographic records of oral traditions, to have expanded from the Sepik Plain out to the off-shore islands displacing other Papuan- and Austronesianspeaking populations in the relatively recent past (Miki, 1980; Roscoe, 1989). We wondered to what extent the distribution of alleles of these populations would reflect the distinct origins of Papuan and Austronesian speakers, recent gene flow across linguistic barriers, and/or the recent Boiken diaspora. Here I summarize frequency data from a malaria-resistant allele, the erythrocyte band
3 gene 27 base pair deletion (B3∆27) and a neutral, mtDNA marker, the region V deletion (Cann and Wilson, 1983) from 800 individuals representing eight of our 13 study populations (Tsukahara et al., 2006) (figure 15.6). The B3∆27 results in an increased rigidity of the red cell cytoskeleton resulting in an oval-shaped erythrocyte known as ovalocytosis (Jarolim et al., 1991). All B3∆27 individuals described to date are heterozygotes, implying homozygote lethality. Although individuals with the B3∆27 do not seem to be protected from uncomplicated malaria or high densities of infection (Allen et al., 1999; Kimura et al., 2002), significant protection from cerebral malaria has been demonstrated (Genton et al., 1995; Allen et al., 1999). Furthermore, four of five tested strains of Plasmodiam falciparum were not able to effectively invade B3∆27 red blood cells in a recent in vitro study (Cortes et al., 2004). Although the specific protective mechanism remains as yet unclear, these data suggest that the B3∆27 is maintained through balancing heterozygote selection in malarious environments. Previous studies within PNG (Mgone et al., 1996; Kimura et al., 2003) revealed that the B3∆27 frequency
Figure 15.6 Map showing the location of the study populations within Papua New Guinea. The ecological zones are indicated by symbols: open circles, off-shore islands; filled circles, coastal plain; triangles, Prince Alexander Ridge; filled squares, Sepik plain.
224
Contributions of Population Origins and Gene Flow
was inversely correlated with altitude and thus, with malaria endemicity. Globally B3∆27 has been found from the Philippines to Malaysia, Indonesia, and PNG to the east and to Mauritius, South Africa, and Madagascar to the west (Jarolim et al., 1991; Ravindranath et al., 1994; Mgone et al., 1996; Kimura et al., 1998; Rabe et al., 2002; Kimura et al., 2003). This distribution is coincident with the diaspora of Austronesian-speaking populations. These data generated two main hypotheses for the current distribution of B3∆27: (1) positive selection through malaria protection in the malarious lowlands of PNG (Mgone et al., 1996), and (2) dispersal via migrating Austronesian-speaking populations from island Southeast Asia through the Pacific and Indian Oceans during the last 3,500 years (Kimura et al., 2003). These two hypotheses are difficult to distinguish because Austronesian-speaking populations are restricted to malarious islands and lowlands of PNG and are not found in the non-malarious highlands. In addition, since there has been extensive gene flow between Austronesian- and Papuan-speaking populations during the initial formation of the Lapita Cultural Complex and/or within the last 3,500 years (Lum et al., 1998, 2002; Merriwether et al., 1999; Cann and Lum, 2004), genetic markers associated with the diaspora of Austronesianspeaking populations in the Pacific are also found in Papuan-speaking coastal and island populations of Near Oceania. To help tease apart these subtleties, we assayed the mtDNA region V deletion thought to be a marker for the Lapita expansion (Hertzberg et al., 1989; Lum et al., 1994; Sykes et al., 1995; Lum and Cann, 1998, 2000; Merriwether et al., 1999) from the same individuals as were screened for the B3∆27. By comparing the distribution of this neutral marker associated with the Austronesian expansion within the Pacific, we were able to assess the relative contributions of population origin and gene flow to the distribution of the malaria selected B3∆27. To investigate the first hypothesis, that positive selection throughout the malarious lowlands of PNG was responsible for current distributions of the B3∆27, we compared population frequencies of the B3∆27 to measures of malaria endemicity estimated from both malaria point prevalence and spleen rate from malariometric surveys of the same populations. The population frequencies of the B3∆27 were not significantly correlated with spleen rate (r = –0.34, p = 0.41) but were significantly inversely correlated with parasite prevalence (r = –0.756, p = 0.03). These data indicate that positive malaria selection alone does not explain the current distributions of the B3∆27. In contrast, the frequencies of the B3∆27 are significantly correlated with proximity to the coast (r = 0.81, p = 0.016). The highest frequencies of the B3∆27 are found in villages of off-shore islands and the coastal plain expected to be
most impacted by Austronesian-speaking populations and their Lapita colonist descendants (Tsukahara et al., 2006). The frequencies of the B3∆27 was also significantly correlated with frequencies of the mtDNA region V deletion (r = 0.88, p = 0.004) (figure 15.7), consistent with both deletions originating with the arrival of Austronesian speakers and limited dispersal into inland PNG. This inference is consistent with oral traditions recounting the expansion of Boiken-speakers from the Sepik Plain to the coast mentioned above (Roscoe, 1989; Miki, 1980). Our results support the maintenance of the B3∆27 in malarious environments, but only within populations genetically descended from the Lapita peoples. In the malaria endemic region of our study site in East Sepik Province, the selective advantage of the B3∆27 was apparently not sufficient to extend its distribution inland against the preponderant flow from the interior of Boiken migrants and their genes.
Malaria Selection within Vanuatu (Remote Oceania) In contrast to the study populations of Near Oceania described above, Vanuatu was settled relatively recently during the initial expansion of the Lapita cultural complex 3.2 kybp (Kirch, 2000) and is the easternmost malarious
Figure 15.7 Correlation between the frequencies of the B3∆27 allele and the mtDNA region V deletion. The symbols indicate the ecological zone of the populations as described in figure 15.6.
225
regional studies and conclusion
archipelago of the Pacific (Kaneko, 1999). Although Vanuatu does not contain Papuan languages, at the time of European contact 110 Austronesian languages were spoken by inhabitants of approximately 100 islands. Some of the larger islands like Santo and Malakula were inhabited by speakers of nearly 40 languages each resulting in Vanuatu being the most linguistically diverse region of the world per capita (Tryon, 1996; Gordon, 2005). An active area of our research examines how substantial linguistic, cultural, and corresponding genetic diversity was maintained among interacting populations on such a small scale and in turn, how this has influenced the ways that populations respond to ecological gradients of malarial endemicity. Previous studies within New Guinea and Island Melanesia revealed a correlation between frequencies of α-thalassemias and both altitude and latitude implicating long-term malaria selection for the frequencies of these resistant alleles (Flint et al., 1986). Within the Pacific, α-thalassemia results from two distinct deletions of 4.2 and 3.7 kilobases (kb). The 4.2 kb deletion is thought to have arisen in PNG, while restriction fragment length polymorphism (RFLP) defined subtypes of the 3.7 kb deletion are thought to have evolved in both Southeast Asia (subtypes I and II) and PNG (subtypes III and IV) (Hill et al., 1985). Vanuatu contains α-thalassemia deletions and subtypes originating in both Southeast Asia and PNG, indicating that alleles from both ancestral populations of Pacific Islanders were brought into Remote Oceanic Island Melanesia during colonization or via post settlement gene flow over the past 3,300 years (Hill et al., 1985, Roberts-Thompson et al., 1996). Within Vanuatu malaria endemicity is highest in the north and decreases with distance away from the equator. Kaneko et al. (1998) showed that the frequency of glucose 6-phosphate dehydrogenase (G6PD) deficiency which confers resistance to malaria likewise shows a north–south cline within Vanuatu with the highest frequencies in the North where malaria selection is most intense. To further investigate the effect of this north–south gradient of malaria selection within the archipelago we (Ubalee et al., 2005) examined two additional loci, one implicated in malaria susceptibility and the other with resistance from 1,074 individuals representing six islands spanning the north–south range of the Vanuatu archipelago (Gaua, Santo, Pentecost, Malakula, Erromango, and Aneityum) (figure 15.8). The susceptibility locus we examined was a promoter polymorphism of the cytokine tumor necrosis factor-α (TNF-α). TNF-α is a pro-inflammatory cytokine that plays an essential role in protection against many infectious diseases, but is also fatal in excess (Knight et al., 1999). Single nucleotide polymorphisms (SNPs) within the TNF-α promoter region (TNFP) are associated with both up and down regulation of transcription.
226
In particular, the TNFP-D allele results in the up regulation of TNF-α and consequently, an increased risk of cerebral malaria in Myanmar (Ubalee et al., 2001). We also determined the frequencies of the resistance conferring αthalassemias (4.2 kb deletion and 3.7 kb deletion subtypes) because they are expected to respond to the malaria endemicity gradient in the opposite way as the frequencies of the susceptible TNFP-D allele. Comparisons of both of these alleles from the same individuals provided an internal control across our malaria selection gradient. Within our six island populations of Vanuatu, we observed four TNFP alleles (A, B, C, and D). The TNFPA, B, and D alleles range in frequency from 0.14 to 0.55 in each population, while the TNFP-C allele was uniformly rare (0.00–0.03). The TNFP-D allele was significantly inversely correlated (r = –0.855, p = 0.03) with malaria endemicities reported by Kaneko et al. (1998), consistent with its up regulation of TNF-α transcription and consequent increased risk of cerebral malaria (Ubalee et al., 2005). The TNFP-D allele has been found previously in Japan and Myanmar and so may be of East Asian origin or common throughout the Pacific region. To explore these alternatives, we are currently extending our surveys of TNFP polymorphisms to our study populations from East Sepik Province, PNG, discussed above. The frequencies of α-thalassemia alleles were significantly inversely correlated with those of TNFP-D (r = −0.962, p = 0.002) (figure 15.9), as expected for alleles responding to malaria selection in opposite ways (Ubalee et al., 2005). The frequencies of α-thalassemias within the six populations ranged from 0.11 to 0.39. The 4.2 kb deletion, thought to have originated in PNG was less common in all populations and ranged from 0.03 to 0.12. Nearly all of the 3.7 kb deleted alleles (>0.97) were subtype III, also believed to have evolved in PNG. Thus, the overwhelming majority of the malaria protective α-thalassemia alleles of Vanuatu (0.99) source to Near Oceania rather than Southeast Asia. Within Southeast Asia α-thalassemia frequencies range from 5–9%, which is relatively low compared to PNG and Vanuatu (9–40%) (O’Shaughnessy et al., 1990). The relatively high frequencies of α-thalassemias inferred to be nearly exclusively of PNG origin are consistent with malaria selecting for resistant alleles arriving via post-settlement gene flow from Near Oceania.
Discussion Patterns of linguistic, cultural, and biological affinities among Pacific Island populations are correlated at some scales but conflict at others, reflecting the complex interactions between the descendants of the Pleistocene settlers
Contributions of Population Origins and Gene Flow
Figure 15.8 Map showing the location, number of individuals examined, and the annual parasite incidence per 1,000 individuals (API) of each Vanuatu population examined.
of Near Oceania and the more recent Holocene expansion of Austronesian-speakers. Here I have reviewed three studies of either neutral or malaria selected autosomal loci to illustrate the contribution of population origins, gene flow, and disease selection to the portioning of genetic variation among current Pacific Island populations. Neutral genetic diversity within populations generally decreases with distance from Southeast Asia, with the exception of Austronesian-speaking populations of Vanuatu that maintain relatively large amounts of diversity possibly accumulated via post-settlement gene flow from Near Oceania (figure 15.2), resulting in a clustering of Melanesian populations (except Fijians) regardless of linguistic affiliation (figure 15.3). On the other hand,
malaria has apparently maintained the B3∆27 through the positive selection of heterozygotes, but only among the descendants of Lapita colonists who likewise harbor the mtDNA region V deletion (figure 15.7). Within Vanuatu, an archipelago settled during the initial Lapita expansion, a north–south gradient of malaria endemicity has resulted in corresponding gradients of resistance and susceptibility conferring alleles (figure 15.9). Although the origin of the TNFP-D allele is still unknown, nearly all of the α-thalassemia alleles of Vanuatu are of inferred PNG origin, consistent with an accumulation over time of alleles that originated in Near Oceania. These vignettes reveal the importance of evolutionary histories of populations to the understanding of their
227
regional studies and conclusion
Figure 15.9 Correlation between TNFP and α-thalassemia allele frequencies.
response to pathogens. Our East Sepik Province populations were selected at a sufficiently fine scale to correspond to prehistoric population interactions. Unfortunately most genetic studies of Pacific Island populations do not reflect realistic interaction spheres (see figure 15.1). This in part reflects the logistical problems inherent to accurately sampling the more than 1,500 linguistic groups of the Pacific (Clarke, 1992; Foley, 1992), and results in our knowledge of Pacific population genetics remaining thus far in its infancy. A meaningful understanding of the effects of the complex interactions among populations and their environments on their cultural and biological evolution awaits a more fine-scaled examination of the diversity of the Pacific.
Acknowledgments I wish to thank M Amos, A Björkman, R Blust, RL Cann, K Hirayama, FW Hombhanje, I Hwaihwanje, L Jorde, M Kalkoa, A Kaneko, T Kobayakawa, JJ Martinson, A Masta, H Osawa, R Regenvanu, K Rehg, M Sapuri, W Schiefenhovel, M Spriggs, G Taleo, T Tsukahara, F Valentin, J Yaviong for often spirited conversations over the years, all of my other collaborators at the University of Papua New Guinea, Wewak General Hospital, Vanuatu Ministry of Health, and Vanuatu Kaljoral Senta, Tokyo Women’s Medical University, Nagasaki University, the many field assistants, traditional leaders of Rubekul Belau and Council of Pilung, and the
228
thousands of participants without whom these studies would not have been possible.
References Allen SJ, O’Donnell A, Alexander ND, Mgone CS, Peto TE, Clegg JB, Alpers MP, and DJ Weatherall. 1999. Prevention of cerebral malaria in children in Papua New Guinea by Southeast Asian ovalocytosis band 3. American Journal of Tropical Medicine and Hygiene 60: 1056–60. Bender BW. 1971. Micronesian languages. In TA Sebeok (ed.: Current trends in linguistics. The Hague, the Netherlands: Mouton, pp 426–65. Cann RL and JK Lum. 2004. Dispersal ghosts in Oceania. American Journal of Human Biology 16: 440–51. Cann RL and AC Wilson. 1983. Length mutations in human mitochondrial DNA. Genetics 104: 699–711. Clarke R. 1992. Austronesian Languages. In W Bright (ed) International Encyclopedia of Linguistics. Vol. 1. New York, Oxford University Press, pp 142–5. Cortes A, Benet A, Cooke BM, Barnwell JW, and JC Reeder. 2004. Ability of Plasmodium falciparum to invade Southeast Asian ovalocytes varies between parasite lines. Blood 104: 2961–6. Costa Lima MA and MM Pimentel.2004. Dynamic mutation and human disorders: the spinocerebellar ataxias. International Journal of Molecular Medicine 13: 299–302.
Contributions of Population Origins and Gene Flow
Diamond JM. 1988. Express train to Polynesia. Nature 336: 307–8. Flint J, Hill AV, Bowden DK, Oppenheimer SJ, Sill PR, Serjeantson SW, Bana-Koiri J, Bhatia K, Alpers MP, Boyce AJ, Weatherall DJ, and JB Clegg. 1986. High frequencies of alpha-thalassaemia are the result of natural selection by malaria. Nature 321: 744–50. Foley WA. 1992. New Guinea Languages. In W Bright (ed) International Encyclopedia of Linguistics. Vol. 3. New York, Oxford University Press, pp 87–90. Genton B, al-Yaman F, Mgone CS, Alexander N, Paniu MM, Alpers MP, and D Mokela.1995. Ovalocytosis and cerebral malaria. Nature 378: 564–5. Gordon RG Jr (ed). 2005. Ethnologue: Languages of the World, Fifteenth edition. Dallas, SIL International, Online version: http://www.ethnologue.com/. Green RC. 1991. Near and Remote Oceania: disestablishing “Melanesia” in cultural history. In: A Pawley (ed): Man and a half: essays in Pacific anthropology and ethnobotony in honour of Ralph Bulmer. Auckland, The Polynesian Society, Inc., pp 491–502. Green RC. 1999. Integrating historical linguistics with archaeology: insights from research in Remote Oceania. Indo-Pacific Prehistory Association Bulletin 18: 3–16. Groube L, Chappell J, Muke J and D Price.1986. A 40,000 year-old human occupation site at Huon Peninsula, Papua New Guinea. Nature 324: 453–5. Hanihara T. 1993. Population prehistory of East Asia and the Pacific as viewed from craniofacial morphology: the basic populations in East Asia, VII. American Journal of Physical Anthropology 91: 173–87. Hertzberg M, Mickleson KNP, Serjeantson SW, Prior JF, and RJ Trent. 1989. An Asian-specific 9-bp deletion of mitochondrial DNA is frequently found in Polynesians. American Journal of Human Genetics 44: 504–10. Hill AV, Bowden DK, Trent RJ, Higgs DR, Oppenheimer SJ, Thein SL, Mickleson KN, Weatherall DJ, and JB Clegg. 1985. Melanesians and Polynesians share a unique αthalassemia mutation. American Journal of Human Genetics 37: 517–80. Hombhanje FW, Hwaihwanje I, Tsukahara T, Saruwatari J, Nakagawa M, Osawa H, Paniu MM, Takahashi N, Lum JK, Aumora B, Masta A, Sapuri M, Kobayakawa T, Kaneko A, and T Ishizaki. 2005. The disposition of oral amodiaquine in Papua New Guinean children with falciparum malaria. British Journal of Clinical Pharmacology 59: 298–301. Intoh M. 1997. Human dispersal into Micronesia. Anthropological Science 105: 15–28. Jackson FH. 1986. On determining the external relationships of the Micronesian languages. In: Geraghty P, Carrington L, and SA Wurm (eds): FOCAL II: Papers from the Fourth International Conference on Austronesian Linguistics. Pacific Linguistics C-94. Canberra, Australian National University, pp 201–38. Jarolim P, Palel J, Amato D, Hassan K, Sapak P, Nurse GT, Rubin HL, Zhai S, Sahr KE, and SC Liu. 1991. Deletion in erythrocyte band 3 gene in malaria-resistant
Southeast Asian ovalocytosis. Proceedings of the National Academy of Sciences, USA 88: 11022–6. Kaneko A. 1999. Malaria on Islands: Human and parasite diversities and implications for malaria control in Vanuatu. PhD dissertation, Stockholm, Sweden: Karolinska Institutet. Kaneko A, Taleo G, Kalkoa M, Yamar S, Kobayakawa T, and A Bjorkman. 1998. Malaria epidemiology, glucose 6-phosphate dehydrogenase deficiency and human settlement in the Vanuatu Archipelago. Acta Tropica 70: 285–302. Kimura M, Shimizu Y, Settheetham-Ishida W, Soemantri A, Tiwawech D, Romphruk A, Duangchan P, and T Ishida. 1998. Twenty-seven base pair deletion in erythrocyte band 3 protein gene responsible for Southeast Asian ovalocytosis is not common among Southeast Asians. Human Biology 70: 993–1000. Kimura M, Soemantri A, and T Ishida. 2002. Malaria species and Southeast Asian ovalocytosis defined by a 27-bp deletion in the erythrocyte band 3 gene. Southeast Asian Journal of Tropical Medicine and Public Health 33: 4–6. Kimura M, Tamam M, Soemantri A, Nakazawa M, Ataka Y, Ohtsuka R, and T Ishida. 2003. Distribution of a 27-bp deletion in the band 3 gene in South Pacific islanders. Journal of Human Genetics 48: 642–5. Kirch PV. 2000. On the road of the winds: an archaeological history of the Pacific Islands before European contact. Berkeley, CA: University of California Press. Knight JC, Udalova I, Hill AV, Greenwood BM, Peshu N, Marsh K, and D Kwiatkowski.1999. A polymorphism that affects OCT-1 binding to the TNF promoter region is associated with severe malaria. Nature Genetics 22: 145–50. Leahy M. 1991. Exploration into highland New Guinea 1930–1935. Tuscaloosa, AL: University of Alabama Press. Lum JK and RL Cann. 1998. MtDNA and language support a common origin of Micronesians and Polynesians in island Southeast Asia. American Journal of Physical Anthropology 105: 109–19. Lum JK and RL Cann. 2000. mtDNA lineage analyses: origins and migrations of Micronesians and Polynesians. American Journal of Physical Anthropology 113: 151–68. Lum JK, Rickards O, Ching C and RL Cann. 1994. Polynesian mitochondrial DNAs reveal three deep maternal lineage clusters. Human Biology 66: 567–90. Lum, JK, RL Cann, JJ Martinson and LB Jorde. 1998. Mitochondrial and nuclear genetic relationships among Pacific island and Asian populations. American Journal of Human Genetics 63: 613–24. Lum JK, LB Jorde, and W Schiefenhovel. 2002. Affinities among Melanesians, Micronesians, and Polynesians: A neutral, biparental genetic perspective. Human Biology 74: 413–30. Lum JK, A Kaneko, K Tanabe, N Takahashi, A Björkman, and T Kobayakawa. 2004. Malaria dispersal among
229
regional studies and conclusion
islands: human and Plasmodium falciparum gene flow in Vanuatu. Acta Tropica 90: 181–5. Masta A, Lum JK, Tsukahara T, Hwaihwanje I, Kaneko A, Paniu MM, Sapuri M, Takahashi N, Ishizaki T, Kobayakawa T, and FW Hombhanje. 2003. Analysis of Sepik populations of Papua New Guinea suggests an increase of CYP2C19 null allele frequencies during the colonization of Melanesia. Pharmacogenetics 13: 697–700. Merriwether DA, Friedlaender JS, Mediavilla J, Mgone C, Gentz F, and RE Ferrell. 1999. Mitochondrial DNA variation is an indicator of Austronesian influence in Island Melanesia. American Journal of Physical Anthropology 110: 243–70. Mgone CS, Koki G, Paniu MM, Kono J, Bhatia KK, Genton B, Alexander NDE, and MP Alpers. 1996. Occurrence of the erythrocyte band 3 (AE1) gene deletion in relation to malaria endemicity in Papua New Guinea. Transactions of the Royal Society of Tropical Medicine and Hygiene 90: 228–31. Miki J. 1980. How one language came about. Wantok 299: 17. O’Shaughnessy DF, Hill AVS, Bowden DK, Weatherall DJ and JB Clegg. 1990. Globin genes in Micronesia: origins and affinities of Pacific island peoples. American Journal of Human Genetics 46: 144–55. Pawley A and R Green. 1973. Dating the dispersal of the Oceanic languages. Oceanic Linguistics 12: 1–67. Pawley A and M Ross. 1993. Austronesian historical linguistic and cultural history. Annual Review of Anthropology 22: 425–59. Pietrusewsky M. 1990. Craniofacial variation in Australasian and Pacific populations. American Journal of Physical Anthropology 82: 319–40. Rabe T, Jambou R, Rabarijaona L, Raharimalala L, Rason MA, Ariey F, and D Dhermy. 2002. South-East Asian ovalocytosis among the population of the Highlands of Madagascar: a vestige of the island’s settlement. Transactions of the Royal Society of Tropical Medicine and Hygiene 96: 143–4. Ravindranath Y, Goyette G Jr, and RM Johnson. 1994. Southeast Asian ovalocytosis in an African-American family. Blood 84: 2823–4. Roberts-Thomson JM, Martinson JJ, Norwich JT, Harding RM, Clegg JB, and B Boettcher. 1996. An ancient common
230
origin of Aboriginal Australians and New Guinea Highlanders is supported by β-globin haplotype analysis. American Journal of Human Genetics 58: 1017–24. Roscoe BR. 1989. The flight from the fen: the prehistoric migrations of the Boiken of the East Sepik Province, Papua New Guinea. Oceania 60: 139–54. Schneider S, Roessli D, and L Excoffier. 2000. Arlequin: a software for population genetics data analysis Ver. 2.000. Genetics and Biometry Lab, Department of Anthropology, University of Geneva. Spriggs M and A Anderson. 1993. Late colonization of East Polynesia. Anitquity 67: 200–17. Sykes B, Leiboff A, Low-Beer J, Tetzner S, and M Richards. 1995. The origins of the Polynesians: and interpretation from mitochondrial lineage analysis. American Journal of Human Genetics 57: 1463–75. Terrell JE, Kelly KM, and P Rainbird. 2001. Forgone conclusions? An analysis of the concepts of “Austronesians” and “Papuans.” Current Anthropology 42: 97–124. Tryon D. 1996. Dialect chaining and the use of geographical space. In: Bonnemaison J, Kaufmann C, Huffman K, and D Tryon (eds) Arts of Vanuatu.Honolulu, University of Hawai’i Press, pp 170–3. Tsukahara T, Hombhanje FW, Lum JK, Hwaihwanje I, Masta A, Kaneko A, and T Kobayakawa. 2006. Austronesian origin of the 27-bp deletion of the erythrocyte band 3 gene in East Sepik, Papua New Guinea inferred from mtDNA analysis. Journal of Human Genetics (Jan 21, Epub ahead of print). Ubalee R, Suzuki F, Kikuchi M, Tasanor O, Wattanagoon Y, Ruangweerayut R, Na-Bangchang K, Karbwang J, Kimura A, Itoh K, Kanda T, and K Hirayama. 2001. Strong association of a tumor necrosis factor-α allele with cerebral malaria in Myanmar. Tissue Antigens 58: 407–10. Ubalee R, Tsukahara T, Kikuchi M, Lum JK, Dzodzomenyo M, Kaneko A, and K Hirayama. 2005. Associations between frequencies of a susceptible TNF-alpha promoter allele and protective alpha-thalassaemias and malaria parasite incidence in Vanuatu. Tropical Medicine and International Health 10: 544–9. Wickler S and M Spriggs. 1988. Pleistocene human occupation of the Solomon Islands, Melanesia. Antiquity 62: 703–6.
16 Conclusion Jonathan S. Friedlaender
William W. Howells once wrote that Melanesian biological diversity was so Protean as to defy analysis. He would be pleased to know this is no longer the case. A comprehensible structure of the variation has emerged, and we are beginning to understand the dynamics responsible. Testing different explanatory models has commenced and this effort should gain power and sophistication in the next couple of years with our extensive sample set.
Specific findings 1. Detection of ancient genetic variants. A number of haplogroups and alleles have developed in particular sub-regions of the Southwest Pacific. These have remarkably restricted distributions. With regard to the mtDNA and NRY, many of these are clearly very old, certainly dating to the Upper Pleistocene, and some very likely date back to initial settlement times (for the mtDNA, haplogroups M27, M28, M29, Q1, Q2, Q3, P1, P2, P3, and P4 are such candidates; for the NRY, haplogroups M2a,K5, K6, K7, and C2b). Some autosomal variants have similarly restricted distributions and may be of similar ages (e.g. GM*A,X G and GM*A B). 2. Recently introduced/modified variants. There are clear genetic signals of relatively recent influence from Island Southeast Asian, perhaps Taiwanese sources, very likely related to the appearance of the Oceanic branch of Austronesian languages and the development of the Lapita Cultural Complex. While this is most apparent in the mtDNA (haplogroups B4a1a and E1), variation at the GM locus echoes the same pattern (GM*A,F B). However, Y chromosome
variation reflects surprisingly little recent Island Southeast Asian influence (7–10% of the Northern Island Melanesian series, vs. >30% for the mtDNA “Polynesian Motif”). A number, but not all, of the recently introduced Island Southeast Asian genetic types have developed local variants in Near Oceania since their appearance in the late Holocene. The full “Polynesian Motif” (B4a1a1) has its greatest diversity in New Guinea, New Ireland, and Bougainville, and therefore probably developed in this region (a caution: current haplotype distributions and age estimates cannot establish ancient origin locations with any greater precision than historical linguistics—as in the case of Indo-European, or the Finns). Although it is found as far west as central Indonesia, E1b has its greatest frequency in certain New Britain populations. Among the malaria-selected variants discussed in chapter 15, some Vanuatu α-thalassemia variants first developed in Southeast Asia, while others developed in lowland New Guinea. The allele associated with malaria-protecting ovalocytosis, B3∆27, is another apparently unchanged Southeast Asian variant that is associated with the Austronesian Diaspora, at least as far east as the New Guinea lowlands. 3. Population (genetic) distinctions. Populations from the largest landmasses in the Southwest Pacific differ genetically to a remarkable degree. Southwest Pacific populations are genetically separate from Island Southeast Asian ones; aboriginal Australia has only very ancient genetic connections with New Guinea; and within Near Oceania, the major islands as far southeast as Bougainville are significantly different
231
regional studies and conclusion
one from another. New Guinea, New Britain, New Ireland, and Bougainville populations each have differing allele frequency profiles. The islands to the southeast are smaller, were settled later, and have been heavily influenced by recent population movements, as shown by the genetic evidence from Malaita, Santa Cruz, and Vanuatu. Not only is there evidence of extensive contact between neighboring groups throughout the region, but the distribution of Western Oceanic, with its distinct demarcation in the Solomons, indicates a secondary expansion of Western Oceanic-speaking peoples from the New Britain/New Ireland region some time after the Lapita period (as discussed in chapter 8). This could be responsible for a great deal of the mixing evident over much of Northern Island Melanesia. Neutral variation decreases from west to east across the Pacific, with the notable exception of Near Oceania (see chapter 15). The main reason it does not decrease in Near Oceania is because the islands there are big and rugged, were accessible from major dispersal points, and were therefore settled early. 4. Language and genetics. It is certain Papuan-speaking groups, and particularly inland ones, that are genetically the most distinctive from other groups in Island Melanesia. The Aita and Rotokas of north Bougainville, the Baining (Mali and Kaket) of New Britain, and also the Kuot and Nagovisi, appear to constitute three separate and important poles of variation. The Aita/Rotokas of north Bougainville and the two Baining groups seem to retain especially old and distinctive profiles. Whether or not one accepts the age estimates for Baining divergences presented in chapter 13, that study shows just how great the genetic divergence among certain closely situated Papuan groups can be. One of the most perplexing situations involves the Kuot and Nagovisi extreme “pole” in the mtDNA MDS plot (chapter 4). Why should speakers of two unrelated Papuan languages on two different islands be united in having extremely high frequencies of the so-called Polynesian Motif, or variants thereof? Does this mean that all the extreme population variation shown in that MDS plot is simply the result of isolation over the last 3,000 years, rather than being a much older residue of Upper Pleistocene times? Other populations, whether Papuan- or Oceanicspeaking, show signs of greater intermixture. While language distinctions therefore clearly are informative with regard to (genetic) population variation, simple models of language–gene associations yield significant but weak overall signals (see chapter 9). It could be that different dynamics underlie language
232
versus population differentiation, but it may be simply the independent transmission of variation in the two systems that results in the low level of correspondence. Both the linguistic and genetic evidence do show a very strong signal of population intermixture in much of the region, especially in the Oceanic-speaking areas. Neither models of simple bifurcative tree-like diversification nor simple geographic distance can account, by themselves, for the major patterns of variation. 5. Extensive within-island variation. In most single gene studies, the variation among populations within an island is very great, at least as considerable as the variation from island to island. This has been established for this region for a long time (Green, 1967; Ward, 1967; Friedlaender, 1971). What has caused this pattern is not simple to explain. Certainly, census sizes had to have been small. Marital migration distances, especially among inland populations, were generally very short. But populations also periodically moved around for a number of reasons. At least at this time, the amongpopulation variation continues to be significant, although it is unlikely to continue much longer because of displacements caused by increasing mobility (development projects and associated roads, urban centers, and so on). Why polygenic characters such as skin color do not show comparable extensive within-island variation is not entirely clear (chapter 6), but it may simply be because variation at a number of loci are involved and average each other out to a degree. As mentioned, some degree of stabilizing selection for a certain level of melanization in the region would likely be involved as well. 6. Contacts between islands. In spite of this genetic patterning, total isolation of populations and groups probably never existed here. Evidence from a variety of sources attests to inter-island contacts and expansions, not only within Near Oceania, but extending to Remote Oceania after it was finally settled at ~3,200 YBP. That includes trade and the intentional (and unintentional) introduction of plants and animals to different regions. New Guinea and the Bismarcks were centers of dispersal for a long time, but Bougainville appears to have remained surprisingly isolated until relatively recently.
Problems and Prospects 1. Differences in male and female marital and demographic patterns. The clearest example of this problem is the distinction between the mtDNA and nonrecombining Y (NRY) signals of recent Island
Conclusion
Southeast Asian/Taiwan (Austronesian) influence. While about 30% of the Northern Island Melanesian mtDNA is haplogroup B4a1a1 (the “Polynesian Motif”), less than 10% of the NRY samples (the total frequency of O:O3 has a frequency of 7%) appear to derive from there recently. This contradiction appears to extend to Polynesian samples as well, although coverage there is not entirely satisfactory. One proposed explanation is that the ancient Austronesian-speaking populations were “... matriarchal and matrilocal (as the Amis tribe still is in Taiwan) whereby the Y chromosome pool of the initial migrants was lost after being repeatedly diluted on the way toward Polynesia” (Trejaut et al., 2005). On the face of it, this argument conjures up images of Austronesian women sailing ever eastward, discarding Austronesian men and picking up Papuan consorts as they traveled. More explicitly, Cann and Lum (2004) proposed an explanation that stressed a major distinction between Austronesian initial settlement patterns and subsequent overlays: ...if post settlement gene flow was predominantly male-biased, as suggested by matrilocal patterns of land tenure (Hage and Marck, 2003), malespecific navigation magic (Haddon and Hornell, 1975 (reprint)) and the vagaries inherent with open ocean travel and landfall on a foreign island, maternal genetic patterns may largely reflect settlement patterns, paternal genetic patterns may reflect post settlement gene flow, and the biparental genome may reveal an intermediate pattern. As a possible example of this, I was told that the Cartarets population (Kilinailau Island) just over the eastern horizon from Bougainville was once a Polynesian Outlier group, but a Bougainville man discovered it by chance when his canoe was driven there by a storm, later returned with his kinsmen, killed the Polynesian males, took the women as wives, and settled in. This would have produced subsequent generations with “Polynesian” mtDNA (all B4a1a1) and “Papuan” NRY variants. Could a similar happening have led to the initial colonization of Remote Oceania? Hypothetical arguments as to what caused early Austronesian wanderlust had focused on the role of rank and competition among sons, with second and third sons inclined to leave home (presumably with their own wives) in search of more salubrious islands less under the thumbs of their first-born brothers (Pawley, 1982; Pawley and Ross, 1993). A number of authors have stressed the founder-focused ideology and rank enhancement that would have been
particularly important in the colonization process (Bellwood, 1996; Kirch and Green, 2001; Green, 2002). This should have led to an initial settlement pattern with a few successful Southeast Asian-derived Y chromosome variants, and also a few highly successful mtDNA variants (i.e., a series of very strong bottleneck effects, with loss of variation and possible fixation of particular haplotypes in the mtDNA and NRY, but certainly no predicted selection against all Southeast Asian variants on either the mtDNA or NRY). I, at least, would have predicted a reverse scenario, with more Southeast Asian-derived NRY variants than in the mtDNA, although Cann and Lum are suggesting a compensating overlay that may provide a plausible way out. It has other attractions. Besides accounting for the mtDNA/NRY discrepancy in the Pacific, it could be used to argue that the mtDNA haplotypes retained the most ancient, “settlement” signature in the Papuan areas as well. Certainly the mtDNA distributions are old and distinctive by region here (which would make this region an exception to apparent trends elsewhere). However, the high frequency of the “Polynesian Motif” in south Bougainville Papuan speakers suggests exactly the opposite of the Cann/Lum model—that “Motif”bearing women married (or were kidnapped) into this Papuan-speaking, matrilocal region comparatively recently. In other studies, distinctions between matrilocal and patrilocal residence patterns have been successfully associated with differences in mtDNA and NRY patterns of variation, with matrilocality being associated with greater mtDNA than NRY diversity among populations, and vice versa for patrilocal groups (e.g., Oota et al., 2001). But as detailed in chapter 13, a key factor in such discrepancies may well be the considerably smaller effective population size of men than women: in the Baining, the effective male population size is estimated to be a fourth of women’s, and the suggestion is that men who moved were especially productive fathers. Clearly, the explanation for the mtDNA/NRY discrepancy in Near and Remote Oceania is not yet resolved. Nor is the very high frequency of mtDNA haplogroup B4a1a1 in certain Papuan-speaking groups any better explained. Some may suggest that differential natural selection may account for the high frequencies of this haplotype (e.g., Mishmar et al., 2003), but this is something of a post-hoc hypothesis, and not satisfying as a result. In terms of reliability, the mtDNA database is both far more extensive than that for the NRY in this region,
233
regional studies and conclusion
because to date, reliable NRY results require better quality DNA samples. Perhaps more important in the long run, mtDNA sequence results are unbiased while the NRY SNP data are not, and may therefore still be yielding a poor record of variation in the region (see chapters 5 and 9). Extensive sampling from the biparental genome will help to resolve some of these issues, or at least put them in a larger perspective. Part of the problem has been the poor definition of what constitutes “Polynesian” population variation. The analyses of a very large battery of STRs and in/dels on these and other Pacific samples are now underway by our group. 2. Dating problems. Most prehistory datasets lose their power the further back in time reconstructions are attempted. Ideally, the archeological record would be the lynch pin, but hard evidence for earlier periods drops off at an almost exponential rate, especially in this wet decaying tropical environment where rising sea levels have concealed a substantial portion of the prehistoric record. While classical historical linguistics can identify language origins, contacts, and even aspects of ancient environments and cultural complexes, its power also deteriorates rapidly over longer time depths as the evidence decays. The accepted horizon for confidently identifying family linguistic relationships and proto-language forms with the Comparative Method is only in the vicinity of 6,000–10,000 years. This is inadequate for most of the occupational period of Island Melanesia, although as chapter 8 shows, other methods are being developed which offer hope for the establishment of older links. Biological anthropological approaches to reconstructions of ancient population migrations and expansions have, until recently, been even less informative. One could argue that a photo tour through the region reveals as much about population biological affinities as most biological studies carried out prior to the last decade (see the book cover and portraits in the color insert). For example, in my early Bougainville research, I was happy to discern a marked division between south and north Bougainville (Papuan) groups in body and head shapes with a multiple discriminant analysis of a number of measurements: northerners had broader faces and bodies than southerners. This really only confirmed what a number of discerning Bougainvilleans had told me themselves (see Friedlaender, 1975, Preface). Decades of work on detailed measurements of Oceanic crania, teeth, bodies, finger and palm prints, and classical blood polymorphisms can be summarized fairly succinctly.
234
Eastern Polynesians form a homogeneous group, indicating Southeast Asian and also modest Island Melanesian affinities; Micronesians vary from west to east, with considerably more Asian influence in the west and Island Melanesian and Polynesian affinities to the east; Australian Aboriginals are very distinctive, showing no Asian influences; and New Guinea and Island Melanesian populations are ambiguous in their relationships, but show an inordinate variability from one region to the next. There are indications of a distant relationship between some of these Near Oceanic populations (especially New Guinea Highlanders) and Australian Aboriginals, some clearer suggestions of Southeast Asian affinities in some regions, and a lot of residual distinctiveness that has been difficult to link to populations elsewhere. An example of this sort of array is the excellent craniometric work of Michael Pietrusewsky, most recently reported in a collection on East Asian prehistory (Pietrusewsky, 2005). Island Melanesians occupy the broad and ambiguous middle ground in the multivariate craniometric space his skulls define (my interpretation, not his). The same has been generally true for studies of dermatoglyphics (finger and palm prints), measurements on living individuals, and dental characteristics, at least where sampling has been adequate. Genetic work suggesting the same general patterns has been recently reviewed by Attenborough (2005) and Main et al. (2005), including the highly polymorphic HLA system, which has been particularly difficult to interpret in the past before DNA sequencing methods were developed. Although coverage with newer techniques is still spotty, New Guinea populations show a great deal of variability, with some suggestions of ancient Australian Aboriginal ties to Eastern Highlanders and a possible series of pre-Austronesian population distinctions in different regions. The major problem of these biological comparisons is that they cannot be linked directly to any chronological sequence. There is no biological stratigraphy, no nested set, only similarities. There are also clear limits to the new power provided by the analysis of the non-recombining male- and female-inherited mtDNA and NRY analyses. Age estimates for lineage divergence times, coalescence times, or population expansions, whether for the mtDNA NRY, or JCV, should be regarded as relative dates only, not anything approaching a real chronology. The ages presented in the mtDNA and NRY chapters are generally uncomfortably old, as noted. Their calibration
Conclusion
depends either on (differing) chimpanzee/human divergence estimates, or on ancient human population bifurcation dates, assumed to be undiluted by subsequent mixing. This is especially troubling in some NRY age estimates, based in part on very tenuous, if not misleading and outdated reconstructions of Cook Islander, Samoan, and Maori prehistory (Roger Green, personal communication). There is also the problem of demonstrated rate heterogeneity among mtDNA haplogroup lineages (Howell et al., 2004). Even more troubling, it has recently been suggested that all mutation rates show an apparent speed-up in the last 1–2 million years. Ho et al (2005) have called this the “time dependency of molecular rates.” The exact causes of the time dependency of molecular rates are not entirely clear, but it is most likely to be a combination of (i) purifying selection against deleterious and slightly deleterious mutations; and (ii) saturation at mutational ‘hotspots’. Many of the polymorphisms within species do not persist over long time frames because they are removed by purifying (background) selection or by chance (drift). These polymorphisms, however, contribute to the elevated rate estimates made from population-level sequence data, leading to an apparent increase in molecular rates towards the present. (Ho et al., 2005). Of course, this suggested rate revision would apply to all molecular dates for all species within roughly the last million years, which would require major revisions. A promising alternative mtDNA dating scheme, relying solely on synonymous transitions in the coding region, has been proposed in the last few months (Kivisild et al., 2006). An alternative explanation for the apparent recent acceleration is that mutations occurring during a range expansion can be driven to high frequency or even fixation by “surfing” on the wave front of expansion by benefiting from the repeated bottlenecking that characterizes expansions (Edmonds et al., 2004: Klopfstein et al., 2006) . Under expansion conditions the substitution rate can be elevated. For example, the many Island Melanesian variants on the Polynesian Motif may have accumulated faster than otherwise would have been expected. There appear therefore to be three likely dating scenarios (unfortunately not necessarily entirely mutually exclusive). First, if the standard molecular techniques are correct, the oldest set of haplogroups in Near Oceania evolved before the currently accepted time of first settlement of Sahul (40–50,000 YBP), perhaps somewhere in Island Southeast Asia where
they have subsequently been obliterated. Also, the late-arriving mtDNA and NRY variants predate the development of Lapita and would suggest the Austronesian expansion was earlier than supposed. Archeological evidence suggests the Austronesian bearers of a Neolithic tradition did not enter Taiwan until 5,500 YBP, and the Philippines until at least 4,000 YBP. Second, if Ho and company are correct, the molecular dates for both the Upper Pleistocene and Holocene would be expected to be too old by about two-fold. This would, of course, eliminate the disquieting disparities with the archeological dates. Their suggestion is to rely on local calibration points, which would be 3,300–3,500 YBP for those variants that are likely to have appeared with the development of the Lapita Cultural Complex, and 40,000–50,000 YBP for the suite of oldest regional variants. This simply admits the inadequacy of the molecular clock and relies entirely on hopefully appropriate archeological pegs. The third scenario involves the warping of the mutation rate by the dynamics of population expansion, but not its entire rejection (Edmonds et al., 2004; Klopfstein et al., 2006). We might call this the “Salvador Dali” molecular clock. This will be the most interesting to test when the anticipated wealth of unbiased biparental markers becomes available in the near future. Certainly, the situation in Near Oceania has become an interesting example of the problems of the molecular clock (recent and more distant). 3. Austronesian settlement scenarios. Green (2003) tabulated a series of different hypotheses concerning the Austronesian expansion, which has been such a contentious issue. The two extreme hypotheses are the “Fast Train” (Diamond, 1988), which envisions the derivation of Polynesians from a Taiwanese source with almost no Melanesian intermixture, and the “Entangled Bank” (Terrell et al., 2001), which sees the Austronesian expansion into Remote Oceania as essentially an outgrowth of Near Oceanic complexity, with relatively little identifiable Island Southeast Asian (Taiwanese) contribution (see chapter 2). At this stage, the molecular data obviously present conflicting results, with the mtDNA still suggesting something of a fast train, and the NRY suggesting a very slow boat. The molecular genetics data rule out the Entangled Bank, as does historical linguistics (Greenhill and Gray, 2005). Again, the biparental nuclear markers should be most important in clarifying the general ancestral composition of the Polynesians and their proto-Oceanic forbears. The mtDNA and NRY data, while highly informative,
235
regional studies and conclusion
represent only two segments of the immense human genetic vocabulary. I will be surprised if the biparental data end up contradicting the general relationships already established with biometrics: that Polynesians do have a major Island Southeast Asian ancestral component, but with an identifiable fraction of Near Oceanic genetic ancestry as well. 4. General relevance for modern human evolution. With the exception of the genetic variants introduced in the mid to late Holocene from Island Southeast Asia, the diversity described in this book apparently developed within the Southwest Pacific over tens of thousands of years. Relationships with other Eurasian, African, or even Native American variants remain extremely tenuous at this point. See, for example, the contradictory positions of the single small south Nasioi, Bougainville population in the worldwide comparisons of Shriver et al. (2005), Rosenberg et al. (2002), Calafell et al. (1998), or the older study of Bowcock et al. (1987). The most probable candidate ancestral variants are from south Asia, but the phylogenetic signal is very weak (see, for example, the M tree in chapter 4). At this point, we do not even know if the melanin candidate alleles that look to be associated with dark pigmentation in Island Melanesia are the same as those in Africans, or independently evolved in the Pacific. This book does provide a perspective, however, on the extensive diversity that can accumulate in a set of small human populations living in semiisolation for an extended time. This can serve as a model for the levels of diversity in Upper Pleistocene populations elsewhere in Eurasia and Africa. A currently popular reconstruction of the exodus of modern humans from Africa has a single relatively homogeneous population moving rapidly eastward along the coast to the Southwest Pacific continent of Sahul in the period of approximately 60,000–40,000 years ago. It is the simplest hypothesis that fits the currently available data, but its sufficiency does not by any means equate with certainty. There are many reasons to suspect that the small populations expanding in various directions across South Asia and into the Southwest Pacific were not particularly homogeneous, and did not all come simultaneously in one ever-eastward-moving wave. Daniel Nettle (1999) has also made the intriguing argument that greater genetic and linguistic diversity in equatorial populations should be predicted because of their ecological self-sufficiency. The model I prefer is that moving out of Africa into Europe, Asia, the Southwest Pacific, and the
236
Americas represented a sequence of population expansions. The initial populations of each area was a subsample of the last, each expanding and diversifying rapidly in their newly settled regions. An apt analogy would be the non-African continents as a series of empty rooms. Each was in turn colonized by very small groups, followed by periods of rapid expansion and diversification over the course of approximately 10,000 years. That early expansion and diversification is what we have described in this book for the Southwest Pacific.
Acknowledgments Most of the authors of chapters in this book had a chance to read a draft of this Conclusion, and a number responded with helpful suggestions and corrections, so that it is something of a group effort. This book itself has been far more of a collaborative venture than the usual volume of contributed chapters. However, I particularly wish to thank my wife, Françoise, whose admirable analytic and technical skills enabled the consistency of the book, and whose graphic skills have unified its visual expression. I cannot imagine this book developing as it has without her intimate involvement.
References Attenborough R. 2005. Introduction to the chapters on biological anthropology and population genetics. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan Pasts: cultural, linguistic, and biological histories of Papuan-speaking peoples. Canberra: Oceanic Linguistics, pp 673–91. Bellwood P. 1996. Hierarchy, founder ideology and Austronesian expansion. In: Fox JJ, Sather C, editors. Origins, Ancestry and Alliance: Explorations in Austronesian Ethnography. Canberra: Australian National University, pp 18–40. Bowcock AM, Bucci C, Hebert JM, Kidd JR, Kidd KK, Friedlaender JS, Cavalli-Sforza LL. 1987. Study of 47 DNA markers in five populations from four continents. Gene Geography 1: 47–64. Calafell F, Shuster A, Speed WC, Kidd JR, Kidd KK. 1998. Short tandem repeat polymorphism evolution in humans. European Journal of Human Genetics 6: 38–49. Cann RL, Lum JK. 2004. Dispersal ghosts in Oceania. American Journal of Human Biology 16: 440–51. Diamond JM. 1988. Express train to Polynesia. Nature 326: 307–8. Edmonds CA, Lillie AS, Cavalli-Sforza LL. 2004. Mutations arising in the wave front of an expanding population. Proceedings of the National Academy of Sciences USA 101: 975–9.
Conclusion
Friedlaender JS. 1971. Isolation by distance in Bougainville. Proceecdings of the National Academy of Sciences USA 68: 704–7. Friedlaender JS. 1975. Patterns of Human Variation: The Demography, Genetics, and Phenetics of Bougainville Islanders. Cambridge, MA: Harvard University Press. Green RC. 1967. The Immediate Origins of the Polynesians. In: Highland GA, Force RW, Howard A, Kelly M, Sinoto Y, editors. Polynesian Culture History. Honolulu: Bishop Museum. Green RC. 2002. Rediscovering the social aspects of ancestral Oceanic societies through archaeology, linguistics, and ethnology. In: Bedford S, Sand C, Burley D, editors. Fifty Years in the field: Essays in Honour and Celebration of Richard Shutler Jr.’s Archaeological Career. Auckland: New Zealand Archaeological Association. Green RC. 2003. The Lapita horizon and traditions— Signature for one set of Oceanic migrations. In: Sand C, editor. Pacific Archaeology: Assessments and Anniversary of the First Lapita Excavation (July 1952) Koné, Nouméa, 2002. Nouméa, New Caledonia: Le Cahiers de l’Archéologie en Nouvelle-Calédonie, pp 95–120. Greenhill S, Gray RD. 2005. Testing dispersal hypotheses: Pacific settlement, phylogenetic trees and Austronesian languages. In: Mace R, Holden C, Shennan S, editors. The Evolution of Cultural Diversity: Phylogenetic Approaches. London: UCL Press, pp 31–52. Haddon AC, Hornell J. 1975 (reprint). Canoes of Oceania. Honolulu, Hawai’i: Bishop Museum Press. Hage P, Marck JC. 2003. Matrilineality and the Melanesian Origin of Polynesian Y Chromosomes. Current Anthropology 44, Supplement: 121–7. Ho SY, Phillips MJ, Cooper A, Drummond AJ. 2005. Time Dependency of Molecular Rate Estimates and Systematic Overestimation of Recent Divergence Times. Molecular Biology and Evolution 22: 1561–8. Howell N, Elson JL, Turnbull DM, Herrnstadt C. 2004. African Haplogroup L mtDNA sequences show violations of clock-like evolution. Molecular Biology and Evolution 21: 1843–54. Kirch PV, Green RC. 2001. Hawaiki, Ancestral Polynesia. Cambridge: Cambridge University Press. Kivisild T, Shen P, Wall DP, Do B, Sung R, Davis K, Passarino G, Underhill PA, Scharfe C, Torroni A, Scozzari R, Modiano D, Coppa A, de Knijff P, Feldman M, Cavalli-Sforza LL, Oefner PJ. 2006. The role of selection in the evolution of human mitochondrial genomes. Genetics 172: 373–87. Klopfstein S, Currat M, Excoffier L. 2006. The fate of mutations surfing on the wave of a range expansion. Molecular Biology and Evolution 23: 482–90.
Main P, Attenborough R, Gao X. 2005. The origin of the Papuans: the HLA story. In: Pawley A, Attenborough R, Golson J, Hide R, editors. Papuan Pasts: cultural, linguistic, and biological histories of Papuan-speaking peoples. Canberra: Oceanic Linguistics, pp 757–70. Mishmar D, Ruiz-Pesini E, Golik P, Macaulay V, Clark AG, Hosseini S, Brandon M, Easley K, Chen E, Brown MD, Sukernik RI, Olckers A, Wallace DC. 2003. Natural selection shaped regional mtDNA variation in humans. Proceedings of the National Academy of Sciences USA 100: 171–6. Nettle D. 1999. Linguistic Diversity. Oxford: Oxford University Press. Oota H, Settheetham-Ishida W, Tiwawech D, Ishida T, Stoneking M. 2001. Human mtDNA and Y-chromosome variation is correlated with matrilocal versus patrilocal residence. Nature Genetics 29: 20–1. Pawley A. 1982. Rubbish-man, commoner, big-man, chief? Linguistic evicence for hereditary chieftainship in Proto-Oceanic society. In: Siikala J, editor. Oceanic Studies: Essays in Honor of Aarne A. Koskinen. Helsinki: Finnish Anthropological Society, pp 33–52. Pawley AK, Ross HM. 1993. Austronesian historical linguistics and culture history. Annual Review of Anthropology 425–59. Pietrusewsky M. 2005. The physical anthropology of the Pacific, East Asia, and Southeast Asia: A mutlivariate craniometric analysis. In: Sagart L, Blench R, Sanchez-Mazas A, editors. The Peopling of East Asia: Putting together archaeology, linguistics, and genetics. London and New York: Routledge Curzon, pp 201–29. Rosenberg N, Pritchard J, Weber J, Cann H, Kidd K, Zhivotovsky L, Feldman M. 2002. Genetic structure of human populations. Science 298: 2381–5. Shriver MD, Mei R, Parra EJ, Sonpar V, Halder I, Tishkoff SA, Schurr TG, Zhadanov SI, Osipova LP, Brutsaert T, Friedlaender J, Jorde LB, Watkins WS, Bamshad MJ, Gutierez G, Loi H, Matsuzaki H, Kittles RA, Argyropoulos G, Fernandez JR, Akey JM, Jones K. 2005. Large-scale SNP analysis reveals clustered and continuous human genetic variation. Human Genomics 2. Terrell J, Kelly KM, Rainbird P. 2001. “Foregone conclusions.” An analysis of the concepts of ‘Austronesians’ and ‘Papuans’. Current Anthropology 42: 97–124. Trejaut JA, Kivisild T, Loo JH, Lee CL, He CL, Hsu CJ, Li ZY, Lin M. 2005. Traces of Archaic Mitochondrial Lineages Persist in Austronesian-Speaking Formosan Populations. Public Library of Science. Biology 3: e247. Ward RH. 1967. Genetic studies on Fijians. In: Department of Anthropology. Auckland: Auckland University, pp 108.
237
This page intentionally left blank
Index
Abelmoschus manihot, 191 Admiralties, 53 animal introductions, 16–17 language family, 124, 131, 132 See also Manus admixture mapping signals, 103 adoption, 200, 206 Adwe, 26 AFLP See amplified fragment length polymorphism Africa, out-of-Africa expansion, 114 Agostini, HT, 171, 173 agouti signaling protein (ASIP), 103–109 agriculture, 182 and language/population spread, 52 origins of in the Pacific, 185 shift to, 26, 38–39 spread of, 39, 40 See also arboriculture; herbiculture; vegeculture AIDS, 172 Aita (Bougainville, Papuan lg.) indel polymorphism, 117 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143, 146, 149, 213 Y, 90, 91, 92, 93 Akolet (New Britain, Oceanic lg.), 131 albinism, 103 Allaby, R, 181–194 Allen, G, 130 Allen, J, 11, 12, 16, 20, 29, 39 Alor, 36, 47 See also Wallacea α-globin, 213 α-thalassemia, 78, 226, 227, 228 deletions and subtypes associated with malaria resistance, 226 Alu elements, 201 Ambalam (New Guinea, Papuan lg.), 224 ambilocality, 199
Amis, 89 AMOVA of mtDNA variation, 76–77 population genetic analyses, 144, 149–150 of Y variation, 90–91 See also F statistics; Φ statistics; RST amplicon, 113 amplified fragment length polymorphisms (AFLP), 183, 185, 186, 188, 192 Amto-Musian language family, 47 AN See Austronesian languages Analysis of Molecular Variance See AMOVA Analysis of Variance See ANOVA ancestry informative allele (AIM), 103 Aneityum, 226, 227 Anêm (New Britain, Papuan lg.) indel polymorphism, 117 language, 119, 130, 131 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143, 146 Y, 86, 89, 92 animal transfers/introductions/ translocations, 157–167 discussion, 165–167 first, 14–17 Lapita sites and remains, 166 linguistic evidence, 164 Pleistocene, 157 post-Pleistocene introductions, 157–164 See also bandicoot; chicken; cuscus; dog; pig; rats; wallaby Anir, 24 anopheles mosquito, 223 ANOVA by island, 99 by latitude and UVR, 99–100 by neighborhood, 99 of hair pigmentation, 98–100, 106–108 of skin pigmentation, 98–100, 106–108 Anson, D, 25
anthocyanin, 183 Apalo, 26, 160 apical protein-like Xenopus laevis (APXL), 200–206 Arapesh (New Guinea, Papuan lg.), 224 Arawe Islands, 24, 26, 182, 184 arboriculture, 182, 187–189 See also agriculture archeobotany, 182–183 molecular, 183 Arlequin software package, 144, 221 Asaro Valley, 221 ascertainment bias, 92 Asiatic sugar (Saccharum sinense), 190 Asmat-Kamoro language group, 49 Ata (New Britain, Papuan lg.) indel polymorphism, 115, 116, 117 language, 119, 120, 131 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143, 146 Y, 89, 92 Attenborough, R, 234 Atui, 131 Austin, F, 208–216 Australian Aborigines, 65, 68, 208, 212, 213–215 Austronesian languages, 118 homeland, 124 lexicostatistical diversity, 50 models for expansion/colonization, 22–24 See also Oceanic languages Avau, 131 Awyu-Dumut language group, 49 Baegu (Malaita, Oceanic lg.), mtDNA, 66 Bahlo, M, 205 Baining culture, 199–200 population genetics, 200–206
239
Index
Baining (Continued) conclusions, 205–206 polymorphism levels and patterns, 201–202 population differentiation, 202 sex-specific factors, 202–205 Baining family of languages, 45, 119, 120, 131, 199–206 Baining (Kaket) (New Britain, Papuan lg.) indel polymorphism, 115, 117 mtDNA, 66 population characteristics (isolation, etc), 143, 152 Y, 89, 92 Baining (Mali) (New Britain, Papuan lg.) indel polymorphism, 115, 117 mtDNA, 66 population characteristics (isolation, etc), 143, 152 Y, 89, 92 Bali, 88, 89 Bali-Vitu (New Britain, Oceanic lg.), 122 Balof, 161 banana (Musa sp.), 182, 189–190, 193 Australimusa section, 189 domestication, 18, 189, 190 Eumusa section, 189 origins, 189, 190 See also herbiculture band 3 gene 27 bp deletion (B3∆27), 224–225 falciparum resistance, 224 and ovalocytosis, 224 bandicoot (Echymipera kalubu), 16, 157 Baniata See Touo Banoni (Bougainville, Oceanic lg.), 140 Bariai languages, 130 basenji, 160, 161 bats, 11, 12, 15, 18 Bayliss-Smith, T, 26 Bebeli, 131 Bellwood, P, 52, 159, 191 betalain, 183 betel nut (Areca catechu), 21, 191 See also arboriculture bilingualism, and linguistic contact, 129 Bilua (Solomons, Papuan lg.), 120, 130, 132 biogeography, biographical regions, and constraints on population movements, 3–4 See also Bougainville, Greater; Buxton’s Line; island size; Near Oceania; Northern Island Melanesia; Remote Oceania; Sahul; Sundaland; Wallacea; Wallace Line bird-shaped pestles, 19, 20 Bird’s Head, 46, 51, 52, 132 Bischofia javanica, 191 Bismarck Archipelago, 3, 52 colonization, 10–13 languages, 132 See also New Britain; New Hanover; New Ireland; Manus; Mussau BKV, 172 Boetcher, B, 212
240
Boiken (New Guinea, Papuan lg.), 224, 225 Booth, P, 208–216 borrowing, linguistic, 129, 130–131 bottle gourd (Lagenaria siceraria), 165, 183, 193 dual origin of, 192, 193 Bougainville, 3, 36, 53, 119, 211 Greater, 36–37, 135 isolation, 17, 135 languages, 120, 128 and Malaita, 89 and New Ireland, 28 population movements, 4 and West Africa, 109 See also Aita; Buka; Eivo; Nagovisi; Nasioi; Rotokas; Saposa; Simeku; Siwai; Teop; Torau Bowcock, AM, 236 breadfruit (Artocarpus altilis, A. camensis, A. mariannensis), 182, 188 domestication, 18, 188 See also arboriculture Buang Merabak, 11, 14, 20 Buin (Bougainville, Papuan lg.), 120 Buka Island, 120, 160 isolation, 17 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 Bulmer, R, 159 Bulmer, S, 21, 39, 53, 159, 160 Burckella sp., 189 See also arboriculture burning, 39 Buxton’s Line, 223 Calafell, F, 236 candle nut (Aleurites moluccana), 182, 191 See also arboriculture Cann, R, 233 Capell, A, 42, 128 Cartarets, 233 See also Kilinailau Island Cavalli-Sforza, LL, 142, 150 caves, as occupation sites, 11 Cenderawasih Bay, 47, 118 center of dispersal, 46, 50, 123, 124 Central-Eastern Oceanic, 124, 125 See also Central Pacific linkage; Micronesian linkage; Southeast Solomonic linkage; Southern Oceanic linkage Central Pacific linkage, 124 chicken (Gallus gallus) introductions, 164 linguistic evidence, 164 Chimbu Province, 50 Chimbu-Waghi language group, 49 Clark, R, 164 Clarke, A, 165 Cleve, H, 212 climate change, 2, 18, 38 effect
climate change (Continued) on coastlines, 18 on exchange, 18 Holocene, 18 warming trend, 18 Clouse, DA, 48 coalescence, 204, 234 coconut (Cocus nucifera), 182, 187–188, 193 domestication, 182, 187 introduction to Island Melanesia, 17 origin, 184 See also arboriculture colonization, 10–13 origins of colonists, 13 commensal animals See animal transfers/ introductions/translocations Comparative Method (CM), 40–41, 43, 44, 120, 123–125, 127–128 Comrie, BA, 43 Cordia subcordata, 182 correlation linguistic-geographic, 133–134 linguistic spatial-structural, 133–134 Corynocarpus sp., 189 See also arboriculture cotton (Gossypium tomentosum), 191 cowrie shell, 19 Crane, GG, 212 craniometry, 78, 234 crops, See also arboriculture; herbiculture; vegeculture culture, linguistic reconstructions of material culture, 125 cuscus Admiralty cuscus (Spilocuscus kraemeri), 16, 157 Gray Cuscus / Northern Common Cuscus (Phalanger orientalis), 14, 15, 39, 135, 157 See also animal transfers/introductions/ translocations cutnut (Barringtonia), 182, 188 See also arboriculture cycad (Cycas circinalis), 182 cytochrome oxidase 3 gene, 201 cytokine tumor necrosis factor-α polymorphism (TNFP-D), 226 and malaria susceptibility, 226 Czarnecki, J, 171–177 Daniels, C, 190 Daniels, J, 190 Dard, P, 208 dating molecular, 62, 160 radiocarbon, 13, 184 See also glottochronology Dauar Island, 186 decay, lexical, 41–42 De Langhe, E, 190 De Maret, P, 190 Denham, T, 187, 193 dental characteristics, 234 dermatoglyphics, 234 dialect geography, 41
Index
Diamond, JM, 219, 220 diet, 11 dingo, 159, 160 distances geographic, 144, 146–148 linguistic, 144, 147–149 marital migration, 4, 232 matrix, 144 phenotypic, 147–148 population genetic See genetic distance diversity heterozygosity, 92, 201 See also FST; population structure dog (Canis familiaris) Australian dingo, 159, 160 Austronesian association, 53 domestication, 160 introductions, 159–161, 165 linguistic evidence, 164 New Guinea singing dog, 160 pre-Lapita evidence for, 159 domestication, 18, 157–161 centers of, 187 definition of, 161 Dongan, 21 Donohue, M, 54 down-the-line exchange model, 15 Dracontomelon dao, 182 Dunn, M, 118–136, 141–153 Dutton, T, 53 dystrophin gene (DMD44), 200–206 East Bird’s Head language family, 47 Eastern Highlands, Papua New Guinea ENGH, 212 language family, 44 Province, 50 East New Britain language family, 45 East Papuan language phylum, 121, 127–128 East Sepik Province, 223–225 ecological zones, 12 effective population size (Ne), 8, 152, 201 See also sex distinctions Egloff, B, 28 Eivo (Bougainville, Papuan lg.) language, 120 mtDNA, 66 endocytosis, 102 endogamy, 199, 206 See also inbreeding Engan language group, 49 “Entangled Bank” hypothesis, 235 environment, human alteration of, 12 forest clearing, 12, 18 See also burning Erianthus sp., 190 Erromango, 226, 227 “Ethnologue,” 212 ethnophoresy, 14 eumelanin, 101 evolution, tree-like See trees exchange between New Guinea and Bismarcks, 20–21 bird-shaped pestles as evidence for, 19
exchange (Continued) interaction within New Guinea, 19–20 networks, 29 See also interaction; trade Express train to Polynesia Model See Fast (Express) Train model family tree model, 40, 41 Fast (Express) Train model, 22, 165, 219, 235 Fauran, P, 212 Fergusson Island, 28 Fichin Tradition, 21 Fiji, 27, 223 Finisterre-Huon language group, 49 fish, dietary, 26 Flannery, T, 160 Foley, WA, 43, 44, 46, 54, 126 folic acid hypothesis, 96 foraging, 16, 48, 49 broad-spectrum foragers, 38 zones, shift from, 16 Friedlaender, F, 61–78, 81–94, 96–109, 141–153 Friedlaender, JS, 3–9, 12, 53, 61–78, 81–94, 141–153, 171–177, 201, 215, 231–236 frogs, dietary, 12 F statistics genetic diversity, 202, 203, 221–222 hair reflectance, 150 pigmentation SNPs, 105–106, 107 skin reflectance, 150 See also AMOVA Fudenberg, HH, 212 Fullagar, R, 25–26 Gainj, 4 Gajdusek, DC, 208–216 Gapapaiwa (Milne Bay, Oceanic lg.), 140 Gaua, 226, 227 Gazelle Peninsula, 120, 199 Geelvink Bay, 47 language family, 48 See also Cenderawasih Bay genealogical model, 40, 41 gene flow, 223, 233 See also marital migration, distance; migration genetic distance, 144, 146–149, 153 pairwise FST, 76–7, 221 pairwise ΦST, 144 pairwise RST, 144 genetic diversity, 201–202, 231–232 loss of, 220–221 See also AMOVA; per nucleotide heterozygosity; population structure; θ (theta) genetic drift, 202, 206 founder effect, 233 See also isolation genetic heterogeneity, 109 heterozygosity, 92, 201 See also population structure genetic homogeneity, 206
genetic variation, 149–150 within-population, 144, 150 See also AMOVA; ANOVA; heterozygosity Genetree software, 204, 205 genomic scanning techniques, 183 See also AFLP; RAPD; RFLP genotyping, 114 Giles, E, 212, 213, 215 glottochronology, 41 glucose 6-phosphate dehydrogenase (G6PD) deficiency, 226 malaria resistance of, 226 GM haplotypes, 208–216 distribution, 210, 211, 212–213 differences between PAP and AN speakers, 214 GM*A B, 209, 211, 212 GM*A,F B, 209, 211, 212–213 GM*A G, 209, 211, 212 GM*A T, 209, 213 GM*A,X G, 209, 211, 212 relevance for population relationships and past migrations, 213–216 Southwest Pacific sample, 211–212 See also immunoglobulin allotypes Golo Cave, 13 Golson, Jack, 18, 20, 39, 53 Gorecki, P, 20, 158 Goroka (New Guinea highlands), 48, 49 Gosden, C, 12, 15, 26, 29 Goulden, R, 130 grasses, 190, 191 Green, RC, 23, 192, 193, 220, 235 Greenberg, J, 43, 120, 121, 127, 128 Griffiths, RC, 205 Groube, L, 12 group boundaries establishment of, 12 See also interaction barriers; isolation Groves, C, 158 Guadalcanal, 3, 164 Guam, 177 Gulf of Papua, 47 Haberle, SG, 18, 188 Halia (Buka, Oceanic lg.), 131, 132 Halmahera, 36, 158, 159, 162 See also Wallacea Hamilton, G, 205, 206 Hammer, M, 81–94, 199–206 Harries, H., 187, 188 Harrison, SP, 121 Hather, JG, 181 Hattori, S, 45 Hatwell, JN, 174 Hawaiian Ti plant (Cordyline sp.), 165, 182, 190 Hawkes, JG, 181 Hay, AR, 185 Hays, T, 53 herbiculture, 189–190 See also agriculture Herpes simplex virus type 1 (HSV-1), 172 See also viruses
241
Index
Herpes simplex virus type 2 (HSV-2), 172 See also viruses heterozygosity, 92, 201 Hibiscus rosa-sinensis, 191 highlands Papua New Guinea, 52 changes in, 18–19 See also Eastern Highlands; Western Highlands Hinkle, AE, 165 historical linguistics See Comparative Method; Papuan languages HLA, 78, 234 Ho, SY, 62, 235 Hodgson, JA, 61–78 Holm, S, 144 Hornabrook, R, 208–216 horticulture, 182 See also agriculture Howells, WW, 231 Hudson, E, 164 human leukocyte antigens (HLAs), 78, 234 Human papillomavirus type 16 (HPV-16), 171–172 See also viruses Human T-lymphotropic virus type 1 (HTLV-1), 171 See also viruses Hunley, K, 141–153 hunting prey, 12 See also bandicoot; bats; frogs; possum; wallaby Huon Gulf, 53 Huon Peninsula, 38, 55, 56 Hurd, C, 127 identical by descent (IBD), 113 immunoglobulin allotypes, 208–216 KM*1, 212, 213, 216 See also GM haplotypes inbreeding, 206 See also endogamy indel, 9.1 kb, polymorphsims, chromosome 22, 113–117 frequency distributions Island Melanesia, 115–116, 117 world-wide, 114 India, 68 Indian Mulberry (Morinda citrifolia), 191 See also arboriculture Indigenous Bismarck model, 22–23 Indo-European language family, 50, 125 lexicostatistical diversity, 50 Indonesia/Indonesian archipelago, 185 Indo-Pacific language phylum, hypothesis, 43, 121, 127 Indrayana, H, 212 informed consent, 201 insertion/deletion polymorphism See indel polymorphism interaction barriers, 13 See also isolation interaction, Papuan-Austronesian, 53 See also exchange
242
inter-island variation See AMOVA; ANOVA International System of Genetic Nomenclature, 208 irrigation channels, 184 See also Kuk Tea Plantation Irwin, GJ, 159 ISEA See Island Southeast Asia island intervisibility, 3 Island Melanesia, 17, 119, 221, 223 post-Lapita connections, 27 See also Northern Island Melanesia island size effect on genetic variation, 90, 99 See also biogeography Island Southeast Asia, 3 isolation of populations, 6, 141 See also interaction barriers isolation by distance model analytical methods, 144 discussion, 152 materials, 142 results, 146–148 isozyme variation, 185–186, 188–190 Ivuyo, Baiva, 29 Japan, 226 JC Virus (JCV), 172–177, 234 agno-gene deletion, 174 distribution, 173, 174–177 epidemiology, 172 frequency, 172 maximum parsimony tree, 174, 175 methodology, 173–174 migration reconstructions with, 173, 177 mode of infection, 172 See also trees, phylogenetic Jobes, DV, 174 Job’s tears (Coix lacryma-jobi), 191 Kafiavana, 19 Kainantu-Goroka language group, 49 Kairiru (New Guinea, Oceanic lg.), 140, 224 Kaket See Baining (Kaket) Kalam, 159 Kamgot, 160 Kaneko, A, 226 Kapingamarangi, 221 Kapugu (Mussau, Oceanic lg.) indel polymorphism, 115, 117 melanin, 98 Karafet, T, 81–94 karaka (Corynocarpus laevigatus), 189 Kaufman, T, 129 Kaulong (New Britain, Oceanic lg.), 140 kava (Piper methysticum, P. wichmanii), 182, 186–187 domestication, 186–187 See also vegeculture keratinocytes, 102 Keriaka (Bougainville, Papuan lg.), 120 Kilinailau Island, 233 Kilivila (Milne Bay, Oceanic lg.), 140 Kilu Cave, 11, 17
Kimura, M, 152 Kirch, PV, 159, 160 Kofler, A, 213 Koil Island, 28 Koki, G, 4, 61–78, 81–94, 96–109, 141–153 Kokota (Solomons, Oceanic lg.), 140 Kol (New Britain, Papuan lg.) indel polymorphism, 117 language, 45, 120, 126, 131 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 Konua (Bougainville, Papuan lg.), 120 Kosipe, 10, 12 Kovai, 130 Kove (New Britain, Oceanic lg.) indel polymorphism, 117 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 86, 89 Kowekau Cave, 17 Kuk Swamp, 184, 185, 187, 188–189 Kuk Tea Plantation, 18–19 and highlands agriculture, 18, 50 kula exchange system, 28 Kumar, S, 212 Kuot (New Ireland, Papuan lg.) indel polymorphism, 115, 117 language, 45, 120, 126, 127, 130, 131, 132 melanin, 97, 98 mtDNA, 66 population characteristics (isolation, etc), 143, 152 Y, 89, 92 Kupona na Dari, 11, 15 Kwaio (Malaita), mtDNA, 66 Kwomtari language family, 47 Lachitu, 21 Lakes Plain languages, 48 Lamasong, 130 Lang, R, 44 Lang Rongrien cave, 13 language barriers, effects on marriage rates/migration, 5 borrowing, 129, 130–131 change, 120 language shift, 130 sound change, 120 contact, 129–134 distributions, relevance, 5 structural features of, 128, 130, 134 typological sketch, 121–123, 126–127 Lapita Cultural Complex contact with Papuan speakers, 40 dates, 235 models for, 22–24, 220 and obsidian, 25–26 origins, 22–24 periods of, 21
Index
Lapita Cultural Complex (Continued) pottery, 21, 23, 24–25 and Proto-Oceanic, 23 regional interaction, 40 settlement patterns, 39, 40 and settlement of Remote Oceania, 21 site locations, 24 and economy, 26 See also Fast (Express) Train model; Indigenous Bismarck model; Slow Train (Slow Boat) model; Triple I model Latham, K, 81–94 Lau (Malaita, Oceanic lg.), mtDNA, 66 Lavongai (New Hanover, Oceanic lg.) indel polymorphism, 115, 117 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92, 93 See also New Hanover Lavukaleve (Solomons, Papuan lg.), 120, 126, 127, 128, 130, 132 Laycock, DC, 47 Left May language family, 47 Lene Hara (East Timor), 13 Lepofsky, D, 182 Lesser Sundas, 36 See also Wallacea Levinson, S, 130 lexicostatistics, 41 Liang Lemdubu, 13 Lilley, I, 159 Lindström, E, 118–136, 141–153 linguistic diversity, 6, 44, 46, 129 causes, 44 linguistic networks, 124 innovation-defined, 124 innovation-linked, 124 linguistic residues, detecting ancient relationships, 40 Long, JC, 4, 152 Lorenz, J, 81–94 Loso (New Britain, Oceanic lg.) indel polymorphism, 117 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 Louisiade Archipelago, 45 Lower Sepik languages, 44, 46 Lower Sepik-Ramu, 46 Lum, JK, 219–228, 233 Lynch, J, 132, 134, 164, 186, 191 M Index, 97 hair, 98, 100, 106–108 skin, 98, 101, 102, 106–108 Madak (New Ireland, Oceanic lg.) indel polymorphism, 117 language, 130 melanin, 98 mtDNA, 66
Madak (New Ireland, Oceanic lg.) (Continued) population characteristics (isolation, etc), 143 Y, 92 Madang language group, 48–49 Madang Province, 46, 50, 213 Main, P, 234 Maisin, 129 Makira, 36, 164 See also Solomon Islands Malaita, 70, 74 and Bougainville, 89 See also Baegu; Kwaio; Lau Malakula Island, 226 malaria, 219, 223–228 epidemiology of, 223 genes conferring resistance to (B3∆27), 224 genetic susceptibility to, 224 parasite drug resistance, differential selection by, 225, 226 parasites (Plasmodium falciparum, P. vivax), 223, 224 selection within PNG, 223–225 selection within Vanuatu, 225–226 Malay Apple (Syzygium malaccense), 189, 191 See also arboriculture Malayo-Polynesian See Austronesian Malécot, G, 152 male vs. female distinctions See sex distinctions Mali See Baining (Mali) Maluku Utara, 159 Mamusi (New Britain, Oceanic lg.) indel polymorphism, 115, 116, 117 language, 119–120 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 89, 92 Mangaia Island, 192 Mangap-Mbula, 130 mangosteen (Garcinia sp.), 189 See also arboriculture Mangseng (New Britain, Oceanic lg.) indel polymorphism, 117 language, 119, 122 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 Manus (Admiralties, Oceanic lg.) mtDNA, 66 population characteristics (isolation, etc), 143 Y, 86 Manus Island, 187 and New Guinea, 28 settlement, implications for water transport, 13, 16 Maori, 162, 189 Marambu, 201–206 See also Baining (Mali)
marital migration bush vs. beach, 4 distance, 4, 232 rate, 4 marital residency patterns, See ambilocality; matrilocality; patrilocality Markham Valley, 97, 211, 213 marriage, 199 marsupial megafauna, 12 See also hunting prey Mass Comparison method, 120–121, 127 See also Greenberg, J Matenkupkum Cave, 11, 14 Matisoo-Smith, E, 147–157 matrilocality, 205, 233 McAdams, K, 208–216 McGrath, S, 61–78 Melamela (New Britain, Oceanic lg.) indel polymorphism, 117 language, 120 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 See also Meramera Melanesia, 219 See also Central–Eastern Melanesia; Island Melanesia; Northern Island Melanesia melanin, 97 biosynthesis, 100–103 Melanin Index See M Index melanocytes, 101 melanosomes, 102 Mengen (New Britain, Oceanic lg.) indel polymorphism, 115, 117 language, 120, 122, 123, 131 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 Meramera, 123 See also Melamela Merriwether, DA, 61–78, 141–153 Meso-Melanesian linkage, 119, 124 St. George linkage, 120, 131 Willaumez linkage, 119 metatypy, 129 Metrosideros sp., 189 See also arboriculture Micronesian language family, 124 Central-Eastern, 221 microsatellites age estimates using, 90 markers, 85 NRY, 90, 91 and plant domestication, 194 See also STR loci migration effects on genetic variation, 153, 202, 206 linguistic reconstructions of, 120, 125 multiple, 124–125
243
Index
migration (Continued) rates, 152, 202–203 secondary migrations in Island Melanesia, 17 migration parameter (m), 204, 205 Milne Bay, 136, 213 Minett, JW, 132 Miscanthus sp., 190 Misima mtDNA, 66, 71 Y, 81 mitochondrial DNA See mtDNA mitochondrial sequence variation, 142 molecular clock, 235 molecular dates and problems with, 234–235 molecular rate, time dependency of, 235 Moluccas, 53 See also Wallacea Mondol, Robert, 29 monogamy, 199 Mopir, obsidian source, 15, 20, 34 Morobe Province, 50, 211, 213 morphology, linguistic, 125, 126, 127 Motu Hiri, 28 Motuna See Siwai mtDNA, 61–78, 233–235 Baining, 200–206 dating and issues, 62 haplogroups B, 214, 215 B4a1a1 (“Polynesian Motif”) and precursor, 72–75, 94, 177 B4b, 76 defining mutations, 69 E, 75–76 E1a, 75–76 E1b, 75, 76 E2, 75–76 introduced from Southeast Asia, 72 M, 68, 69, 70, 71, 215 M7, 76 M27, 70, 71, 135, 214 M28, 70–71, 135, 214 M29, 71–72, 214 oldest in Near Oceania, 65–72 P, 65, 68, 214, 215 P1, 65 P2, 65 P3, 65 P4, 65 Q, 70, 71, 214, 215 Q1, 70 Q2, 12, 70 Q3, 70 Y, 76 and inland Papuan isolation, 78 lineage frequencies, 64, 65, 66–67 links to Australian Aborigines, 65, 68, 214 MDS plots, 77, 150, 151 Northern Island Melanesia population comparisons, 76–77 phylogeny in Near Oceania and Island Melanesia, 63–76
244
mtDNA (Continued) plant, 183 regional haplotype distributions, 63 region V deletion, 224, 225 Taiwan connection, 72, 74 Mt. Witori eruption, 25 effects of, 20 multidimensional scaling (MDS), 76–77, 91–92, 144, 150–151 Multilateral Comparison See Mass Comparison method Mussau, 24, 93, 159 mtDNA, 66 plant exploitation, 182, 184, 187 population characteristics (isolation, etc), 143 mutation rate, 235 mtDNA, 62 Y chromosome, 81 See also molecular clock, molecular rate Myanmar, 226 Nagaoka, L, 161 Nagovisi (Bougainville, Papuan lg.) language, 120 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 Nakanai (New Britain, Oceanic lg.) indel polymorphism, 115, 117 language, 119, 123 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 Nakhleh, L, 132 Nalik (New Ireland, Oceanic lg.) indel polymorphism, 117 language, 140 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 NAN See Papuan languages Nasioi (Bougainville, Papuan lg.) indel polymorphism, 115, 116, 117 language, 120, 126 melanin, 109, 236 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 89 natural selection, 99 differential, 96, 99, 108 effects of, 7 See also malaria, selection within PNG, selection within Vanuatu; sexual selection Ndu languages, 47 Near Oceania, 3, 6, 7, 9, 36–37, 219 as plant domestication center, 183–190 prehistory, 37–40
Near Oceania (Continued) settlement of, 51 small population sizes in, 38 Nehan, 132 Neighbor Joining (NJ) trees, 142, 143–146, 163 Nei, M, 144, 212 Nettle, D, 236 New Britain, 52, 53, 199 languages, 119–120 and New Guinea, 28 and New Ireland, 28 population movements, 4 See also Anêm; Ata; Baining; Bismarck Archipelago; Kol; Kove; Mamusi; Mangseng; Melamela; Sulka; Tolai; Yombon New Caledonia, 27, 165–167 New Guinea colonization, 10–13 languages, 36, 46–47 and Manus, 28 and New Britain, 28 New Guinea highlands, 4, 88 See also Papua New Guinea New Guinea lizard (Lipinia noctua), 165 New Hanover, 16, 89, 99 See also Lavongai New Ireland, 53 animal introductions, 14–16 and Bougainville, 28 languages, 120 and New Britain, 28 See also Kuot; Madak; Nalik; Notsi; Tigak Nggela, 36 See also Solomon Islands Niah Cave, 13, 30 Nichols, J, 134, 136 Nissan, 17, 28, 34 Ni Vanuatu, 221, 222 See also Vanuatu Nombe, 10, 12, 18, 31 non-Austronesian languages See Papuan languages Nordlund, J, 101–102 North America, 152 North Bougainville language family, 45, 120 Northern Island Melanesia, 3, 36–37, 52, 53 prehistory, 39–40, 152–153 northern pademelon (Thylogale browni), 158 See also wallaby North New Guinea linkage (NNG), 119, 124, 131, 132 Northwest Solomonic linkage, 120, 125 Norton, H, 96–109, 141–153 Notsi (New Ireland, Oceanic lg.) indel polymorphism, 115, 117 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 90, 92
Index
NRY lineages C2b-M208, 87–88 C2-M38, 86–87 C4-P55, 88 C-RPS4Y, 85–86 frequencies, 82, 83–84 K1-M177, 89 K5-M230, 88–89 K6-P79, 89 K7-P117, 89 K-M9*, 88 M2*-P87, 89 M2a-P22, 89 M-M4*, 89 O1a-M119, 89 O2a, 89 O3-M122, 89 O-M175, 89 NRY structure, 81–90, 233–235 age estimates, 90, 91 AMOVA, 90–91 findings, 82–85 heterogeneity, 90 MDS plots, 93, 150, 151 SNP analysis, 81–82 SNP haplogroup distributions, 85–90 tree, 82 See also NRY lineages obsidian distribution, 25–26 Mopir, 15, 20, 34 source, 19, 25 stemmed tools, 25 translocation of, 14–16 Oceanic languages, 121–125 contact with Papuan languages, 129–134 distribution, 119 history reconstruction, 123–125 subgroups, 124 typological sketch, 121–123 See also Admiralties; Austronesian languages; Western Oceanic; Central-Eastern Oceanic Ok language group, 49 Osmond, Meredith, 44 outrigger canoes with sails, 39 ovalocytosis See band 3 gene 27 bp deletion P See Papuan languages Pacific rat See rats Pakistan, 205–206 Paleo-Indians, 208 Pamwak Cave, 16, 161 Panakiwuk, 161 Pantar, 36, 47 PAP See Papuan languages paper mulberry (Broussenetia papyrifera), 191 Papua New Guinea, 205–206 coast of, 27–28 See also New Guinea
Papuan languages (PAP), 36, 119, 125–129, 212 contact with Oceanic languages, 129–134 distribution, 119 East Papuan, 44, 45 historical linguistics, 42–44, 127–129 New Guinea families, 36, 46–47 typological sketch, 126–127 Papuan Tip linkage, 124, 131, 132 Parkinson, R., 28 parthenocarpy, 182, 189 Pasveer, J, 12 patrilocality, 205, 233 Pavesi, A, 174 Pawley, A, 36–54, 124 pedigree, 113 Pentecost, 226, 227 permission See informed consent per nucleotide heterozygosity, 201 phagocytosis, 102 pheomelanin, 101, 102 Philippines, 159, 235 Φ statistics, 92, 149–150 See also AMOVA phonology, 121, 126, 130 See also language, typological sketch photo-protection hypothesis, 96 Phylip, 144 phylogenetics, 183 structural linguistic, 121, 127–128, 132 phylogeography, 63, 183 phylum, 42 phytoliths, 182–183 π (pi), 201 Pietrusewsky, M, 234 pig (Sus scrofa, S. celebensis, S. barbatus, S. verrucosus), 39 Austronesian association, 53 bones of, 20, 158 controversy on Lapita association, 20, 159 as evidence for domestication, 158 introductions, 20, 157–159, 165 linguistic evidence, 164 pre-Lapita evidence of, 39, 53 pigmentation, 96–109 correlation with latitude or UVR, 99–100 genetics of, 100–108 genotype–phenotype associations, 106–108 hair, 98–100, 106–108, 153 phenotypic measurements, 97 skin, 98–100, 102, 106–108, 153 subject classification, 97 variation by island and neighborhood, 99 by language phylum, 97–99 by sex, 97 See also melanin; M index pigmentation candidate genes ASIP, 103–109 MATP, 103–105, 109 MC1R, 103–108 OCA2, 103–109
pigmentation candidate genes (Continued) TYR, 103–105, 109 TYRP1, 103–104 plant domesticates, 181–182, 192–193 origins of, 181, 192–193 See also banana; breadfruit; sugarcane; taro; yams plant phytochemicals, 183 plant transfers/introductions/translocations, 181–194 primary introductions, 184–190 secondary introductions, 190–191 tertiary introductions, 191–193 See also agriculture; arboriculture; herbiculture; horticulture; vegeculture Pleistocene animal translocations, 157 life in, 11–12 Pleistocene sites, See Buang Merabak; Huon; Kilu; Kosipe; Kupona na Dari; Lachitu; Liang Lemdubu; Matenkupkum; Nombe; Toe Cave; Yombon PMP See Proto-Malayo-Polynesian PNG See Papua New Guinea POc See Proto-Oceanic Pollen analysis, 39 polygyny, 199, 204 polymerase chain reaction (PCR), 173 polymorphisms, 201–202 See also genetic variation; indel polymorphism Polynesian Motif (B4a1a1), 72–75, 94, 177, 231, 233 frequency, 73 hypothetical relation to Lapita, 73 origin, 75 Polynesian origins, 75 Polynesian Outliers, 27 Polynesian tomato (Solanum repandum), 192 Pometia, 182 See also arboriculture population displacements See population movements population expansion See coalescence population fissions model analytical methods, 143–144 discussion, 150–152 materials, 142 results, 146 population growth parameter (β), 204 population history fissions model for See population fissions model isolation by distance model for See isolation by distance model population increase, 38 effects of, 15–16 population isolation, 6, 141 See also genetic drift; marital migration population movements/mobility, 5 constraints on spread, 174 dynamics, 4–5 increase over time, 5 population-mutation parameter (θ), 201, 204
245
Index
population stratification See population structure population structure, 199–206 possessives, 122–123, 131 See also language, typological sketch possum, 12, 32 pottery Fichin Tradition, 21 introduction of, 20–21 Lapita dentate-stamped, 21, 23 post-Lapita, 27, 28 pre-Lapita, 20–21, 53 relation to Lapita, 24–25 predation levels, 12 primer, 113, 115 Principal Components Analysis (PCA), 215–216 progressive multifocal leukoencephalopathy (PML), 173 proportion of shared sites (PSS), 202 proto-languages, 41, 124 proto-linkages, 124 Proto-Malayo-Polynesian (PMP), 123 Proto-Oceanic (POc), 23, 118, 121, 123, 134 linguistic reconstruction, 125, 184 questionnaire, kinship, 4 RST, 144 See also AMOVA R271 See indel polymorphism, 9.1 kb, chromosome 22 Ramu languages, 46 Ramu River, 6, 18, 52 random amplified polymorphic DNA (RAPD), 183, 185, 192 Rangulit, 201–206 See also Baining (Kaket) Rathje, WL, 25 rats Asian rat (Rattus tanezumi), 164 commensal species, 162 introductions, 161–164 large spiny (Rattus praetor), 15, 16, 157, 161, 164 linguistic evidence, 164 Melomys rat (Melomys rufenscens), 157 Norwegian rat (Rattus norvegicus), 161, 162 Pacific rat (Rattus exulans), 161, 162, 164, 167 (Rattus mordax), 157, 161 (Rattus rattus), 161, 162 See also animal transfers/introductions/ translocations RDA, 113 red hair color (RHC), 103 Reesink, G, 118–136, 141–153 Relethford, J, 144 Remote Oceania, 3, 21, 36–37, 162, 219 Renfrew, C, 52 Rensch, KH, 192 Representational Difference Analysis (RDA), 113 reproductive success, 203–204 residual zones, 134
246
resource distribution, 16 See also foraging, zones restriction fragment length polymorphisms (RFLP), 186, 187, 226 rice (Oryza sativa ssp.), 191 RMET, 144 Roberts-Thompson, JM, 213 Robledo, R, 113–117 Rogers, A, 109 root crops, processing Alocasia, 17, 18 Cyrtosperma, 17 Roscoe, P, 39 Rosenberg, N, 236 Rossel Island, 76, 134 Ross, MD, 44, 45, 47, 49, 50, 51, 121, 124, 125, 127, 129, 130, 132, 135, 136 Rotokas (Bougainville, Papuan lg.) language, 120, 126 mtDNA, 66 population characteristics (isolation, etc), 143, 213 Roviana (Solomons, Oceanic lg.), 140 Saave, J, 208–216 sago palm (Metroxylon sagu, M. rumphii), 189, 190, 194 See also arboriculture Sahul, 3, 36–37, 183, 209 Pleistocene settlement of, 38, 209, 236 routes to enter, 13 Sahul-Armature shelf, 38 Samoa, 187, 221 sampling strategy, 93, 203 sandalwood (Santalum), 182 See also arboriculture Sandaun Province, 47 Santa Cruz, 182 Austronesian language, 45 settlement of, 45 Santa Ysabel (Ysabel), 164 Santo, 226, 227 Saposa (Bougainville, Papuan lg.) indel polymorphism, 117 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143, 146 Y, 92 Sardinia, 113 Sauer, CO, 181, 193 Savolainen, P, 160 Savosavo (Solomons, Papuan lg.), 120, 132 Schanfield, M, 208–216 Scheinfeldt, L, 81–94, 141–153 Schurr, TG, 61–78 screw pine (Pandanus antaresensis, P. brosimos, P. julianettii, P. spurious, P. tectorious, P. whitmeanus), 182, 188–189 domestication, 182, 188 See also arboriculture sea almond (Terminilia catappa, T. copelandii, T. kaernbachii), 182, 189 See also arboriculture
seafaring, 13, 125, 130, 135 See also outrigger canoes with sails sea levels, 10, 13, 18, 36 Second World War, effects on demography, 4 segregation analysis, 113 Senft, G, 136 Sepik languages See Lower Sepik-Ramu; Ramu languages; Sepik-Ramu language group; Yuat Sepik Plain, 224 Sepik-Ramu basin, 43, 51, 187, 191 Sepik-Ramu language group, 43, 46 sequential-rejective approach, 144 Serjeantson, S, 208–216 sex distinctions in effective population sizes, 205 in marital migration rates, 4 in melanization, 97 in reproductive success, 203–204 in reproductive variance, 204 sexual selection, 96, 100, 112 Sharp, PM, 174 shore resources (diet), 38 Shortland Islands, 164 short tandem repeat loci, see STR loci Shriver, MD, 236 Siar (New Ireland, Oceanic lg.), 140 Simeku (Bougainville, Papuan lg), mtDNA, 66 Simian virus 40 (SV40), 172 single nucleotide polymorphisms See SNPs Sisiqa (Solomons, Oceanic lg.), 140 Siwai (Bougainville, Papuan lg.) language, 120 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 Sko language family, 47 Slatkin, M, 144 Slow Train (Slow Boat) model, 23–24 SNPs, 81 non-synonymous, 103 recurrence, 81 synonymous, 103 See also NRY structure Sohks, 29 Solomon Islands, 3, 53, 119 languages, 120 mtDNA, 66 Papuan language family, 53 sound correspondence, regular, 40–41 South Africa, 213 South America, 142 South Bougainville language family, 45, 120 Southeast Asia, earliest occupation, 13 Southeast Papuan language group, 49 Southeast Solomonic language family, 124, 125 Southern Oceanic linkage, 124 Southwest New Britain linkage, 119 Specht, J, 16, 17, 29 spread zones, 134 Spriggs, M, 27, 39, 54, 159 SPSS software package, 91, 144
Index
starch grain analysis, 182, 185 Statistica, 215 Stavely, JD, 212 Stebbins, T, 54, 119 Steinberg, AG, 211, 212, 215 St. George linkage, 120, 131 St. Matthias See Mussau Stoltz, M, 61–78 Stoner, G, 171–177 storage organ crops, 181 Storey, AA, 164 STR loci, 220 See also microsatellites structural linguistic features, 118, 143 trees See phylogenetics, structural subsistence base, 11 See also hunting prey Sudest (Milne Bay, Oceanic lg.), 140 sugar cane (Saccharum sp.), 182, 190 domestication, 18, 190 origins, 183, 184, 190 See also herbiculture Sugimoto, C, 173, 174 Sulawesi, 36 See also Wallacea Sulka (New Britain, Papuan lg.) indel polymorphism, 117 language, 45, 120, 126, 130, 131, 132 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 Summerhayes, GR, 10–29, 54 Sundaland, 3, 36 Swadling, P, 19, 21, 29, 54, 158 sweet almond (Canarium ovatum, C. vulgare, C. lamii, C. harveyii, C. decamanum, C. indicum, C. baileyanum, C. littorale, C. zeylanicum, C. solomonense), 16, 17, 182, 187, 193 See also arboriculture sweet potato (Ipomoea batatas), 182, 183, 192–193 domestication, 192 origins, 165, 183, 192, 193 See also root crops syntax, 123, 129, 131 See also language, typological sketch Tahitian chestnut (Inocarpus sp.), 191 See also arboriculture Taiof (Bougainville, Oceanic lg.), 140 Taiwan, 3, 124, 177, 235 Aborigines, 89, 177 See also Amis Tajima’s D, 202 Takia (New Guinea, Oceanic lg.), 129 Tamura, K, 144 Taora, 21 Tarophagus proserpina, 185 taros (Colocasia esculenta, Cyrtosperma chamissonis, Alocasia macrorrhizus), 165, 182, 183, 185 dasheen, 185 domestication, 184
taros (Continued) in Near Oceania, 18, 52, 185 in Southeast Asia, 185 eddoe, 185 introduction to Island Melanesia, 17 origins, 183, 184–185 triploid, 185 tattooing, 25 Teop (Bougainville, Oceanic lg.) indel polymorphism, 117 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143, 146 Y, 87, 92 Terrill, A, 118–136, 141–153 territoriality, 16 territorial ranges, 16 Thailand, 161 Thomason, SG, 129 Thurston, WR, 130 Tigak (New Ireland, Oceanic lg.) indel polymorphism, 115, 117 melanin, 98 mtDNA, 66 population characteristics (isolation, etc), 143 Y, 92 Tikopia, 160, 164 Timor, 36, 47, 50, 53 dogs in, 159 See also Wallacea TMRCA, 203 See also coalescence TNFP alleles, 226 TNG hypothesis See Trans New Guinea language family Toe Cave, 11 Tolai (New Britain, Oceanic lg.) indel polymorphism, 117 language, 120, 123 mtDNA, 66 population characteristics (isolation, etc), 143, 152 Y, 86, 89, 92 Tomoip (New Britain, Oceanic lg.), 120 Tonga, 164 Torau (Bougainville, Oceanic lg.) melanin, 98 mtDNA, 66 Torrence, R, 16 Torres Straits, 38 Torricelli language family, 47, 132 Touo (Solomons, Papuan lg.), 120, 126–127, 130, 132 trade, 152 transformation, in societies, 13 Trans New Guinea language family/phylum (TNGP), 42–44, 47–53, 134, 212 dispersal dating, 50 distribution, 212 divisions, 212 grouping, 48–49 location of proto-TNG, 49–50 TNG hypothesis, 42–43, 47 critiques of case for, 43–44
tree crops, 219 See also arboriculture; banana; coconut; screw pine; sweet almond trees fitting, 146, 152 genetic, 143–146, 204 haplogroup, 204 linguistic, 123–124, 143–146 Neighbor Joining (NJ), 142, 143–146, 163 phylogenetic, 128–129, 173, 174 treeness, 146, 152 Trejaut, JA, 73 tripartite hypothesis, 192, 193 Triple I model, 23, 220 Trobriand Islands, 221 Tungag (New Ireland, Oceanic lg.), 140 turmeric (Curcuma sp.), 191 See also root crops typology, linguistic See language, structural features of; language, typological sketch ultraviolet radiation See UVR uniparental inheritance, 200, 206 Upper Wahgi Valley, 38–39 See also Kuk Tea Plantation Utupua, 124 UVMED levels, 99, 101 UVR, 96 and melanogenesis, 102 Vanikoro, 124 van Loghem, E, 208 Vanuatu, 27, 165–167, 221–223, 225–226 as kava domestication origin, 186–187 malaria selection within, 225–226 populations, 221 settlement by Lapita peoples, 221, 225 See also Ni Vanautu Varicella zoster virus (VZV), 172 See also viruses vegeculture, 181–182, 184–187 See also agriculture; storage organ crops Vilà, C, 160 viruses, 171–177 See also BKV; Herpes simplex virus type 1 (HSV-1); Herpes simplex virus type (HSV-2); Human papillomavirus type 16 (HPV-16); Human T-lymphotropic virus type 1 (HTLV-1); JC virus (JCV); Simian virus 40; Varicella zoster virus (VZV) vitamin D hypothesis, 96 Vitiaz Strait, 13 volcano eruptions, 10, 15, 17, 18, 44 See also Mt. Witori Voorhoeve, CL, 49 waisted axes, 10, 12 wallaby (Thylogale browni, Dorcopsis), 17 animal transport, 20, 158 distribution of, 21
247
Index
wallaby (Continued) as prey, 12 See also northern pademelon (Thylogale browni) Wallacea, 37, 63 Wallace Line, 63, 183 Wanelek, 21 Wang, WS-Y, 132 warfare, effects of, 44 Warnow, T, 132 Watom Island, 164 Watterson, GA, 201 wax gourd (Benincasa hispida), 193 Webb, J, 26 Weiss, G, 152 West Bomberai-Timor language group, 49 Western Highlands Province, Papua New Guinea, 50 Western New Guinea language group, 49 Western Oceanic, 124, 125, 131 See also Meso-Melanesian linkage; North New Guinea linkage; Papuan Tip linkage
248
West Irian, 47 West New Britain language family, 45 West Papuan language family, 47 Wetef Cave, 13 whalers, 29 White, JP, 17, 159, 161 Wi Apple (Spondias dulcis), 191 See also arboriculture Wide Bay, New Britain, 120 Wigler, M, 113 Wilder, J, 199–206 wild ginger (Zingiber zerumbet), 186, 187, 191 wild rice (Oriza nivara; O. rufipogen), 191 Willaumez linkage, 119 Willaumez Peninsula, 118 obsidian source, 15, 20 Wise, R, 185 Woodfield, G, 208–216 Wright, S, 152, 202, 204 Wurm, SA, 20, 42, 43–44, 45, 121, 127–128, 134, 135, 136, 212
Yabem (New Guinea, Oceanic lg.), 140 yams (Dioscorea esculenta, D. alata, D. bulbifera, D. nummularia, D. pentaphylla, D. hastifolia, D. transversa, D. alata, D. sativa), 182, 183, 186, 193 distribution, 186 introduction to Island Melanesia, 17 New Guinea origin, 26, 186 Yanagihara, R, 176 Y chromosome, 81, 153 markers, 85 See also NRY lineages; NRY structure Yélî Dnye (Rossel Island, Papuan lg.), 45, 121, 126, 128, 130, 131 position, 135–136 Yen, D, 17, 160, 165, 182, 188, 189, 192, 193 Yombon, 20 occupation, 11 YSTR See microsatellites Yuat language group, 46 Yuku, 19, 20, 158
X-linked genes, 200–201
Zhivotovsky, LA, 205