You are on page 1of 99

SPRINGER BRIEFS IN ARCHAEOLOGY

Emily Lena Jones

In Search of the
Broad Spectrum
Revolution
in Paleolithic
Southwest Europe
123

SpringerBriefs in Archaeology

More information about this series at http://www.springer.com/series/10186

Emily Lena Jones

In Search of the Broad


Spectrum Revolution
in Paleolithic Southwest
Europe

Emily Lena Jones


Department of Anthropology
University of New Mexico
Albuquerque, NM, USA

ISSN 1861-6623
ISSN 2192-4910 (electronic)
SpringerBriefs in Archaeology
ISBN 978-3-319-22350-6
ISBN 978-3-319-22351-3 (eBook)
DOI 10.1007/978-3-319-22351-3
Library of Congress Control Number: 2015949633
Springer Cham Heidelberg New York Dordrecht London
The Author(s) 2016
This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, express or implied, with respect to the material contained herein or for any errors
or omissions that may have been made.
Printed on acid-free paper
Springer International Publishing AG Switzerland is part of Springer Science+Business Media
(www.springer.com)

Acknowledgments

The research presented in this volume has its roots in my dissertation, and thus my
first thanks go to my Ph.D. advisors, Donald K. Grayson and Eric Alden Smith, and
to Franoise Delpech, my mentor at the Universit Bordeaux I. In recent years,
Lawrence Guy Straus has been a mentor as well as a colleague; the portion of this
work that focuses on Iberia could not have been completed without him. Thank you
all for being such superlative guides and role models.
This project involved tracking down large numbers of datasets, and many individuals helped me to identify and locate sources, including J. Emili Aura, Gregory
Bayle, Carmen Cacho Quesada, David Cochard, Iain Davidson, Jess Jord,
Vronique Laroulandie, Ana Belen Marn Arroyo, Andr Morala, Ana Navarro,
Manuel Perez, Manuel Ramos, Sergio Ripoll Lopez, Alfred Sanchis Serra, and Jos
Yravedra. Patricia Crown and Ann Ramenofsky convinced me to present these data
in a longer format; Stephanie Mack assisted with the data compilation; and Caitlin
Ainsworth, Cyler Conrad, David Hurley, Lawrence Straus, and William Taylor read
and commented on chapter drafts.
Finally, I am grateful to Teresa Krauss, Springer Senior Editor of Archaeology
and Anthropology, who helped me see my way toward how best to present these
analyses, and to Hana Nagdimov for her expert assistance in the manuscript submission process.
My most sincere thanks to all of you. Any errors, of course, are my own.
Elevation maps and data in this work were produced using Copernicus data and
information funded by the European Union (EU-DEM layers). Portions of the
research reported here were supported by the Graduate School of the University of
Washington and the Office of Research at the University of New Mexico, by a
Chateaubriand Fellowship from the Mission pour la Science et la Technologie of the
Embassy of France in the United States and by the National Science Foundation
under Grants No. 947900 and 1148146.

Contents

Paleolithic People, Paleolithic Landscapes ...............................................


1.1 Introduction ..........................................................................................
1.2 Why Paleolithic Subsistence Choices? ................................................
1.3 Why Southwest Europe? ......................................................................
1.4 The Late Paleolithic of Southwest Europe...........................................
1.5 Organization of This Volume ...............................................................
References .....................................................................................................

1
1
2
3
5
7
7

Big Game, Small Game: Why It Matters..................................................


2.1 Introduction ..........................................................................................
2.2 Historical Examples of Bigger Is Better...........................................
2.2.1 A Gap or a Mesolithic? ............................................................
2.3 The Broad Spectrum Revolution ......................................................
2.4 The Prey Choice Model and the Body-Size Proxy ..............................
2.5 Alternatives to Traditional Applications of the Prey
Choice Model .......................................................................................
2.5.1 Prey Mobility and Energetic Return ........................................
2.5.2 Increasing Hunting Efficiency Through Innovation ................
2.5.3 Gender, Age, and Prey Rank ....................................................
2.5.4 Niche Construction Theory ......................................................
2.6 Is Bigger Always Better? .....................................................................
References .....................................................................................................

9
9
10
11
12
13

Climate and Environment in Late Paleolithic Southwest Europe ..........


3.1 Introduction ..........................................................................................
3.2 Climate .................................................................................................
3.2.1 Ice Cores and Oxygen Isotopes................................................
3.2.2 Marine Cores............................................................................
3.2.3 Fossil Pollen .............................................................................
3.2.4 Climate in Late Paleolithic Southwest Europe ........................

23
23
24
24
26
26
26

14
14
15
15
17
18
18

vii

viii

Contents

3.3 Environment and Landscape ................................................................


3.3.1 The Iberian Peninsula ..............................................................
3.3.2 The Pyrenees ............................................................................
3.3.3 Southern France .......................................................................
3.4 Regionalization and Paleolithic People................................................
References .....................................................................................................

27
27
30
30
31
31

Human Subsistence and the Archaeofaunal Record


of Late Paleolithic Southwest Europe .......................................................
4.1 Introduction ..........................................................................................
4.2 Site Location and Human Subsistence.................................................
4.3 Regionalization and Archaeofaunas.....................................................
4.3.1 Similarity Indices and Cluster Analysis...................................
4.3.2 Archaeofaunal Similarity in the Upper Paleolithic ..................
4.3.3 Archaeofaunal Similarity in the Epipaleolithic........................
4.4 Persistent Regionalization: But with a Twist .......................................
References .....................................................................................................

37
37
38
43
48
49
53
55
55

Archaeofaunal Diversity and Broad Spectrum Diets


in Late Paleolithic Southwest Europe .......................................................
5.1 Introduction ..........................................................................................
5.2 Measuring Changing Archaeofaunal Diversity....................................
5.2.1 Diversity 1: Richness ...............................................................
5.2.2 Diversity 2: Evenness...............................................................
5.2.3 Lagomorphs, Diet Breadth, and Changing Diversity ...............
5.2.4 Diversity, Diet Breadth, and Faunal Turnover .........................
5.3 Nestedness and Faunal Turnover .........................................................
5.3.1 Understanding Zooarchaeological Nestedness ........................
5.3.2 Calculating Nestedness: The NODF ........................................
5.3.3 Regional Nestedness in Southwest Europe ..............................
5.4 Landscape Transformations and Resource Stress ................................
References .....................................................................................................

61
61
63
63
66
68
72
72
72
73
73
74
75

Was There a Broad Spectrum Revolution in Southwest Europe? ..........


6.1 Introduction ..........................................................................................
6.2 What the Archeofaunas Say .................................................................
6.2.1 Mediterranean Iberia ................................................................
6.2.2 Euro-Siberian Iberia .................................................................
6.2.3 Southern France .......................................................................
6.3 Environmental Change and Resource Stress .......................................
6.4 Conclusions ..........................................................................................
References .....................................................................................................

79
79
79
80
80
81
82
83
83

Glossary ............................................................................................................. 85
Index ................................................................................................................... 89

About the Author

Emily Lena Jones is Assistant Professor of Anthropology at the University of New


Mexico. A zooarchaeologist and environmental archaeologist, her research focuses
on humanenvironment interactions, particularly human responses to climate change
and landscape variability. She has worked in the Paleolithic of Southwestern Europe
for over 15 years.

ix

Chapter 1

Paleolithic People, Paleolithic Landscapes

1.1

Introduction

During the academic year, I spend the majority of my time in the urban landscape
of Albuquerque, New Mexico. My day is structured by trips between my home and
the University of New Mexico campus; I usually travel by bicycle, which takes
about 20 min. I pedal past Vietnamese restaurants and taco trucks (sometimes stopping to buy lunch) and through blocks of single-family houses and apartments.
Occasionally, I catch a glimpse of the volcanoes on the horizon to the west, or the
leafy green treetops that mark the passageway of the Rio Grande, or the Sandia
Mountains to the east. Closer to campus, I pass a grocery store (which I may visit
on my way home, to pick up ingredients for dinner) and student rentals (identifiable
by couches on the porches and pizza boxes protruding from the garbage cans). Once
on campus, Ill often stop for coffee before heading to my office.
When Im elsewhere and think of Albuquerque, these are the things I think of:
my neighborhood and the student neighborhood surrounding the University; the
volcanoes, the Rio Grande, and the Sandias; the Vietnamese sandwich shop where
I so often buy lunch and the cheese section at the co-op grocery. There are other
parts of Albuquerque, and I go to them as wellbut it is these places, places I see
or visit everyday, that make up my personal landscape of Albuquerque.
The term landscape has its roots in geography. The famous geographer Carl
Sauer once defined landscape as the expression of interaction between humans and
their environment (1925). Put another way, landscape is the environment (that is,
the world that surrounds us) at the scale at which humans experience it. Landscape
isnt only environment; it is rather those parts of the environment with which we
interact, and how we interact with them.
This book is about Paleolithic people and their landscapesspecifically, about the
huntergatherers who inhabited the Iberian Peninsula and Southern France during the
last part of the Pleistocene, between about 30,000 and 10,000 years ago. The landscapes of Late Paleolithic Southwest Europe would have looked quite different from
The Author(s) 2016
E.L. Jones, In Search of the Broad Spectrum Revolution in Paleolithic Southwest
Europe, SpringerBriefs in Archaeology, DOI 10.1007/978-3-319-22351-3_1

1 Paleolithic People, Paleolithic Landscapes

those of today. Many of the differences relate to climate change: because much of the
worlds water was locked up in glaciers during the Late Paleolithic, coastlines were
lower; in many places, there was more tundra- and steppe-like grassland and less forest; and cold-adapted animals such as reindeer lived as far south as the Iberian
Peninsula. I explore these climate-driven differences at length in Chap. 3. But some
of the differences in landscape are due to cultural changes. Today, Southwest
European landscapes show the legacy of millennia of farming and grazing, but in the
Late Paleolithic, these landscapes were the domain of foragers, people who neither
grew crops nor raised livestock. The vineyards, the fields, the towns, and grazing
lands that define Southwest Europe today were far in the future.

1.2

Why Paleolithic Subsistence Choices?

I opened this chapter with an example meant to demonstrate that landscapes shape
the human experience, even in the twenty-first century when many of us live in
urban environments and spend most of our time inside. But youll note that I came
back to issues of foodmy daily experience of landscape involves restaurants, coffee, and the grocery store. What we eatwhat archaeologists and anthropologists
term subsistenceis a major part of how we interact with our environment. Whether
we live in a rural setting or an urban one, whether we are farmers or gardeners or
huntergatherers, we all need to find some way to eat to live. And our landscape
shapes what food we eat as well as how we interact with our environment to obtain
such food.
A hunter-gatherers experience of landscape is in some ways more direct than
mine. I purchase most of the food I eat. The bulk of my calories come from the
grocery store, and even when I buy local products, the food I obtain goes through
a chain of interactions (grower to wholesaler to store) before it comes home with
me. Modern huntergatherers often do trade with nearby farmers or herders, but
their calories primarily come from food they themselves gather. Before the adoption
of agriculture, in the period archaeologists call the Paleolithic, everyone was a
hunter-gatherer, and there were no farmers or herders with whom to trade. In the
Paleolithic, subsistence would have been a direct tie to the environment.
My research draws on this link, using information about past hunter-gatherer
subsistence to better understand past environments and human interactions with
those environments. The particular case study presented in this volume focuses on
the analysis of animal bone assemblages (or archaeofaunas) from Late Paleolithic
archaeological sites in Southwest Europe. Most (though not all) of these faunas are
the remains of Paleolithic mealsin other words, the garbage pile. Through identification and analysis, we can learn about the Late Paleolithic diet, and how it
changed in response to both shifting environment (something that happened a lot
during the Paleolithic) and cultural changes and innovations. This type of analysis
is called zooarchaeology.

1.3 Why Southwest Europe?

Because of the nature of the hunter-gatherer relationship with landscape, zooarchaeological analysis is a great way to learn about Paleolithic landscapes. In studying animal remains from Paleolithic sites, not only do we learn about the dietary
choices these ancient people made, we also learn about their relationships with the
landscapes they lived inlandscapes very different than those of today, but nevertheless (like today) landscapes that were constantly changing in response to climate
(and possibly human modification as well).
Theres another, less anthropological reason to study archaeological faunas from
the Paleolithic. Archaeofaunas can tell us not only about landscapes, but with careful interpretation they can tell us about ancient environments as well. When humans
hunt, we select certain animals that we see as good to eat from the full suite of
resources available. Research suggests huntergatherers make these choices in
rational, predictable ways (e.g., Smith and Winterhalder 1992; Winterhalder and
Smith 2000). As long as we have a handle on the filter with which humans greet the
environment, we can use these data to understand what prehistoric environments
looked like. While our understanding of climate and environment in the late
Pleistocene has grown in recent years (see for example lvarez-Lao and Garca
2010; Cuenca-Bescos et al. 2009; Gonzalez-Samperiz et al. 2006; Snchez Goi
et al. 2002; Sommer and Nadachowski 2006; Wu et al. 2007), much of this data is
at a global or regional scale. Zooarchaeology provides information about the environment as people experienced itthat is, their landscape.

1.3

Why Southwest Europe?

I hope Ive convinced you that Paleolithic zooarchaeology is a worthwhile subject of


studythat it can help us better understand the nature of humanenvironment interactions throughout history. But this book deals with Paleolithic zooarchaeology in a
very specific place: Southwest Europe, here defined as the western portion of southern France and the Iberian Peninsula (Fig. 1.1). Why so focused, you might ask?
There are several good reasons to concentrate on Southwest Europe. The first,
and most practical, is one of scale. At the end of the last section I argued that zooarchaeological data operates at the scale in which people experience environments;
the flip side of this is that because zooarchaeological assemblages (and depositional
contexts) vary spatially (e.g., Jones 2015), comparisons between different regions
can be challenging.
A second reason for this regional focus has to do with the particular location of
Southwest Europe. Biogeographers identify five macroregions in Southwest Europe
today based on the distributions of different biological taxa (or biological groups,
such as phyla, orders, families, genera, or species; see Fig. 1.1), and two bioclimatic
regions based on differences in temperature and precipitation (Council of Europe
Directorate-General for Environment 2011; Rivas-Martnez et al. 2004). This
regional diversity results from geography and climate history: the Iberian Peninsula

1 Paleolithic People, Paleolithic Landscapes

Fig. 1.1 Southwestern Europe today, showing bioregions as defined by the Council of Europe
Directorate-General for Environment (2011)

is a relatively small, jutting piece of land, separated from the rest of Europe by the
Pyrenees. Today this barrier is a relatively minor one, but in the Paleolithic, when
(like other high-elevation locations) the Pyrenees were glaciated and there was no
motorized transport, the barrier would have been much more significant (Fig. 1.2).
During cold periods, southwestern France, too, was relatively isolated, bounded by
glaciers both to the north and east and by a periglacial desert to the west (Bertran
et al. 2013). So these two areas, bisected by the Pyrenees, were closer to each other
than to other parts of Eurasia. Because of this relative isolation, the Paleolithic
archaeology of Southwest Europe is particular to this region.
And finally, there is a deep history of archaeological research in Southwest
Europe, which is ideal for the type of comparative study Ill be presenting here.
Formal archaeological work in Southwest Europe extends into the nineteenth century; scientists such as Edouard Lartet and Edouard Piette were excavating in southwestern France from 1863 on, and their finds were critical for the establishment of
human antiquity (Grayson 1983; Trigger 2006). Iberia was not far behind, with
excavations in Portugal in the 1860s by J. F. Nery Delgado and work at the Upper
Paleolithic cave site of Altamira by Marcelino Sanz de Sautuola in the 1870s (Straus
1992). While we certainly dont know the entire archaeological record from
Southwest Europe, the long rich history of research means that we know more than
we do for many other regions of the world.

1.4 The Late Paleolithic of Southwest Europe

Fig. 1.2 Southwest Europe with reconstructed coastlines and glaciers for the Last Glacial
Maximum (Council of Europe Directorate-General Enterprise and Industry (DG-ENTR) 2013;
Ray and Adams 2001)

1.4

The Late Paleolithic of Southwest Europe

Before archaeologists had radiocarbon or other means of absolute dating, they


used a combination of changing artifact styles and stratigraphy to establish relative
chronologies. In those areas of the world with a long history of archaeological
inquiry (like Southwest Europe) we now have a fairly robust chronology that combines the detailed knowledge of stratigraphy and changing artifact types with radiocarbon datesyet another reason that this is an excellent area for a regional study.
The end result is a set of culture-historical periods, constructed based on a combination of changing technological style, discrete stratigraphic unit, and absolute dates.
Of course, this upside comes with a corresponding downside. A long research
history often means that there is a proliferation of names for these culture-historical
periods. Table 1.1 shows some of the common names and units of time used for the
Paleolithic of Southwest Europe; I will use the units listed in the This work column throughout this text. Ive tried to keep them as simple and consistent as possible without sacrificing accuracy.

Upper Paleolithic

This work
Epipaleolithic

Gravettian

Solutrean

Magdalenian

Vasco-Cantabria
Azilian
10,5009000 BP
Upper Magdalenian
13,00010,500 BP
Lower Cantabrian
Magdalenian
17,00013,000 BP
Solutrean
21,00017,000 BP
Gravettian
28,00020,000 BP

Aquitaine
Azilian

11,0009000 BP
Upper Magdalenian
13,00011,000 BP
Lower and Middle
Magdalenian
18,00013,000 BP
Solutrean
21,00018,000 BP
Gravettian
29,00021,000 BP

Levantine Spain
Microlaminar
Epipaleolithic
10,5008000 BP
Upper Magdalenian
12,50010,500 BP
Solutreo-Gravettian
and Early Magdalenian
17,50013,000 BP
Solutrean
21,00017,500 BP
Gravettian
29,00021,000 BP

18,00012,500 BP
Solutrean
21,00018,000 BP
Gravettian
29,00021,000 BP

10,50008500 BP
Late Magdalenian
12,50010,500 BP
Late Magdalenian

Portugal
Final Magdalenian/Epipaleolithic

Table 1.1 Late Paleolithic culture-historical periods in Southwest Europe; dates are approximate uncalibrated radiocarbon years before present
(Jones 2004; Straus 1992, 1996, 2005)

6
1 Paleolithic People, Paleolithic Landscapes

References

A final note about dates: those presented in Table 1.1 and elsewhere in this book,
unless otherwise stated, are uncalibrated radiocarbon years. Because the amount of
radiocarbon in the environment has changed through time, radiocarbon dates must
be calibrated to provide a calendar age. There are a number of times in the Late
Paleolithic (notably at the beginning and at the end) for which calibration is difficult, and therefore many scholars present their dates without calibrating them.
Those dates are in radiocarbon years before present; in calendar years, the dates
would be about 2000 years older (though the difference fluctuates through time).
Because I rely on the work of scholars using uncalibrated dates in this book, I follow their lead and likewise use radiocarbon years. So if, for example, I say something occurred at about 11,000 years ago, be aware that in calendar years, that
would be about 12,500 years ago!

1.5

Organization of This Volume

This book describes the results of my research over the last 10 years, and it is organized exactly as my research process was: it begins with theory, then moves through
environmental background before testing theory against available data. Chapter 2
introduces the history of research on broad spectrum diets and explores models for
understanding the increasing dietary diversity, and Chap. 3 describes the environments of Late Paleolithic Southwest Europe and how those environments may have
impacted prehistoric huntergatherers. Chapter 4 tests for evidence of dietary transitions using site location and archaeofaunal similarity data. In Chapter 5, I use archaeofaunal data from different regions in Southwest Europe to test for increasing dietary
diversity from the Upper Paleolithic to the Epipaleolithic. And finally, Chap. 6
assesses whether there was a Broad Spectrum Revolution in Late Paleolithic Southwest
Europe or not, and what role climate-driven resource stress may have played in dietary
change in this region. All original datasets from this project are archived in the
University of New Mexicos data repository, LoboVault (http://repository.unm.edu/),
and are freely available for use.
This volume thus explores the question: what does dietary diversityarchaeological and modernreally mean? While the answers I uncover are specific to the Late
Paleolithic of Southwest Europe, I have found over the last 10 years of research that
theyve provided me with much food for thought about present-day dietary choices
and how they relate to climate and environment. I hope they will do the same for you!

References
lvarez-Lao, D. J., & Garca, N. (2010). Chronological distribution of Pleistocene cold-adapted
large mammal faunas in the Iberian Peninsula. Quaternary International, 212, 120128.
Bertran, P., Sitzia, L., Banks, W. E., Bateman, M. D., Demars, P.-Y., Hernandez, M., et al. (2013).
The Landes de Gascogne (southwest France): Periglacial desert and cultural frontier during the
Palaeolithic. Journal of Archaeological Science, 40, 22742285. doi:http://dx.doi.org/10.1016/j.
jas.2013.01.012

1 Paleolithic People, Paleolithic Landscapes

Council of Europe Directorate-General Enterprise and Industry (DG-ENTR). (2013).


EU-DEM. Retrieved May 3, 2014, from http://www.eea.europa.eu/data-and-maps/data/ds_
resolveuid/HQ5ZD47A6I
Council of Europe Directorate-General for Environment. (2011). Biogeographical regions.
Retrieved May 3, 2014, from http://www.eea.europa.eu/legal/copyright
Cuenca-Bescos, G., Straus, L. G., Morales, M. R. G., & Pimienta, J. C. G. (2009). The reconstruction of past environments through small mammals: From the Mousterian to the Bronze Age in
El Miron Cave (Cantabria, Spain). Journal of Archaeological Science, 36, 947955.
doi:10.1016/j.jas.2008.09.025.
Gonzalez-Samperiz, P., Valero-Garces, B. L., Moreno, A., Jalut, G., Garcia-Ruiz, J. M., MartiBono, C., et al. (2006). Climate variability in the Spanish Pyrenees during the last 30,000 yr
revealed by the El Portalet sequence. Quaternary Research, 66, 3852. doi:10.1016/j.
yqres.2006.02.004.
Grayson, D. K. (1983). The establishment of human antiquity. New York: Academic.
Jones, E. L. (2004). Broad spectrum diets and the European rabbit (Oryctolagus cuniculus):
Dietary change during the Pleistocene-Holocene transition in the Dordogne, Southwestern
France. Ph.D. dissertation, University of Washington, Seattle, WA.
Jones, E. L. (2015). Archaeofaunal evidence of human adaptation to climate change in Upper
Paleolithic Iberia. Journal of Archaeological Science: Reports, 2, 257263. doi:http://dx.doi.
org/10.1016/j.jasrep.2015.02.008
Ray, N., & Adams, J. M. (2001). A GIS-based vegetation map of the world at the Last Glacial
Maximum (25,00015,000 BP). Internet Archaeology, 11. doi:http://dx.doi.org/10.11141/
ia.11.2
Rivas-Martnez, S., Penas, A., & Daz, T. E. (2004). Bioclimatic map of Europe, bioclimates.
Cartographic service. Len, Spain: University of Len.
Snchez Goi, M. F., Cacho, I., Turon, J., Guiot, J., Sierro, F., Peypouquet, J., et al. (2002).
Synchroneity between marine and terrestrial responses to millennial scale climatic variability
during the last glacial period in the Mediterranean region. Climate Dynamics, 19, 95105.
doi:10.1007/s00382-001-0212-x.
Sauer, C. O. (1925). The morphology of landscape (University of California publications in geography, Vol. 2, no. 2). Berkeley, CA: University of California Press.
Smith, E. A., & Winterhalder, B. (Eds.). (1992). Evolutionary ecology and human behavior
(Foundations of human behavior). New York: Aldine de Gruyter.
Sommer, R. S., & Nadachowski, A. (2006). Glacial refugia of mammals in Europe: Evidence from
fossil records. Mammal Review, 36, 251265. doi:10.1111/j.1365-2907.2006.00093.x.
Straus, L. G. (1992). Iberia before the Iberians: The Stone Age prehistory of Cantabrian Spain.
Albuquerque, NM: University of New Mexico Press.
Straus, L. G. (1996). The archaeology of the Pleistocene-Holocene transition in southwest Europe.
In L. G. Straus, B. V. Eriksen, J. M. Erlandson, & D. R. Yesner (Eds.), Humans at the end of
the Ice Age: The Archaeology of the Pleistocene-Holocene transition (Interdisciplinary
Contributions to Archaeology, pp. 83100). New York: Plenum Press.
Straus, L. G. (2005). The Upper Paleolithic of Cantabrian Spain. Evolutionary Anthropology, 14,
145158. doi:10.1002/evan.20067.
Trigger, B. G. (2006). A history of archaeological thought (2nd ed.). Cambridge, MA: Cambridge
University Press.
Winterhalder, B., & Smith, E. A. (2000). Analyzing adaptive strategies: Human behavioral ecology at twenty-five. Evolutionary Anthropology, 9, 5172.
Wu, H. B., Guiot, J. L., Brewer, S., & Guo, Z. T. (2007). Climatic changes in Eurasia and Africa at
the last glacial maximum and mid-Holocene: Reconstruction from pollen data using inverse
vegetation modelling. Climate Dynamics, 29, 211229. doi:10.1007/s00382-007-0231-3.

Chapter 2

Big Game, Small Game: Why It Matters

2.1

Introduction

West of Albuquerque on the north side of Interstate 40 stands a billboard advertising


a restaurant in Tucumcari, New Mexico, 170 miles to the east. Like countless other
billboards along American roadways, this sign lures diners with the promise of
large portions to be had cheaply: a customer who can eat an entire 72 oz. steak, the
sign proclaims, will get the steak for free.
Such advertising is so familiar as to be unremarkable. From chain restaurants
super-large portion sizes (whether offered as a dare, like the 72 oz. steak, or
advertised as best value) to warehouse club stores like Costco and Sams Club,
marketers prey on our sense that we save by buying more. In recent years this type
of advertising has drawn the ire of public health officials and nutrition activists, who
claim that large portion sizes are a contributor to the modern obesity epidemic.
Some have advocated for laws regulating portion sizes, but the food and beverage
industry has argued that such laws would restrict consumer freedom, and (as of this
writing) the American courts have agreed.
The same idea, that bigger equals better, is ingrained in studies of Paleolithic
subsistenceand, as with large portion sizes in the restaurant industry, this assumption has been the subject of considerable debate in recent years. In many parts of the
worldincluding Southwest Europethe Upper Paleolithic is portrayed as a time
of abundance for human huntergatherers, largely based on the fact that zooarchaeological assemblages contain the remains of large herbivores. Textbooks wax
romantic over these environments:
The rich European grasslands and mixed forest habitats supported great numbers of herbivores, including reindeer, deer, bison, wild ox, ibex, woolly rhinoceros, and mammoths.
.This lovely part of the worldoffered excellent places to live. (Wenke 1999: 213)

In southwestern France, reindeer (Rangifer tarandus) often occupy the spotlight;


south of the Pyrenees, the archaeological focus is on red deer (Cervus elaphus), but
the emphasis, when authors stress the richness of these environments, is on the
The Author(s) 2016
E.L. Jones, In Search of the Broad Spectrum Revolution in Paleolithic Southwest
Europe, SpringerBriefs in Archaeology, DOI 10.1007/978-3-319-22351-3_2

10

Big Game, Small Game: Why It Matters

abundance of large mammals. The evidence for this abundance is almost entirely
drawn from faunal assemblages recovered from archaeological contexts.
Similarly, in times and places characterized by archaeofaunas dominated by
small game archaeofaunal assemblages are used as evidence that environments
were resource-poor.
Hunting remained important in the Tardiglacial and Postglacial, but was directed at a wider
faunal spectrum in the context of greatly expanded dietary breadth. Remains of small
mammals, birds, invertebrates, fish, tubers, fruits and other assorted plant remains have
been recovered from sites dating to this time .[T]he expansion of the plant and animal
resource base, obtained from more limited foraging territories, marks a period of more
intensive exploitation of the environment than before. (Villotte et al. 2010)

In other words, a diet containing small mammals, birds, invertebrates and fish (not
to mention plants) is a sign of intensification and stress. Without some constraint
(be it a depleted environment, more limited foraging territories due to human population growth, or some combination of the two), people do not eat smaller game.
In recent years this line of thinking has been the subject of much argument, with
different bodies of theory being brought in to support or counter. But the argument
over bigger equals better has a much longer history than these recent debates.
This chapter covers the historical underpinnings of the modern size debate, and then
explores recent theoretical approaches to this question.

2.2

Historical Examples of Bigger Is Better

The steak example with which this chapter opened is an example of late twentieth/
early twenty-first century consumer culture rather than Paleolithic behavior. But the
concept behind the advertisingthat people think in economic terms when making
choices about dietcan be applied to Paleolithic people as well as modern ones,
and the roots of this concept in the way archaeologists think about diet extend back
into the early years of archaeology as a discipline (Grayson 1984). The assumption
that large fauna are desirable prey for hunters, and small undesirable, can be seen in
a multitude of historical examples.
For instance, in early excavations (and in some cases, as late as the 1960s),
archaeologists routinely only collected larger fauna; smaller fauna was discarded
along with the backdirt (Jones and Gabe 2015; Peres 2010). Only larger fauna, it
was assumed, could reveal anything about human diet. Similarly, the recognition
that humans co-existed with extinct megafauna (a key discovery in the establishment of human antiquity) led some immediately to conjecture over the role human
hunting may have played in these extinctions (Grayson 1984), a debate which has
continued unabated into the twenty-first century; but the smaller taxa that went
extinct at the Pleistocene-Holocene boundary, such as the Aztlan hare (Aztlanolagus)
and the diminutive pronghorn (Capromeryx) have received much less attention.
In some cases where smaller taxa have been considered, animal body size has been
used as a proxy for likeliness that humans might have hunted a particular taxon

2.2

Historical Examples of Bigger Is Better

11

(e.g., Koch and Barnosky 2006). But one of the best examples of bigger is better
comes from Western Europe: the late nineteenth/early twentieth debate over the
presence of a Mesolithic.

2.2.1

A Gap or a Mesolithic?

Today, archaeologists use the term Mesolithic to indicate a cultural period in the
Holocene, just after the period of Paleolithic hunters of the Pleistocene, in which
foragers adapted to changing climate by adjusting their subsistence and mobility
patterns. Mesolithic subsistence is generally portrayed as incorporating smaller
mammals, fish, and plant foods, as opposed to the large game focus of the European
Upper Paleolithic. Today, although definitions of the term vary slightly in different
regions, the term Mesolithic and its sister term Epipaleolithic (see Table 1.1)
are widely used and attract little attention. In the past, however, both the presence of
huntergatherers in the European Holocene prior to the Neolithic spread of agriculture, and (if there were any) their overall significance in European prehistory were
hotly debated.
In 1865, Sir John Lubbock defined the four great epochs of prehistoric archaeology: the Paleolithic (Old Stone Age), when humans co-existed with a suite of
now-extinct species, particularly megafauna; the Neolithic (New Stone Age), characterized by beautiful weapons and instruments made of flint and contemporary
with extant plants and animals; the Bronze Age, when bronze was the primary material used in the manufacture of tools; and the Iron Age, when iron replaced bronze
as the preferred material for tools and weapons (Lubbock 1865: pp. 23). Note that
Mesolithic is not one of these epochs. For Lubbock, the development of tools
reflected the natural evolutionary progression of cultures. Tools in the Paleolithic
were crude flakes, replaced by more sophisticated stone tools in the Neolithic,
replaced by the superior bronze tools, which were in turn superseded by tools of
iron (Lubbock 1865).
Although Mesolithic sites were not well known in the nineteenth century, there
was a series of sitesthe Danish shellmoundswhich were problematic for
Lubbocks scheme, as even he acknowledged. He dealt with these assemblages by
suggesting they might be considered an early Neolithic. In 1872, Westropp coined
the term Mesolithic to account for these and other shellmound sites, but he too
was troubled; from a cultural evolution standpoint, the Mesolithic sites should be
intermediate in progress between Paleolithic sites and Neolithic ones. Instead, the
shellmound sites contained medium-sized and small fauna rather than mammoths
(Mammuthus) and reindeer, and small microlith tools rather than large and impressive spear points. The assemblages of stone trapezes characteristic of the Mesolithic,
when compared with the polished tool types and dramatic cave art of the late
Paleolithic, could hardly be seen as anything but degenerate in a Victorian
evolutionist framework (Childe 1925, 1939, 1947; Obermaier 1924; Price 2000).
The period between the dramatic cave art of the Paleolithic and Neolithic pastoral-

12

Big Game, Small Game: Why It Matters

ists was widely described as one in which people regressed, undergoing a dark
age (see discussion in Clark 1980).
The dark age approach to the Mesolithic is perhaps best illustrated by the
Neolithic revolution of V. Gordon Childe. As others had before him, Childe noted
that all the domestic species used by European Neolithic agriculturalists originated
in southwest Asia. This led him to conclude that the Mesolithic huntergatherers
were mere remnants of a Paleolithic way of life, soon wiped out by Neolithic bearers of civilization, who migrated from the east (Childe 1925, 1939, 1947). Mesolithic
people, Childe argued, were few in number and had no real significance in European
prehistory. In general the mesolithic cultures just described fill gaps in time and
prove the occupation of parts of Europe from the glorious days of mammoth hunting (Childe 1939: 13). In Childes view, Mesolithic populations, with their focus
on small game, were degraded remnants of the Paleolithic hunters.

2.3

The Broad Spectrum Revolution

Childes view did not go uncontested, of course. Most notably, Grahame Clark
argued that the Mesolithic peoples of Europe were complex foragers rather than
degraded vestiges, and that the differences in their technology and subsistence were
evidence of adaptation to climate change (Clark 1932, 1936, 1980). Clarks work at
Starr Carr in the mid-twentieth century inspired an abundance of research on
Mesolithic lifeways in Europe, all with a strong focus on changing environments
and the ways in which Mesolithic foragers adapted to them.
It was in this contexta focus on adaptationthat New Archaeology came to
be interested, more globally, in the change from large-game-dominant diets to smallprey dominant ones. In contrast to Clark, however, the idea that a switch to smaller
prey indicated some sort of crisis was generally accepted by these scholars. Binford
(1968) suggested that in high-latitude Europe there was a shift from large-game
focused diets to broad spectrum ones at the PleistoceneHolocene transition, and
that this was a result of increased population pressure. Flannery (1969) made a similar argument for western Asia, arguing that when foragers broadened their diet to
include small prey in response to shortages of larger prey, they entered into a feedback loop wherein human populations grew and ever more low-return (i.e., small)
items were added to the diet in response, eventually resulting in the adoption of
agriculture (and consequent return to narrower diets) (Fig. 2.1). He named this transition from large prey-focused diets to broad spectrum ones the Broad Spectrum
Revolution.
The Broad Spectrum Revolution is both descriptivea broadening of diet
occurred at the Pleistocene-Holocene transitionand predictive, stating that this
transition (a) should be in response to human impacts rather than climate change;
and (b) should lead eventually to agriculture. It implies several testable questions.
Did a broadening of diet occur as Flannery (and Binford) hypothesized, and if so,
when did this happen? Was broadening a result of human overharvest? Did this
transition truly lead to agriculture? For nearly 50 years archaeologists have been

2.4 The Prey Choice Model and the Body-Size Proxy

13

Fully agricultural economy

Forager population density

Emergence of
agricultural/
domestication
economies

Early Holocene
broad-spectrum
foraging economies

Specialized foraging economies of


the Late Pleistocene

Niche Width
Fig. 2.1 A schematic representation of Flannerys Broad Spectrum Revolution process
(after Clark 1999)

exploring these questions in various forms and in numerous locations around the
globe (Stiner 2001; Zeder 2012). And in recent years, many of themparticularly
zooarchaeologistshave done so using the same tool: the prey choice model of
optimal foraging theory.

2.4

The Prey Choice Model and the Body-Size Proxy

The prey choice model predicts the resources a forager will pursue once that potential resource, or prey type, has been encountered (Grayson and Cannon 1999; Kelly
1995; Stephens and Krebs 1986; Winterhalder and Smith 2000). In this model, foragers chose to pursue prey (or not) based on the resources place in a rank-ordered
set. Ranking is usually based on prey energy return ratesthat is, how much energy
(or other value currency) is returned from prey of a given type per unit pursuit and
handling. According to the prey choice model, a forager interested in maximizing
foraging efficiency will pass up prey only if he or she has a sufficiently high probability of encountering higher-ranked prey. Resources will be added to, or dropped
from, a foragers optimal set according to rank order. As encounter rates with highranked resources decline, a wider and wider array of increasingly lower-ranked prey
types will be taken to make up the difference. This means that increasing numbers
of lower-ranked prey types within a foragers diet will (all else being equal) indicate
a scarcity of higher-ranked resources.

14

Big Game, Small Game: Why It Matters

The primary goal of most archaeological studies using the prey choice model has
been to identify cases of prehistoric exploitation resource depression (Charnov et al.
1976), or reductions in prey capture rates by foragers due to the foragers own subsistence activities. The primary driver for such reductions is an increase in forager
population sizemaking the prey choice model an obvious tool for anyone interested in Broad Spectrum Revolution research.
Archaeologists have had to make some adjustments to use the prey choice model
with archaeological data, however (Grayson and Cannon 1999). One of the most
significant for the purposes of the Broad Spectrum Revolution concerns how to rank
prey (e.g., Bird et al. 2009; Jones 2004; Stiner 2001; Stiner and Munro 2002; Ugan
and Simms 2012). Ethnographers using optimal foraging theory can measure the
exact caloric return of each prey type and rank accordingly; archaeologists cannot,
as the foragers we study died long ago. However, ethnographers who measure return
rates from foragers have observed a correlation: up to a certain point, energetic
return generally tracks prey body size (Broughton 1994; Smith 1991). Archaeologists
using the prey choice model often make use of this correlation and rank prey according to size, with larger prey types assumed to be higher-ranked and smaller ones,
lower-ranked. This is known as the body-size proxy. For researchers applying the
prey choice model in this way, increasing proportions of lower-ranked (i.e., smaller)
prey may be evidence for the Broad Spectrum Revolution (Lupo 2007; Richards
et al. 2001; Rillardon and Brugal 2014; Zhang et al. 2013).

2.5

Alternatives to Traditional Applications


of the Prey Choice Model

If the body-size proxy sounds familiar, it should: this is the bigger is better assumption introduced earlier, though this time with quantitative evidence supporting it.
Despite the quantitative backing, however, the body-size proxy has led to significant
debate among zooarchaeologists using the prey choice model (Bird et al. 2013;
Broughton et al. 2011; Codding et al. 2010a, b; Jones 2004; Ugan 2005; Ugan and
Simms 2012), as well as critiques from archaeologists working outside optimal foraging theory (e.g., Zeder 2012).

2.5.1

Prey Mobility and Energetic Return

One of the first points of argument against the body-size proxy has to do with prey
mobility. Many zooarchaeologists working within the optimal foraging framework
equate prey type with Linnean speciesbut this is not actually how the prey choice
model predicts foragers interact with prey. In optimal foraging terms, a prey type is
not necessarily equivalent to a species; it is defined instead by return per unit handling
time (the amount of time needed to successfully pursue, capture, and process the prey).

2.5 Alternatives to Traditional Applications of the Prey Choice Model

15

Whether the prey moves quickly (like a hare) or slowly (like turtles or shellfish) thus
will impact its ranking (Bird et al. 2009).
Some zooarchaeologists have addressed this problem by using prey mobility to
rank prey, grouping taxa as slow or fast, with the former assumed to be higherranked (e.g., Starkovich 2012, 2014; Stiner 2001; Zhang et al. 2013). Using this
index, Stiner and colleagues have identified declining ratios of fast-to-slow taxa in
the Mediterranean basin far earlier than the Pleistocene-Holocene transition, suggesting some increase in diet breadth in this region as far back as 40,000 years ago
(Stiner and Munro 2002, 2011; Stiner et al. 2000).

2.5.2

Increasing Hunting Efficiency Through Innovation

Prey mobility, however, only gets at one aspect of the amount of time it takes to
pursue, capture, and process game. Huntergatherers both past and present often
use technology and other cultural innovations to decrease handling time, and this
too can impact prey rankings. A wide array of innovations can reduce handling
time. Somesuch as spear-throwers (Marlowe 2005), the bow and arrow (Codding
et al. 2010b), snares (Lombard 2005; Smith 1991), and domestic dogs (Koster
2008)typically assist hunters in the pursuit of individual prey. Otherssuch as
nets and group hunting events (e.g., Jones 2006; Lupo and Schmitt 2002; Marshall
1987; Mrite 1927; Speck and Schaeffer 1950)allow hunters to increase the numbers of individuals taken at one time; these techniques are known collectively as
mass capture or mass harvest strategies. Still other techniques use landscape
modification strategies such as controlled burning (Bliege Bird et al. 2008) to alter
encounter rates for hunters and/or to decrease pursuit time after a resource has been
encountered.
Hunting innovations vary so widely in the taxa they target, the number of hunters
who participate in them, and how they impact energetic return that it is extremely
challenging to generalize about them (e.g., Ugan 2005). A selection of mass harvest
techniques illustrates this difficulty (Table 2.1). However, all do have at least the
potential to alter the relationship between prey body size and return (Jones 2004),
and they can therefore invalidate both the body-size proxy and the inverse relationship between prey mobility and rank (Grayson and Cannon 1999; Kuchikura 1988).
Identifying such innovations in the archaeological record is thus critical for any
zooarchaeological study using the prey choice model.

2.5.3

Gender, Age, and Prey Rank

Forager gender, too, can impact applications of the prey choice model. Ethnographers
have observed a nearly universal sexual division of labor: men generally focus on
higher return, less predictable resources (often large game), while women focus on
the more reliable but lower return plants and smaller animals (Hurtado et al. 1985;
Marlowe 2007; Winterhalder and Smith 1981). Foraging theorists suggest this is

Great Basin

American West

Western Europe

South Pacific
American
Pacific Northwest
Arctic
Arctic
Arctic
Central Congo
Australia
Great Basin

Found in windrow

Rabbit drive

Warren harvest

Eel traps
Salmon weir
Freshwater/anadromous fish
Freshwater/anadromous fish
Freshwater/anadromous fish
Mix of small mammals
Mix of bird species
Mix of fish species

Wild European Rabbit (Oryctolagus


cuniculus)
Eel (Anguilla spp.)
Salmon (Salmonidae)

Target prey type(s)


Grasshopper
(Anabrus simplex)
Grasshopper
(Anabrus simplex)
Jackrabbit (Lepus spp.)

Net
Net
Net
Net
Net
Basket

Variable:
corral, net, none
Variable: fire,
water, net, none
Basket or net
Weir; gill net

None

Technology
Pit

Nelson (1986); Smith (1991)


Nelson (1986); Smith (1991)
Nelson (1986); Smith (1991)
Lupo and Schmitt (2002)
Satterthwait (1987)
Lindstrom (1996)

Marshall (1987)
Hunn (1990)

Schaffer and Gardiner (1995),


Speck and Schaeffer (1950)
Callou (2003), Jones (2006)

Madsen and Kirkman (1988)

Source
Jones and Madsen (1991)

Ice netting
Fall netting
Spring netting
Net hunt
Net hunt
Basket hunt

Location
Great Basin

Method
Driving into pit/stream

Table 2.1 Selected examples of mass collecting techniques

16
Big Game, Small Game: Why It Matters

2.5 Alternatives to Traditional Applications of the Prey Choice Model

17

Fig. 2.2 Fourteenth century illustration of women mass-harvesting the European rabbit
(Oryctolagus cuniculus) assisted by a ferret (from the Queen Mary Psalter, held by the British
Museum)

because the costs and benefits faced by male and female foragers are different, and
so, in prey choice terms, their prey rankings differ (Bliege Bird 1999; Hurtado et al.
1985).
Because zooarchaeologists by definition focus on the animaland often the
large gameportion of the diet, the differing ranked lists of men and women are not
a problem as long as the sexual division of labor was present in the past. However,
as mentioned above, women do sometimes hunt. An increase in relative abundance
of smaller, more reliable resources such as rabbits and shellfish in an archaeofaunal
assemblage could reflect an increase in womens hunting rather than a decrease in
availability of higher-ranked prey (Bailey and Aunger 1989; Bliege Bird and Bird
2008; Codding et al. 2010a). Childrens hunting can also contribute to the zooarchaeological recordand they, too, forage differently from adult men and women
(Bird and Bliege Bird 2000). In addition, women (and children) often participate in
mass harvest activities, as attested not only by ethnographic data but by historic
photographs and illustrations (e.g., Fig. 2.2).
The gender and age problem is important because an increase in the relative
abundance of small prey, which a prey choice-inclined zooarchaeologist might
interpret as evidence for resource depression, could indicate instead an increase in
womens and/or childrens hunting. The literature on the potential impact of gendered foraging choices on zooarchaeological applications of the prey choice theory
is vast (Bliege Bird et al. 2008; Codding et al. 2010a; Hildebrandt and McGuire
2003; Owen 2005; Zeanah 2004), but because archaeologists cannot know who
contributed what to the zooarchaeological record, the only practical solution to this
problem at this point is to be aware of the ways in which gender and age may impact
archaeofaunas.

18

2.5.4

Big Game, Small Game: Why It Matters

Niche Construction Theory

In recent years, much of the criticism of optimal foraging approaches in archaeology has been leveled from researchers using niche construction theory (Laland and
OBrien 2010; Riel-Salvatore 2010; Smith 2007, 2009; Smith and Zeder 2013;
Zeder 2012). Niche construction explicitly rejects the bigger equals better argument and instead sees human environmental engineering as the key to understanding dietary change (Zeder and Smith 2009). While some of the problems with
optimal foraging identified by adherents of niche construction are primarily problems with unwarranted applications of the body-size proxy (as discussed earlier in
this section), considering dietary change from this perspective has allowed researchers to identify instances of stability and sustainability in the archaeological record
(e.g., Smith 2011a, b) rather than long-term resource depression. In terms of Broad
Spectrum Revolution research, niche construction theory suggests the switch to
broad spectrum diets might be best viewed through a lens of opportunity rather than
one of restriction.

2.6

Is Bigger Always Better?

Although many scholars equate bigger is better with optimal foraging theory, this
assumption is neither intrinsic to optimal foraging theory nor restricted to it. The
equation of zooarchaeological assemblages rich in large mammals with good times
and plentiful resources extends back into the nineteenth century; and while many
zooarchaeologists using the prey choice model rely on the body-size proxy, the
body-size proxy is not the same as optimal foraging theory or even the prey choice
model. The bigger equals better assumption has, however, been fundamental to
Broad Spectrum Revolution research since Flannery first coined the name.
This has been changing, however. A lively debate on the appropriateness of the
body-size proxyincluding discussion of alternative methods for evaluating prehistoric prey choicesprung up in the literature in the early years of the twenty-first
century and has continued through the present. The addition of niche construction
theory to these debates furthered theory and broadened the range of approaches
available to archaeologists assessing dietary transitions such as the Broad Spectrum
Revolution, directly tackling the question, is bigger always better? As a result, a
new debate has emerged: do diets broaden out of necessity, opportunity, or some
combination of the two? These new approaches and new questions have led to
increasingly nuanced work on the Broad Spectrum Revolution and an awareness
that contextboth cultural and environmentalis a critical component of any
search for dietary transitions.

References

19

References
Bailey, R. C., & Aunger, R. (1989). Net hunters vs. archers: Variation in womens subsistence
strategies in the Ituri forest. Human Ecology, 17, 273297.
Binford, L. R. (1968). Post-Pleistocene adaptations. In S. R. Binford & L. R. Binford (Eds.), New
perspectives in archaeology (pp. 313341). Chicago: Aldine.
Bird, D. W., & Bliege Bird, R. L. (2000). The ethnoarchaeology of juvenile foraging: Shellfishing
strategies among Meriam children. Journal of Anthropological Archaeology, 19, 461476.
Bird, D. W., Bliege Bird, R., & Codding, B. F. (2009). In pursuit of mobile prey: Martu hunting
strategies and archaeofaunal interpretation. American Antiquity, 74, 329.
Bird, D. W., Codding, B. F., Bliege Bird, R., Zeanah, D. W., & Taylor, C. J. (2013). Megafauna in
a continent of small game: Archaeological implications of Martu Camel hunting in Australias
Western Desert. Quaternary International, 297, 155166. doi:http://dx.doi.org/10.1016/j.
quaint.2013.01.011
Bliege Bird, R. L. (1999). Cooperation and conflict: The behavioral ecology of the sexual division
of labor. Evolutionary Anthropology, 8, 6575.
Bliege Bird, R., Bird, D. W., Codding, B. F., Parker, C. H., & Jones, J. H. (2008). The fire stick
farming hypothesis: Australian Aboriginal foraging strategies, biodiversity, and anthropogenic fire mosaics. Proceedings of the National Academy of Sciences, 105, 1479614801.
doi:10.1073/pnas.0804757105.
Bliege Bird, R., & Bird, D. W. (2008). Why women hunt: Risk and contemporary foraging in a
western desert aboriginal community. Current Anthropology, 49, 655693. doi:10.1086/587700.
Broughton, J. M. (1994). Declines in mammalian foraging efficiency during the late Holocene, San
Francisco Bay, California. Journal of Anthropological Archaeology, 13, 371401.
Broughton, J. M., Cannon, M. D., Bayham, F. E., & Byers, D. A. (2011). Prey body size and ranking in zooarchaeology: Theory, empirical evidence, and applications from the northern Great
Basin. American Antiquity, 76, 403428.
Callou, C. (2003). De la garenne au clapier. Histoire et archologie du lapin europen. Paris:
Publications scientifiques du Musum.
Charnov, E. L., Orians, G. H., & Hyatt, K. (1976). Ecological implications of resource depression.
The American Naturalist, 110, 247259.
Childe, V. G. (1925). The dawn of European civilization. New York: Knopf.
Childe, V. G. (1939). The dawn of European civilization (3rd ed.). London: Keegan Paul.
Childe, V. G. (1947). The dawn of European civilization (4th ed.). London: Keegan Paul.
Clark, J. G. D. (1932). The Mesolithic age in Britain. Cambridge, MA: Cambridge University
Press.
Clark, J. G. D. (1936). The Mesolithic settlement of northern Europe. Cambridge, MA: Cambridge
University Press.
Clark, J. G. D. (1980). Mesolithic prelude: The Palaeolithic-Neolithic transition in old world prehistory. Edinburgh, England: Edinburgh University Press.
Clark, G. A. (1999). Le Msolithique de la cte Atlantique Ibrique: Tendances rcentes. In
A. Thvenin (Ed.), LEurope des derniers chasseurs: pipalolithique et Msolithique
(pp. 5358, 5e Colloque International UISPP, 1823 Septembre 1995). Paris: Editions de
CTHS.
Codding, B. F., Bird, D. W., & Bliege Bird, R. (2010a). Interpreting abundance indices: Some
zooarchaeological implications of Martu foraging. Journal of Archaeological Science, 37,
32003210.
Codding, B. F., Porcasi, J. F., & Jones, T. L. (2010b). Explaining prehistoric variation in the abundance of large prey: A zooarchaeological analysis of deer and rabbit hunting along the Pecho
Coast of Central California. Journal of Anthropological Archaeology, 29, 4761.
Flannery, K. V. (1969). Origins and ecological effects of early domestication in Iran and the Near
East. In P. J. Ucko & G. W. Dimbleby (Eds.), The domestication and exploitation of plants and
animals (pp. 73100). Chicago: Aldine.

20

Big Game, Small Game: Why It Matters

Grayson, D. K. (1984). Nineteenth-century explanations of Pleistocene extinctions: a review and


analysis. In P. S. Martin & R. G. Klein (Eds.), Quaternary extinctions: A prehistoric revolution
(pp. 539). Tucson, AZ: University of Arizona Press.
Grayson, D. K., & Cannon, M. D. (1999). Human paleoecology and foraging theory in the Great
Basin. In C. Beck, D. Rhode, & R. Elston (Eds.), Models for the millennium: Great Basin
anthropology today (pp. 141151). Salt Lake City, UT: University of Utah Press.
Hildebrandt, W. R., & McGuire, K. R. (2003). Large-game hunting, gender-differentiated work
organization, and the role of evolutionary ecology in California and Great Basin prehistory:
A reply to Broughton and Bayham. American Antiquity, 68, 790792.
Hunn, E. S. (1990). Nchi-wna, the big river: Mid-Columbia Indians and their land. Seattle,
WA: University of Washington Press.
Hurtado, A. M., Hawkes, K., Hill, K., & Kaplan, H. (1985). Female subsistence strategies among
Ache hunter-gatherers of eastern Paraguay. Human Ecology, 13, 128.
Jones, E. L. (2004). Dietary evenness, prey choice, and human-environment interactions. Journal
of Archaeological Science, 31, 307317.
Jones, E. L. (2006). Prey choice, mass collecting, and the wild European rabbit (Oryctolagus
cuniculus). Journal of Anthropological Archaeology, 25, 275289.
Jones, E. L., & Gabe, C. (2015). The promise and peril of older collections: meta-analyses in the
American Southwest. Open Quaternary, 1, 1-13, doi:http://doi.org/10.5334/oq.ag.
Jones, K., & Madsen, D. B. (1991). Further experiments in native food procurement. Utah
Archaeology, 4, 6877.
Kelly, R. L. (1995). The foraging spectrum: Diversity in hunter-gatherer lifeways. Washington,
DC: Smithsonian Institution Press.
Koch, P. L., & Barnosky, A. D. (2006). Late quaternary extinctions: State of the debate. Annual
Review of Ecology Evolution and Systematics, 37, 215250. doi:10.1146/annurev.
ecolsys.34.011802.132415.
Koster, J. M. (2008). Hunting with dogs in Nicaragua: An optimal foraging approach. Current
Anthropology, 49, 935944. doi:10.1086/592021.
Kuchikura, Y. (1988). Efficiency and focus of blowpipe hunting among Semaq Beri HunterGatherers of Peninsular Malaysia. Human Ecology, 16, 271305. doi:10.2307/4602888.
Laland, K. N., & OBrien, M. J. (2010). Niche construction theory and archaeology. Journal of
Archaeological Method and Theory, 17, 303322. doi:10.1007/s10816-010-9096-6.
Lindstrom, S. (1996). Great Basin fisherfolk: Optimal diet breadth modelling the Truckee River
aboriginal subsistence fishery. In M. G. Plew (Ed.), Prehistoric hunter-gatherer fishing strategies (pp. 114179). Boise, ID: Boise State University Printing Services.
Lombard, M. (2005). Evidence of hunting and hafting during the Middle Stone Age at Sibidu
Cave, KwaZulu-Natal, South Africa: A multianalytical approach. Journal of Human Evolution,
48, 279300. doi:http://dx.doi.org/10.1016/j.jhevol.2004.11.006
Lubbock, J. (1865). Pre-historic times, as illustrated by ancient remains, and the manners and
customs of modern savages. London: Williams and Northgate.
Lupo, K. D. (2007). Evolutionary foraging models in zooarchaeological analysis: Recent applications and future challenges. Journal of Archaeological Research, 15, 143189. doi:10.1007/
s10814-007-9011-1.
Lupo, K. D., & Schmitt, D. N. (2002). Upper Paleolithic net-hunting, small-prey exploitation, and
womens work effort: A view from the ethnographic and ethnoarchaeological record of the
Congo Basin. Journal of Archaeological Method and Theory, 9, 147179.
Madsen, D. B., & Kirkman, J. E. (1988). Hunting hoppers. American Antiquity, 53, 593604.
Marlowe, F. W. (2005). Hunter-gatherers and human evolution. Evolutionary Anthropology:
Issues, News, and Reviews, 14, 5467. doi:10.1002/evan.20046.
Marlowe, F. W. (2007). Hunting and gathering: The human sexual division of foraging labor.
Cross-Cultural Research, 41, 170195. doi:10.1177/1069397106297529.
Marshall, Y. (1987). Maori mass capture of freshwater eels: An ethnoarchaeological reconstruction of prehistoric subsistence and social behaviour. New Zealand Journal of Archaeology, 9,
5579.

References

21

Mrite, E. (1927). Diffrents engins pour la capture des animaux. Ethnographie, 15(16), 2326.
Nelson, R. K. (1986). Hunters of the northern forest: Designs for survival among the Alaskan
Kutchin (2nd ed.). Chicago: University of Chicago Press.
Obermaier, H. (1924). Fossil man in Spain. New Haven, CT: Yale University Press.
Owen, L. R. (2005). Distorting the past: Gender and the division of labor in the European Upper
Paleolithic (Tbingen publications in prehistory). Tbingen, Germany: Kerns.
Peres, T. M. (2010). Methodological issues in zooarchaeology. In A. M. VanDerwarker & T. M.
Peres (Eds.), Integrating zooarchaeology and Paleoethnobotany: A consideration of issues,
methods, and cases (pp. 1536). New York: Springer.
Price, T. D. (2000). Europes first farmers: An introduction. In T. D. Price (Ed.), Europes first
farmers (pp. 118). Cambridge, MA: Cambridge University Press.
Richards, M. P., Pettitt, P. B., Stiner, M. C., & Trinkaus, E. (2001). Stable isotope evidence for
increasing dietary breadth in the European mid-Upper Paleolithic. Proceedings of the National
Academy of Sciences of the United States of America, 98, 65286532.
Riel-Salvatore, J. (2010). A niche construction perspective on the Middle-Upper Paleolithic transition in Italy. Journal of Archaeological Method and Theory, 17, 323355. doi:10.1007/
s10816-010-9093-9.
Rillardon, M., & Brugal, J.-P. (2014). What about the Broad Spectrum Revolution? Subsistence
strategy of huntergatherers in Southeast France between 20 and 8 ka BP. Quaternary
International, 337, 129153. doi:http://dx.doi.org/10.1016/j.quaint.2014.01.020
Satterthwait, L. (1987). Socioeconomic implications of Australian aboriginal net hunting. Man,
22, 613636.
Schaffer, B. S., & Gardiner, K. M. (1995). The rabbit drive through time: Analysis of the North
American ethnographic and prehistoric evidence. Utah Archaeology, 8, 1325.
Smith, E. A. (1991). Inujjuamiut foraging strategies: Evolutionary ecology of an Arctic hunting
economy (Foundations of human behavior). Hawthorne, NY: Aldine de Gruyter.
Smith, B. D. (2007). Niche construction and the behavioral context of plant and animal domestication. Evolutionary Anthropology, 16, 188199. doi:10.1002/evan.20135.
Smith, B. D. (2009). Resource resilience, human niche construction, and the long-term sustainability of Pre-Columbian Subsistence Economies in the Mississippi River Valley Corridor.
Journal of Ethnobiology, 29, 167183. doi:10.2993/0278-0771-29.2.167.
Smith, B. D. (2011a). The cultural context of plant domestication in eastern north America.
Current Anthropology, 52, S471S484. doi:10.1086/659645.
Smith, B. D. (2011). General patterns of niche construction and the management of wild plant
and animal resources by small-scale pre-industrial societies. Philosophical Transactions of the
Royal Society of London B. Biological Sciences, 366, 836848.
Smith, B. D., & Zeder, M. A. (2013). The onset of the Anthropocene. Anthropocene, 4, 813,
doi:http://dx.doi.org/10.1016/j.ancene.2013.05.001
Speck, F. G., & Schaeffer, C. E. (1950). The deer and the rabbit hunting drive in Virginia and the
Southeast. Southern Indian Studies, 2, 320.
Starkovich, B. M. (2012). Intensification of small game resources at Klissoura Cave 1 (Peloponnese,
Greece) from the Middle Paleolithic to Mesolithic. Quaternary International, 264, 1731.
doi:http://dx.doi.org/10.1016/j.quaint.2011.10.019
Starkovich, B. M. (2014). Optimal foraging, dietary change, and site use during the Paleolithic at
Klissoura Cave 1 (southern Greece). Journal of Archaeological Science, 52, 3955. doi:http://
dx.doi.org/10.1016/j.jas.2014.08.026
Stephens, D. W., & Krebs, J. R. (1986). Foraging theory (Monographs in behavior and ecology).
Princeton, NJ: Princeton University Press.
Stiner, M. C. (2001). Thirty years on the Broad Spectrum Revolution and Paleolithic demography.
Proceedings of the National Academy of Sciences of the United States of America, 98, 69936996.
Stiner, M. C., & Munro, N. D. (2002). Approaches to prehistoric diet breadth, demography, and prey
ranking systems in time and space. Journal of Archaeological Method and Theory, 9, 181214.

22

Big Game, Small Game: Why It Matters

Stiner, M. C., & Munro, N. D. (2011). On the evolution of diet and landscape during the Upper
Paleolithic through Mesolithic at Franchthi Cave (Peloponnese, Greece). Journal of Human
Evolution, 60, 618636. doi:10.1016/j.jhevol.2010.12.005.
Stiner, M. C., Munro, N. D., & Surovell, T. A. (2000). The tortoise and the hare: Small-game use,
the broad-spectrum revolution, and Paleolithic demography. Current Anthropology, 41, 3973.
Ugan, A. (2005). Does size matter? Body size, mass collecting, and their implications for understanding prehistoric foraging behavior. American Antiquity, 70, 7590.
Ugan, A., & Simms, S. (2012). On prey mobility, prey rank, and foraging goals. American
Antiquity, 77, 179185.
Villotte, S., Churchill, S. E., Dutour, O. J., & Henry-Gambier, D. (2010). Subsistence activities and the
sexual division of labor in the European Upper Paleolithic and Mesolithic: Evidence from upper
limb enthesopathies. Journal of Human Evolution, 59, 3543. doi:10.1016/j.jhevol.2010.02.001.
Wenke, R. J. (1999). Patterns in prehistory: Humankinds first three million years. New York:
Oxford University Press.
Westropp, H. M. (1872). Pre-historic phases; or, introductory essays on pre-historic archaeology.
London: Bell & Daldy.
Winterhalder, B., & Smith, E. A. (1981). Hunter-gatherer foraging strategies: Ethnographic and
archeological analyses (Prehistoric archeology and ecology). Chicago: University of Chicago
Press.
Winterhalder, B., & Smith, E. A. (2000). Analyzing adaptive strategies: Human behavioral ecology
at twenty-five. Evolutionary Anthropology, 9, 5172.
Zeanah, D. W. (2004). Sexual division of labor and central place foraging: A model for the Carson
Desert of western Nevada. Journal of Anthropological Archaeology, 23, 132.
Zeder, M. A. (2012). The Broad Spectrum Revolution at 40: Resource diversity, intensification,
and an alternative to optimal foraging explanations. Journal of Anthropological Archaeology,
31, 241264. doi:10.1016/j.jaa.2012.03.003.
Zeder, M. A., & Smith, B. D. (2009). A conversation on agricultural origins: Talking past each
other in a crowded room. Current Anthropology, 50, 681690. doi:10.1086/605553.
Zhang, Y., Zhang, S., Xu, X., Liu, D., Wang, C., Pei, S., et al. (2013). Zooarchaeological perspective on the Broad Spectrum Revolutionin the Pleistocene-Holocene transitional period, with
evidence from Shuidonggou Locality 12, China. Science China Earth Sciences, 56, 14871492.
doi:10.1007/s11430-013-4584-7.

Chapter 3

Climate and Environment in Late Paleolithic


Southwest Europe

3.1

Introduction

Chapter 2 explored the history of archaeological approaches to the Broad Spectrum


Revolution and how human selectionwhat prehistoric people chose to eatmight
be reflected in the zooarchaeological record. But there is an important factor other
than peoples behavior to consider as well. Human selection does impact the taxonomic composition of animals in archaeological sites, but the environments in
which people livedwhat animals were present on the landscape, and in what frequencydoes so as well. When we sit down to eat today, the food we have on the
table is dictated by availability (in the present-day United States, this often means,
does the supermarket carry this item?) as well as by our preferences. In Late
Paleolithic Southwest Europe, this was even more the case, as the specialized storage options and global food distribution networks that allow us to eat food out of
season today did not exist. In addition, as was an explicit component of both
Binfords (1968) and Flannerys (1969) original arguments, the Pleistocene
Holocene transition was a time of profound and dramatic climate and environmental
change. Understanding what Late Paleolithic archaeofaunal assemblages mean
and whether they show evidence for a Broad Spectrum Revolutiontherefore
requires understanding the changing climate and environment during this period.
The plant and animal environments of Southwest Europe shifted wildly and
sometimes rapidly through time in the Late Paleolithic; they also changed spatially
(lvarez-Lao and Garca 2010; Sommer et al. 2011b; Sommer and Zachos 2009).
Certain locations have been suggested as biotic refugia, or areas where resources
concentrated during the coldest parts of the Pleistocene (Bennett and Provan 2008;
Burke et al. 2014; Feliner 2011; Husemann et al. 2014). Other areas seem to have
been deserts both culturally and resourcewise; in these locations, few archaeological sites are known and little food would have been available (e.g., Bertran et al.

The Author(s) 2016


E.L. Jones, In Search of the Broad Spectrum Revolution in Paleolithic Southwest
Europe, SpringerBriefs in Archaeology, DOI 10.1007/978-3-319-22351-3_3

23

24

3 Climate and Environment in Late Paleolithic Southwest Europe

2013). In this chapter, I review the paleoclimate and paleoenvironmental records for
Late Paleolithic Southwest Europe with an eye towards how these environments
might have shaped the landscapes of the huntergatherers who lived in them.

3.2

Climate

Climatologists who work in recent history often rely on historical records of daily
high and low temperatures, precipitation, and river flow. Such records are not available for the deep past, and so for the Late Paleolithic we must rely on proxy measures of climate. Proxies are preserved physical characteristics of the environment
that can stand in for direct observations. Proxy measures commonly used in climate
reconstruction for Late Paleolithic Europe include ice cores, deep-sea cores, and
fossil pollen. These proxies vary in what they are measuring and therefore in what
they tell us, and they respond to environmental change in different ways and at different scales. Some provide global estimates of climate; others regional summaries
of vegetation; and others a more local signal.

3.2.1

Ice Cores and Oxygen Isotopes

On ice sheets in places like Greenland and Antarctica, snow accumulates yearly,
forming distinct layers. Each layer contains information about the climate when the
snow fell. When researchers drill into an ice sheet and remove a core, they thus
obtain a long record of paleoclimatic information. The chemical composition of the
snow can be particularly revealing. For example, the ratio of oxygen isotopes18O
to 16Oin these layers has been shown to correspond with temperature; there is less
18
O in snow in cold periods, and higher concentrations of 16O. This temperaturedriven variation in oxygen isotopes reflects the overall global hydrological cycle.
By calculating oxygen isotope ratios along the length of an ice core, it is possible
to construct curves showing the variability in oxygen isotope ratios over many millennia, and thus have a record of changing global temperatures (for instance, Alley
et al. 1993; Johnsen et al. 2001; Stuiver and Grootes 2000). Greenlands GISP2 ice
core provides one such record (Table 3.1).
Can ice core-derived climate reconstructions tell us about Late Paleolithic climates in Southwest Europe? Yes, but with a caveat. Ice core oxygen isotope records
are global proxies, not local or regional ones. Because the oxygen isotope ratios are
driven by the global hydrological cycle, the climatic information they provide
reflects worldwide averages. If most of the world was cold between 23 and 22 kya,
for instance, but Southwest Europe was comparatively warm, the ice core record
would not show us this.

MIS
1
2

Alboran Sea coreb


Colderwarmer

Southern Francec
Colderwarmer

Euro-Siberian Iberiad
Colderwarmer

Mediterranean Iberiae
Colderwarmer

Derived from 18O values from the GISP2 ice core, constructed from data in Alley (2004)
Derived from 18O-based mean sea-surface temperatures from the Alboran Sea marine core in the Northwestern Mediterranean, derived from Cacho et al. (1999)
c
Derived from fossil pollen-based temperature curve for Southern France from the La Grande Pile pollen core (Davis et al. 2003)the dashed line represents
average summer temperatures, while the unbroken line indicates average winter temperatures
d
Derived from terrestrial pollen-derived temperature curve based on the Laguna de la Roya core (Muoz Sobrino et al. 2013)
e
Derived from marine pollen-based curve showing percent of Mediterranean vegetation, from the SU81-18 core off the coast of Southern Portugal (Snchez
Goi et al. 2008)

This study
Epipaleolithic
Magdalenian
Solutrean
Gravettian

GISP2 Ice corea


Colderwarmer

Table 3.1 Schematic chart of climate reconstructions for the period 1029 kyr BP

3.2
Climate
25

26

3.2.2

3 Climate and Environment in Late Paleolithic Southwest Europe

Marine Cores

A slightly more regional (but still spatially broad) climate proxy comes from deepsea sediment cores. These marine cores are taken by drilling into the sea floor rather
than an ice sheet; they are composted of layers of sediment rather than layers of
annually deposited ice. Like ice cores, marine cores can be used to create an oxygen
isotope record (based on the foraminifera, or shelled fossil protists, in the sediments), although the oxygen isotope signature is the opposite of that in ice cores:
seawater has elevated 18O during cold periods, rather than the depleted 18O found in
ice. These climate reconstructions have been used to establish global marine isotope
climate stages (MIS), which are widely used to identify glacial (even numbers) and
interglacial (odd numbers) in the past (Table 3.1).

3.2.3

Fossil Pollen

Marine cores also provide climate information through the fossil pollen they contain. Pollen grains (produced by all flowering plants) are resistant to decay and (as
allergy sufferers know) disperse widely. Pollen thus accumulates and is preserved in
sediments, including marine sediments. Analysis of the type of pollen in each layer
provides information on what kinds of plants were growing at the time the sediment
was deposited, and this can be used in turn to reconstruct climate.
Marine sediments accumulate pollen from the basin in which the core was taken,
and so reconstructions from these data tend to reflect basinwide conditions. But fossil pollen can also be obtained from cores taken in terrestrial locations, such as lakes
and peatbogs. Terrestrial cores typically trap pollen from a smaller area than do
marine cores, and thus reflect conditions in a more constrained region. Used
together, marine and terrestrial pollen can provide relatively fine-grained vegetation
and climate data for an area.

3.2.4

Climate in Late Paleolithic Southwest Europe

Table 3.1 shows a schematic representation of five different climate reconstructions


for Late Paleolithic Southwest Europe: an oxygen isotope-derived series of temperature reconstructions from Greenlands GISP2 ice core (Alley 2004); an oxygen
isotope-derived temperature reconstruction from a marine core taken in the northwestern Mediterranean (Cacho et al. 1999, 2001); and pollen-based temperature
reconstructions from southern France (Davis et al. 2003), northern Spain (Muoz
Sobrino et al. 2013), and central/southern Iberia (Snchez Goi et al. 2008). Several
major climatic events are apparent in these data. The Last Glacial Maximum (LGM),
when glaciers were at their maximum extent, is reflected by a period of colder than
normal temperatures in the Solutrean. The Oldest Dryas, a cold period in the

3.3

Environment and Landscape

27

Magdalenian, is also discernable, as is the Blling/Allerd interstadial, a warming


period (coinciding with the onset of MIS 1) beginning at 13 kya. The Blling/
Allerd was interrupted at 11 kya by an abrupt cold event known as the Younger
Dryas; and at 10 kya the temperature reconstruction curves show warming, signaling the end of the Younger Dryas and the transition to the Holocene.
On the whole, these proxy records do correspond with one another (Table 3.1;
Snchez Goi et al. 2002) and with other records in Southwest Europe (e.g., Cacho
et al. 1999, 2001; Frigola et al. 2007; Moreno et al. 2005). However, there are also
some regional differences: the Younger Dryas appears to have been briefer and less
intense in Iberia than in southern France, for instance. The fact that there were
regional differences in climate suggests there may have been regional differences in
environment in Late Paleolithic Southwest Europe, too.

3.3

Environment and Landscape

As discussed in Chapter 1, the environments of Southwest Europe today can be


divided into either five biogeographical (Alpine, Atlantic, Continental, and
Mediterranean; see Fig. 1.1) or two bioclimatic (Mediterranean, Temperate/EuroSiberian) regions (Council of Europe Directorate-General for Environment 2011;
Rivas-Martnez et al. 2004; Rivas-Martinez and Rivas-Saenz 2009). The spatial
variations in climate suggest regional divisions in Late Paleolithic Southwest
Europebut were they the same as modern bioregions? Reconstructions of Late
Paleolithic environment and physical landscape (Table 3.2) can help to answer this
question.

3.3.1

The Iberian Peninsula

The Iberian Peninsula, isolated on the extreme southwest edge of Europe, was a
refugium (a location where previously more widespread taxa congregate during
climatically or environmentally challenging times) for the rest of the continent during glacial periods (Gil-Romera et al. 2009; Gmez-Orellana et al. 2013; Hewitt
1999; Naughton et al. 2007; Sommer and Nadachowski 2006; Sommer and Zachos
2009). When Europe got cold, warmer-adapted taxa moved south, and stayed there
until conditions improved. However, although Iberia as a whole seems to have been
home to a variety of refugee taxa, Iberian environments were not uniform during
glacial periods. Different parts of Iberia appear to have contained distinct environments in the Late Paleolithic just as they do now (Jones 2013, 2015; Tarroso 2008).
Present-day Iberia can be divided into three (biogeographic) or two (bioclimatic)
bioregions, all influenced by elevation: the relatively low-elevation central and
southern portion of the peninsula (the Mediterranean region in both schemes)
and the higher-elevation strip along the northern Atlantic coast and Pyrenees
(the Temperate/Euro-Siberian bioclimatic region or the Atlantic and Alpine

2914

Gravettian

LGM

Landes desert is present;


cold-adapted taxa abundant
Landes desert is present;
cold-adapted taxa abundant

Southern France
Vegetation returns to the
Landes desert; extirpation of
reindeer (Rangifer tarandus)
Dune buildup in the Landes
desert; increase in cold-adapted
taxa
Some vegetation returns to the
Landes desert; colonization
by the wild European rabbit;
increase in warm-adapted taxa
(both forest and grassland)
Landes desert is present;
cold-adapted taxa abundant
Peaks in Spruce and Pine
abundance
Increase in steppic vegetation

Expansion of temperate forest;


stable isotope evidence of
some forest constriction
Extirpation of reindeer
(Rangifer tarandus); increase
in forest-adapted taxa; increase
in deciduous trees and
temperate vegetation.
Steppic vegetation

Euro-Siberian Iberia
Expansion of temperate forest

Mix of Mediterranean and


temperate vegetation

Decreasing salinity of
Mediterranean; new marine taxa;
decrease in temperate vegetation
Increase in temperate vegetation

Expansion of Mediterranean
forest continues; limited evidence
of cooling
Initial late glacial expansion of
Mediterranean forest
environments; sharp decline in
temperate vegetation

Mediterranean Iberia
Expansion of Mediterranean
forests

By region; dates are uncalibrated and approximate (data from Aura et al. 2015; Bertran et al. 2013; Cacho et al. 1999, 2001; Callou 2003; Heinz 1991; Heinz
and Barbaza 1998; Jones 2004; Moreno et al. 2005; Snchez Goi et al. 1999, 2000, 2002; Sommer et al. 2008, 2011b, 2013; Sommer and Nadachowski 2006;
Sommer and Zachos 2009; Stevens et al. 2014; Valsecchi et al. 2012)

2418

Solutrean

Oldest Dryas

Blling-Allerd

1311

1813

Younger Dryas

Event
Pleistocene
Holocene transition

1110

Time
(kyr BP)
10

Magdalenian

This study
Epipaleolithic

Table 3.2 Major climatic/environmental events in Late Paleolithic Southwest Europe

28
3 Climate and Environment in Late Paleolithic Southwest Europe

3.3

Environment and Landscape

29

biogeographic regions). Pollen-based vegetation reconstructions for the Iberian


Peninsula suggest the division between the lower-elevation center and south and
higher-elevation north was present in the Late Paleolithic as well (Gmez-Orellana
et al. 2013; Muoz Sobrino et al. 2007; Naughton et al. 2011; Penalba et al. 1997;
Tarroso 2008; Valsecchi et al. 2012). While the vegetation in northern Iberia seems
to have been largely steppic during cold periods, many animal taxa (ranging from
red deer, Cervus elaphus to mountain hare, Lepus timidus) were plentiful in this
region even during the coldest parts of the Pleistocene (Melo-Ferreira et al. 2005;
Sommer et al. 2008). The Younger Dryas seems to have had more of an impact in
the Euro-Siberian portions of Iberia (Stevens et al. 2014) than in the Mediterranean
region (Aura et al. 2015), although in both regions the impacts seem short-lived
compared to those in southern France (as discussed below).
By contrast, temperate vegetation seems to have concentrated in the Mediterranean
Iberian bioregion in glacial periods (Tarroso 2008; though see Muoz Sobrino et al.
2013). In addition the Mediterranean Sea may have provided resources for
Paleolithic people along the coast of Iberia, although the extreme salinity of the
Mediterranean for much of the Pleistocene likely limited available marine taxa
(e.g., Gaspari et al. 2015).
Iberia was thus home to two bioregions in the Late Paleolithic (Fig. 3.1; see
also Jones 2015), roughly corresponding with the bioclimatic regions of today:

Fig. 3.1 Environmental regions, barriers, and corridors in Late Paleolithic Southwest Europe

30

3 Climate and Environment in Late Paleolithic Southwest Europe

a Mediterranean region, comprising the central and southern portions of


the peninsula as well as the Mediterranean coast up to the Pyrenees; and a
Euro-Siberian region in the north. While it is possible there was a third paleobioregion (corresponding with the present-day Alpine biogeographic region), the
glaciers in the Pyrenees make this less likely. The paleobioregions were connected
by a corridor, or an area facilitating interaction: the Ebro basin, a 900-km long
watershed that runs between the two areas (Utrilla et al. 2012).

3.3.2

The Pyrenees

The Pyrenees, the mountain chain separating southern France and the Iberian
Peninsula, are a barrier (Fig. 3.1)in biological terms, an area limiting interaction.
While the Pyrenees are passable (and appear to have been so, at least by humans,
even in the Late Paleolithic; see Utrilla et al. 2012), because of their extent (from the
Bay of Biscay in the west to the Mediterranean in the east, they cover an area greater
than 30,000 km2) and height (on average about 1000 m, but with peaks in the central
ranges over 3000 m) they are formidableand would have been more so in the Late
Paleolithic, when they were thickly glaciated (Palls et al. 2006). Biogeographers
cite this mountain range as the limiting factor in range expansion for a number of
plant and animal taxa (e.g., Sommer et al. 2011a; Vargas et al. 2014).
The Pyrenean barrier may be behind some of the differences in the paleoenvironmental record between France and the Iberian Peninsula (Table 3.2). As mentioned
earlier, in Iberia the Ebro basin facilitates movement between the Euro-Siberian and
Mediterranean bioclimatic regions. This means that during cold periods, warmeradapted taxa in northern Iberia had a corridor to follow to the relatively mild refugia
in the south; when the climate ameliorated, their range could likewise re-expand
relatively quickly. This was not the case in southern France. Any taxon expanding
out of the Iberian Peninsula during a warm period would have to contend with the
Pyrenees to return when the climate turned cold.

3.3.3

Southern France

The environments of southern France are notable for their topographic diversity
(see Figs. 1.2 and 3.1). In the Dordogne valley, broad river valleys cut through limestone cliffs with plateaus on the top. To the west is the Atlantic and to the southwest,
the Mediterranean; and in the east is the high-elevation Massif Central while the
Pyrenees lie to the south. The topographic relief is part of what makes this land so
attractive to both visitors and residents today. In glacial periods, however, the topographic diversity likely meant many areas in this region were only marginally usable
by prehistoric huntergatherers. At the LGM, glaciers would have surrounded
southern France on three sides: the north and east were glaciated as were the

References

31

Pyrenees. In addition, the Landes region of southwestern France was likely a periglacial deserta field of sand dunesfor most of the Late Paleolithic (Bertran et al.
2013). Even the southeast seems to have been notably resource-restricted (Bruxelles
and Jarry 2011). In the coldest periods, huntergatherers may have been restricted
to the river valleyswhich may be one of the reasons that the Dordogne is notably
rich in Paleolithic archaeological sites.
The paleoenvironmental record suggests that biotas as well as humans were
restricted in their distribution (Table 3.2). Some areas (again, particularly in the
Dordogne)seem to have been rich in both plants and animals (Delpech 1999;
Sommer and Nadachowski 2006). Others seem to have been resource-poor
(Bruxelles and Jarry 2011; Jones 2007). And there is a temporal aspect to this as
well: cold-adapted taxa coexist with warm-adapted taxa in some deposits, but other
warm-adapted taxanotably the wild European rabbit, Oryctolagus cuniculus
seem to have only managed a truly successful colonization of southern France with
the Blling-Allerd warming at ca. 13 kya (Callou 2003; Jones 2006). The variability in the record makes any definitive understanding of the environments of this
region challenging. Even the Landes desert seems occasionally to have been home
to microrefugia (de Lafontaine et al. 2014).
Late Paleolithic environments in southern France, in short, were not like those of
today. This region seems to have been what paleoecologists call a non-analogous
environmentan environment for which there is no present-day equivalent
(Huntley 1990; Prentice et al. 1996).

3.4

Regionalization and Paleolithic People

European environments today are marked by a strong latitudinal gradient, with


marked species turnover from north to south (Dobrovolski et al. 2012; Moreno et al.
2014; Svenning et al. 2011). This seems to have been the case in the Late Paleolithic
too, as paleoenvironmental data suggest three distinct paleobioclimatic regions:
Mediterranean Iberia, Euro-Siberian Iberia, and southern France (Fig. 3.1).
Archaeological work suggests that, like flora and fauna, humans often also respond
to climate change regionally (Jones 2015). In Paleolithic Iberia in particular, regional
differences in human mobility, land use, and hunting behavior have been recorded
(Barton et al. 2013; Haws 2012; Jones 2013; Marn Arroyo 2010). Do the paleoenvironmental regions in Late Paleolithic Southwest Europe correspond with differences in human subsistence strategy? I address this question in the next chapter.

References
Alley, R. B. (2004). GISP2 ice core temperature and accumulation data. In I. P. W. D. C. f.
Paleoclimatology (Ed.). Boulder, CO: NOAA/NGDC Paleoclimatology Program.

32

3 Climate and Environment in Late Paleolithic Southwest Europe

Alley, R. B., Meese, D. A., Shuman, A. J., Gow, A. J., Taylor, K. C., Grootes, P. M., et al. (1993).
Abrupt accumulation increase at the Younger Dryas termination in the GISP2 ice core. Nature,
362, 527529.
lvarez-Lao, D. J., & Garca, N. (2010). Chronological distribution of Pleistocene cold-adapted
large mammal faunas in the Iberian Peninsula. Quaternary International, 212, 120128.
Aura, J. E., Jord, J. F., Montes, L., & Utrilla, P. (2015). Human responses to Younger Dryas in the
Ebro Valley and Mediterranean Watershed (Eastern Spain). Quaternary International, 242,
348359.
Barton, C. M., Villaverde, V., Zilho, J., Aura, J. E., Garcia, O., & Badal, E. (2013). In glacial environments beyond glacial terrains: Human eco-dynamics in late Pleistocene Mediterranean
Iberia. Quaternary International, 318, 5368. doi:http://dx.doi.org/10.1016/j.quaint.2013.05.007.
Bennett, K. D., & Provan, J. (2008). What do we mean by refugia? Quaternary Science Reviews,
27, 24492455. doi:10.1016/j.quascirev.2008.08.019.
Bertran, P., Sitzia, L., Banks, W. E., Bateman, M. D., Demars, P.-Y., Hernandez, M., et al. (2013).
The Landes de Gascogne (southwest France): Periglacial desert and cultural frontier during the
Palaeolithic. Journal of Archaeological Science, 40, 22742285. doi:http://dx.doi.org/10.1016/j.
jas.2013.01.012.
Binford, L. R. (1968). Post-Pleistocene adaptations. In S. R. Binford & L. R. Binford (Eds.), New
perspectives in archaeology (pp. 313341). Chicago: Aldine.
Bruxelles, L., & Jarry, M. (2011). Climatic conditions, settlement patterns and cultures in the
Paleolithic: The example of the Garonne Valley (southwest France). YJHEV Journal of Human
Evolution, 61, 538548.
Burke, A., Levavasseur, G., James, P. M. A., Guiducci, D., Izquierdo, M. A., Bourgeon, L., et al.
(2014). Exploring the impact of climate variability during the Last Glacial Maximum on the
pattern of human occupation of Iberia. Journal of Human Evolution, 73, 3546. doi:http://
dx.doi.org/10.1016/j.jhevol.2014.06.003.
Cacho, I., Grimalt, J. O., Canals, M., Sbaffi, L., Shackleton, N. J., Schnfeld, J., et al. (2001).
Variability of the western Mediterranean Sea surface temperature during the last 25,000 years
and its connection with the Northern Hemisphere climatic changes. Paleoceanography, 16,
4052. doi:10.1029/2000PA000502.
Cacho, I., Grimalt, J. O., Pelejero, C., Canals, M., Sierro, F. J., Flores, J. A., et al. (1999).
Dansgaard-Oeschger and Heinrich event imprints in Alboran Sea paleotemperatures.
Paleoceanography, 14, 698705. doi:10.1029/1999PA900044.
Callou, C. (2003). De la garenne au clapier. Histoire et archologie du lapin europen. Paris:
Publications Scientifiques du Musum.
Council of Europe Directorate-General for Environment. (2011). Biogeographical regions.
Retrieved May 3, 2014, from http://www.eea.europa.eu/legal/copyright.
Davis, B. A. S., Brewer, S., Stevenson, A. C., & Guiot, J. (2003). The temperature of Europe during
the Holocene reconstructed from pollen data. Quaternary Science Reviews, 22, 17011716.
de Lafontaine, G., Amasifuen Guerra, C. A., Ducousso, A., & Petit, R. J. (2014). Cryptic no more:
Soil macrofossils uncover Pleistocene forest microrefugia within a periglacial desert. New
Phytologist, 204, 715729. doi:10.1111/nph.12833.
Delpech, F. (1999). Biomasse donguls au Palolithique et infrences sur la dmographie. Palo,
11, 1942.
Dobrovolski, R., Melo, A. S., Cassemiro, F. A. S., & Diniz-Filho, J. A. F. (2012). Climatic history and
dispersal ability explain the relative importance of turnover and nestedness components of beta
diversity. Global Ecology and Biogeography, 21, 191197. doi:10.1111/j.1466-8238.2011.00671.x.
Feliner, G. N. (2011). Southern European glacial refugia: A tale of tales. Taxon, 60, 365372.
Flannery, K. V. (1969). Origins and ecological effects of early domestication in Iran and the Near
East. In P. J. Ucko & G. W. Dimbleby (Eds.), The domestication and exploitation of plants and
animals (pp. 73100). Chicago: Aldine.
Frigola, J., Moreno, A., Cacho, I., Canals, M., Sierro, F. J., Flores, J. A., et al. (2007). Holocene
climate variability in the western Mediterranean region from a deepwater sediment record.
Paleoceanography, 22, PA2209. doi:10.1029/2006PA001307.

References

33

Gaspari, S., Scheinin, A., Holcer, D., Fortuna, C., Natali, C., Genov, T., et al. (2015). Drivers of
population structure of the Bottlenose Dolphin (Tursiops truncatus) in the Eastern Mediterranean
Sea. Evolutionary Biology, 42, 177190. doi:10.1007/s11692-015-9309-8.
Gil-Romera, G., Carrin, J. S., McClure, S. B., Schmich, S., & Finlayson, C. (2009). Holocene
vegetation dynamics in Mediterranean Iberia: Historical contingency and climatehuman
interactions. Journal of Anthropological Research, 65, 271285.
Gmez-Orellana, L., Ramil-Rego, P., & Muoz Sobrino, C. (2013). The response of vegetation at
the end of the last glacial period (MIS 3 and MIS 2) in littoral areas of NW Iberia. Boreas, 42,
729744. doi:10.1111/j.1502-3885.2012.00310.x.
Haws, J. A. (2012). Paleolithic socionatural relationships during MIS 3 and 2 in central Portugal.
Quaternary International, 264, 6177. doi:http://dx.doi.org/10.1016/j.quaint.2011.10.003.
Heinz, C. (1991). Upper Pleistocene and Holocene vegetation in the south of France and Andorra,
adaptations and first ruptures: New charcoal analysis data. Review of Palaeobotany and
Palynology, 69, 288324.
Heinz, C., & Barbaza, M. (1998). Environmental changes during the Late Glacial and Post-Glacial
in the central Pyrenees (France): New charcoal analysis and archaeological data. Review of
Palaeobotany and Palynology, 104, 117.
Hewitt, G. M. (1999). Post-glacial re-colonization of European biota. Biological Journal of the
Linnean Society, 68, 87112. doi:10.1111/j.1095-8312.1999.tb01160.x.
Huntley, B. (1990). European pollen history: Palaeovegetation maps from pollen data13 000 yr
BP to present. Journal of Quaternary Science, 5, 103122.
Husemann, M., Schmitt, T., Zachos, F. E., Ulrich, W., & Habel, J. C. (2014). Palaearctic biogeography revisited: Evidence for the existence of a North African refugium for Western Palaearctic
biota. Journal of Biogeography, 41, 8194. doi:10.1111/jbi.12180.
Johnsen, S. J., Dahl-Jensen, D., Gundestrup, N., Steffensen, J. P., Clausen, H. B., Miller, H., et al.
(2001). Oxygen isotope and palaeotemperature records from six Greenland ice-core stations:
Camp Century, Dye-3, GRIP, GISP2, Renland and NorthGRIP. Journal of Quaternary Science,
16, 299307. doi:10.1002/jqs.622.
Jones, E. L. (2004). Broad spectrum diets and the European rabbit (Oryctolagus cuniculus):
Dietary change during the Pleistocene-Holocene transition in the Dordogne, Southwestern
France. Seattle, WA: Ph.D. dissertation. University of Washington.
Jones, E. L. (2006). Prey choice, mass collecting, and the wild European rabbit (Oryctolagus
cuniculus). Journal of Anthropological Archaeology, 25, 275289.
Jones, E. L. (2007). Subsistence change, landscape use, and changing site elevation at the
Pleistocene-Holocene transition in the Dordogne of Southwestern France. Journal of
Archaeological Science, 34, 344353.
Jones, E. L. (2013). Mobility, settlement, and resource patchiness in Upper Paleolithic Iberia.
Quaternary International, 318, 4652. doi:http://dx.doi.org/10.1016/j.quaint.2013.05.027.
Jones, E. L. (2015). Archaeofaunal evidence of human adaptation to climate change in Upper
Paleolithic Iberia. Journal of Archaeological Science: Reports, 2, 257263. doi:http://dx.doi.
org/10.1016/j.jasrep.2015.02.008.
Marn Arroyo, A. B. (2010). Arqueozoologa en el Cantbrica oriental durante la transicin
Plestoceno/Holoceno. Cantabria, Spain: Santander/PUbliCan, Ediciones Universidad de
Cantabria.
Melo-Ferreira, J., Boursot, P., Suchentrunk, F., Ferrand, N., & Alves, P. C. (2005). Invasion from
the cold past: Extensive introgression of mountain hare (Lepus timidus) mitochondrial DNA
into three other hare species in northern Iberia. Molecular Ecology, 14, 24592464.
doi:10.1111/j.1365-294X.2005.02599.x.
Moreno, A., Cacho, I., Canals, M., Grimalt, J. O., Snchez-Goi, M. F., Shackleton, N., et al.
(2005). Links between marine and atmospheric processes oscillating on a millennial time-scale.
A multi-proxy study of the last 50,000 yr from the Alboran Sea (Western Mediterranean Sea).
Quaternary Science Reviews, 24, 16231636. doi:http://dx.doi.org/10.1016/j.quascirev.
2004.06.018.

34

3 Climate and Environment in Late Paleolithic Southwest Europe

Moreno, A., Svensson, A., Brooks, S. J., Connor, S., Engels, S., Fletcher, W., et al. (2014). A
compilation of Western European terrestrial records 608 ka BP: Towards an understanding of
latitudinal climatic gradients. Quaternary Science Reviews, 106, 167185. doi:http://dx.doi.
org/10.1016/j.quascirev.2014.06.030.
Muoz Sobrino, C., Heiri, O., Hazekamp, M., van der Velden, D., Kirilova, E. P., Garca-Moreiras,
I., et al. (2013). New data on the Lateglacial period of SW Europe: A high resolution multiproxy
record from Laguna de la Roya (NW Iberia). Quaternary Science Reviews, 80, 5877. doi:http://
dx.doi.org/10.1016/j.quascirev.2013.08.016.
Muoz Sobrino, C., Ramil-Rego, P., & Gmez-Orellana, L. (2007). Late Wurm and early Holocene
in the mountains of northwest Iberia: Biostratigraphy, chronology and tree colonization.
Vegetation History and Archaeobotany, 16, 223240. doi:10.1007/s00334-006-0083-5.
Naughton, F., Drago, T., Snchez Goi, M. F., & Freitas, M. C. (2011). Climate variability in the
North-Western Iberian Peninsula during the last deglaciation. In P. Duarte & J. M. SantanaCasiano (Eds.), Oceans and the atmospheric carbon content (pp. 122). Berlin, Germany:
Springer.
Naughton, F., Snchez Goi, M. F., Desprat, S., Turon, J. L., Duprat, J., Malaize, B., et al. (2007).
Present-day and past (last 25 000 years) marine pollen signal off western Iberia. Marine
Micropaleontology, 62, 91114. doi:10.1016/j.marmicro.2006.07.006.
Palls, R., Rods, ., Braucher, R., Carcaillet, J., Ortuo, M., Bordonau, J., et al. (2006). Late
Pleistocene and Holocene glaciation in the Pyrenees: A critical review and new evidence from
10Be exposure ages, south-central Pyrenees. Quaternary Science Reviews, 25, 29372963.
doi:http://dx.doi.org/10.1016/j.quascirev.2006.04.004.
Penalba, M. C., Arnold, M., Guiot, J., Duplessy, J. C., & Beaulieu, J. L. D. (1997). Termination of
the last glaciation in the Iberian Peninsula inferred from the pollen sequence of Quintanar de la
Sierra. Quaternary Research, 47, 205214.
Prentice, I. C., Guiot, J., Huntley, B., Jolly, D., & Cheddadi, R. (1996). Reconstructing biomes
from palaeoecological data: A general method and its application to European pollen data at 0
and 6 ka. Climate Dynamics, 12, 185194.
Rivas-Martnez, S., Penas, A., & Daz, T. E. (2004). Bioclimatic map of Europe, bioclimates.
Cartographic service. Len, Spain: University of Len, Spain.
Rivas-Martinez, S., & Rivas-Saenz, S. (2009). Worldwide Bioclimatic Classification System,
19962009. Retrieved from http://www.globalbioclimatics.org2015.
Snchez Goi, M. F., Cacho, I., Turon, J., Guiot, J., Sierro, F., Peypouquet, J., et al. (2002).
Synchroneity between marine and terrestrial responses to millennial scale climatic variability
during the last glacial period in the Mediterranean region. Climate Dynamics, 19, 95105.
doi:10.1007/s00382-001-0212-x.
Snchez Goi, M. F., Eynaud, F., Turon, J. L., & Shackleton, N. J. (1999). High resolution palynological record off the Iberian margin: Direct land-sea correlation for the last Interglacial complex. Earth and Planetary Science Letters, 171, 123137. doi:http://dx.doi.org/10.1016/
S0012-821X(99)00141-7.
Snchez Goi, M. F., Landais, A., Fletcher, W. J., Naughton, F., Desprat, S., & Duprat, J. (2008).
Contrasting impacts of DansgaardOeschger events over a western European latitudinal transect modulated by orbital parameters. Quaternary Science Reviews, 27, 11361151. doi:http://
dx.doi.org/10.1016/j.quascirev.2008.03.003.
Snchez Goi, M. F., Turon, J.-L., Eynaud, F., & Gendreau, S. (2000). European Climatic Response
to Millennial-Scale Changes in the AtmosphereOcean System during the Last Glacial Period.
Quaternary Research, 54, 394403. doi:http://dx.doi.org/10.1006/qres.2000.2176.
Sommer, R. S., Benecke, N., Lougas, L., Nelle, O., & Schmolcke, U. (2011a). Holocene survival
of the wild horse in Europe: A matter of open landscape? Journal of Quaternary Science, 26,
805812. doi:10.1002/jqs.1509.
Sommer, R. S., Fritz, U. W. E., Sepp, H., Ekstrm, J., Persson, A., & Liljegren, R. (2011b). When
the pond turtle followed the reindeer: Effect of the last extreme global warming event on the

References

35

timing of faunal change in Northern Europe. Global Change Biology, 17, 20492053.
doi:10.1111/j.1365-2486.2011.02388.x.
Sommer, R. S., Kalbe, J., Ekstrm, J., Benecke, N., & Liljegren, R. (2013). Range dynamics of the
reindeer in Europe during the last 25,000 years. Journal of Biogeography, 41, 298306.
doi:10.1111/jbi.12193.
Sommer, R. S., & Nadachowski, A. (2006). Glacial refugia of mammals in Europe: Evidence from
fossil records. Mammal Review, 36, 251265. doi:10.1111/j.1365-2907.2006.00093.x.
Sommer, R. S., & Zachos, F. E. (2009). Fossil evidence and phylogeography of temperate species:
glacial refugia and post-glacial recolonization. Journal of Biogeography, 36, 20132020.
Sommer, R. S., Zachos, F. E., Street, M., Jris, O., Skog, A., & Benecke, N. (2008). Late Quaternary
distribution dynamics and phylogeography of the red deer (Cervus elaphus) in Europe.
Quaternary Science Reviews, 27, 714733.
Stevens, R. E., Hermoso-Buxn, X. L., Marn-Arroyo, A. B., Gonzlez-Morales, M. R., & Straus,
L. G. (2014). Investigation of Late Pleistocene and Early Holocene palaeoenvironmental
change at El Mirn cave (Cantabria, Spain): Insights from carbon and nitrogen isotope analyses
of red deer. Palaeogeography, Palaeoclimatology, Palaeoecology, 414, 4660. doi:http://dx.
doi.org/10.1016/j.palaeo.2014.05.049.
Stuiver, M., & Grootes, P. M. (2000). GISP2 oxygen isotope ratios. Quaternary Research, 53,
277284.
Svenning, J.-C., Fljgaard, C., & Baselga, A. (2011). Climate, history and neutrality as drivers of
mammal beta diversity in Europe: Insights from multiscale deconstruction. Journal of Animal
Ecology, 80, 393402. doi:10.1111/j.1365-2656.2010.01771.x.
Tarroso, P. (2008). Late Quaternary landscape dynamics in the Iberian Peninsula and Balearic
Islands. Oporto, Portugal: University of Oporto.
Utrilla, P., Domingo, R., Montes, L., Mazo, C., Rodans, J. M., Blasco, F., et al. (2012). The Ebro
Basin in NE Spain: A crossroads during the Magdalenian. Quaternary International, 272273,
88104. doi:http://dx.doi.org/10.1016/j.quaint.2012.04.024.
Valsecchi, V., Goi, M. F. S., & Londeix, L. (2012). Vegetation dynamics in the Northeastern
Mediterranean region during the past 23 000 yr: Insights from a new pollen record from the Sea
of Marmara. Climate of the Past, 8, 19411956. doi:10.5194/cp-8-1941-2012.
Vargas, P., Thbaud, C., Burrus, M., Suchet, C., & Liberal, I. (2014). The evolutionary history of
Antirrhinum in the Pyrenees inferred from phylogeographic analyses. BMC Evolutionary
Biology, 14, 114.

Chapter 4

Human Subsistence and the Archaeofaunal


Record of Late Paleolithic Southwest Europe

4.1

Introduction

In Chap. 3, I showed that environments in Late Paleolithic Southwest Europe were


strongly regional, as they are todaybut also that these environments changed
through time. On the Iberian Peninsula (isolated from southern France by the
Pyrenees and the Landes peri-glacial desert) paleoenvironmental regions seem to
have corresponded to modern-day bioclimatic regions (e.g., Rivas-Martnez et al.
2004; Rivas-Martinez and Rivas-Saenz 2009) all the way back to the LGM. The
central and southern portions of the peninsulathose areas in the Mediterranean
paleobioclimatic region (Fig. 3.1)contained concentrations of warmer-adapted
plant and animal taxa during the coldest periods. Some temperate adapted taxa (such
as red deer, Cervus elaphus) ranged into the Euro-Siberian portion of Iberia as well.
In southern France, by contrast, there seems to have been a shifting and patchy environmental mosaic throughout the Upper Paleolithicwhat paleoecologists call a
non-analogous environment or an environment not comparable to any extant today.
In the Epipaleolithic, the warming associated with the Blling/Allerd interstadial
led to an expansion of taxa out of refugia throughout Iberia and associated environmental changes in all regions, but the Younger Dryas cold event had different impacts
in southern France than it did on the Iberian Peninsula (Table 3.2).
How did Late Paleolithic huntergatherers in each of these regions respond to the
environmental changes? What can their subsistence strategies tell us about environmental variability through space and time? And finally, do site location and/or archaeofaunal similarity values suggest a Broad Spectrum Revolution in the Epipaleolithic of
any of these regions? I explore these questions here. I use site location and archaeofaunal data to test for regionalization. Particularly, I ask the question, are the paleobioregions proposed in Chap. 3 (Fig. 3.1) supported by the archaeological data?
Because I am exploring data by paleobioregion, I use the terms Mediterranean
Iberia, Euro-Siberian Iberia, and southern France in the discussion that follows.
It is important to remember that a site I locate in Mediterranean Iberia is not nec The Author(s) 2016
E.L. Jones, In Search of the Broad Spectrum Revolution in Paleolithic Southwest
Europe, SpringerBriefs in Archaeology, DOI 10.1007/978-3-319-22351-3_4

37

38

4 Human Subsistence and the Archaeofaunal Record of Late Paleolithic Southwest

essarily near the Mediterranean coastit may be in the Spanish meseta (the highelevation region of central Spain) or in interior Portugal! But any such site
nevertheless is located in a region today classed as having a Mediterranean climate,
which is why I term it Mediterranean here.

4.2

Site Location and Human Subsistence

Archaeologists working in Late Paleolithic Europe commonly use changes in site


location to ask questions about human adaptation to changing environments (e.g.,
Banks et al. 2008, 2009; Miller 2012; Straus et al. 2000; Wygal and Heidenreich
2014). In Southwest Europe, it is clear site distributions did change somewhat
through time. Figure 4.1 shows site distributions in the Gravettian (a) and Solutrean
(b) (data drawn from Vermeersch 2014). Note that in the Gravettian the bulk of the
sites are in the north. In the Solutrean, during the LGM, sites congregate in the
refugia of Mediterranean Iberia and the river valleys of southern France. Site distribution continued to change in later periods: in the Magdalenian (Fig. 4.2a) sites
expanded to high-elevation areas, a process that continued in the Epipaleolithic
(Fig. 4.2b). The differences are obvious, but what do they mean?
One way to track subsistence change in the archaeological record is through
research into prehistoric mobility (Conard et al. 2012; Henry 1994; Holt 2003;
Jones 2013; Straus 1986; Villaverde et al. 1996). Groups with different subsistence
strategies move across the landscape differently, and so looking at differences in
archaeological site location can provide insight into how past peoples were obtaining their food (Fisher 2002; Grove 2010). To explore mobility, archaeologists often
make use of the logistical/residential continuum (Binford 1980; Kelly 1995). In this
scheme, huntergatherers following a logistical mobility pattern have a base camp
from which small groups or individuals may make short forays to obtain specialized
resources; the main base camp moves infrequently. People following a more residential strategy move more frequently, shifting their base camp from location to
location to exploit predictable resource patches. Residential mobility is more common in highly seasonal environments, as huntergatherers move to exploit resources
that come into season in a particular location, year after year. Frequently, they locate
sites in along an elevation gradient, maximizing access to different resource patches.
Logistical mobility is more likely in areas with less predictable patches; in such a
situation there is little to be gained by having sites in such specific locations. Instead,
sites will be located within range of the widest number of possible patches, often at
mid-range elevations (Aldenderfer 2008; Jones 2007, 2013; Morgan 2008).
Analyses of site elevation variance (Jones 2007, 2013) allow us to test these
hypotheses with the archaeological record. Statistical variance describes the
degree to which a group of numbers is spread out (if all the numbers in the group
are the same, the variance will be zero). Site elevation variance, therefore, shows
whether sites are located at many different elevations, or concentrated in a smaller
range.If site elevation variance is high, sites are dispersed across a variety of elevations; if variance is low, they are concentrated at a particular range of elevations.

4.2

Site Location and Human Subsistence

39

Fig. 4.1 Site distribution in the Gravettian (a) and Solutrean (b). Circles indicate sites in the
Mediterranean Iberia paleobioclimatic region; triangles, sites in Euro-Siberian Iberia; and squares,
sites in southern France

40

4 Human Subsistence and the Archaeofaunal Record of Late Paleolithic Southwest

Fig. 4.2 Site distribution in the Magdalenian (a) and Epipaleolithic (b). Circles indicate sites in
Mediterranean Iberia; triangles, sites in Euro-Siberian Iberia; and squares, sites in southern France

4.2

Site Location and Human Subsistence

41

Predictions for variance in site elevation in Late Paleolithic Southwest Europe are
therefore as follows:
Residential mobility > Predictable resources > Frequent movement > High site variance
Logistical mobility > Less predictable resources > Infrequent movement > Low site
variance
If a region contains high site elevation variance, this suggests a predictably
patchy environment and thus a more residential strategy. If a region shows low site
elevation variance, this suggests resources are scattered across the region less predictably, and thus huntergatherers may have foraged more logistically.
To test for differences in site elevation variance, I used site distribution data
presented in Figs. 4.1 and 4.2 in conjunction with a 25m digital elevation model
(DEM) of Western Europe (Council of Europe Directorate-General Enterprise and
Industry (DG-ENTR) 2013). I used the DEM to extract elevations for each site and
analyzed these data by paleobioregion using the statistical package PAST (Hammer
et al. 2001).
The analysis shows that the three paleobioregions of Late Paleolithic Southwest
Europe proposed in Chap. 3southern France, Euro-Siberian Iberia, and
Mediterranean Iberiahave distinct differences in site elevation variance (Table 4.1,
Fig. 4.3). Sites in Euro-Siberian Iberia have high-elevation variance throughout the
Late Paleolithic. An Analysis of Variance (ANOVA) shows that mean site elevation
in this region stays consistent through time, and a Bartletts test for homogeneity of
variance indicates that site elevation variance likewise does not change (Table 4.2).
This finding concurs with other research (e.g., Marn Arroyo 2008, 2009; Menndez

Table 4.1 Descriptive statistics for changing elevation by chronological period


Region
All regions

Mediterranean
Iberia

Euro-Siberian
Iberia

Southern France

Period
Gravettian
Solutrean
Magdalenian
Epipaleolithic
Gravettian
Solutrean
Magdalenian
Epipaleolithic
Gravettian
Solutrean
Magdalenian
Epipaleolithic
Gravettian
Solutrean
Magdalenian
Epipaleolithic

N sites
152
245
284
322
57
128
53
148
28
57
63
86
79
80
169
159

Mean
elevation
260.82
248.47
276.32
309.61
257.74
268.69
351.84
341.67
423.27
293.69
343.35
339.35
202.25
186.05
218.78
272.99

Range
1912.06
2054.76
2054.76
1797.95
1150.28
1046.00
1165.74
1483.34
1908.91
2054.76
2054.76
1634.19
1141.92
899.54
903.72
1797.95

Std.
deviation
22.80
18.42
17.62
18.17
34.09
24.61
43.20
27.38
85.10
52.15
55.73
38.39
18.69
18.43
15.22
23.67

Variance
78,996.02
83,161.75
88,123.18
106,321.80
66,223.90
77,547.77
98,895.09
110,927.10
202,779.70
155,018.70
195,691.50
126,716.00
27,601.47
27,176.50
39,152.90
89,105.03

42

4 Human Subsistence and the Archaeofaunal Record of Late Paleolithic Southwest


250000

Mediterranean Iberia

Euro-Siberian Iberia
Southern France

Variance

200000

150000

100000

50000

0
Gravettian

Solutrean

Magdalenian

Epipaleolithic

Fig. 4.3 Changing variance by region through time. Variance values differ significantly by region
(F = 19.19, df = 11, p = 0.00)
Table 4.2 Differences in site elevation through time and by region; statistically significant values
are in bold
Region
All regions,
through time
Mediterranean Iberia
Euro-Siberian Iberia
Southern France

Elevation
between regions

ANOVA
F = 2.12,
p = 0.10
F = 2.40,
p = 0.07
F = 0.66,
p = 0.58
F = 3.414,
p = 0.02

Welchs test
F = 2.03,
p = 0.11
F = 2.42;
p = 0.07
F = 0.57;
p = 0.64
F = 2.98;
p = 0.03

KruskalWallis
H = 3.21,
p = 0.36
H = 6.44,
p = 0.09
H = 2.75,
p = 0.43
H = 1.75,
p = 0.63

F = 12.37,
p = 0.00

F = 11.81,
p = 0.00

H = 9.60,
p = 0.00

Bartletts
2 = 8.94,
p = 0.03
2 = 3.58,
p = 0.31
2 = 4.31,
p = 0.23
2 =
60.18,
p = 0.00
2 =
73.78,
p = 0.00

Region indicates sites location within proposed paleobioclimatic regions (e.g., Fig. 3.1)

Fernndez et al. 2005; Straus 1986). In Upper Paleolithic Euro-Siberian Iberia,


huntergatherers seem to have practiced a residential mobility strategy, and
resources seem to have been predictableeven in the coldest days of the LGM and
during the environmental reorganization that characterized the transition to the
Holocene.

4.3

Regionalization and Archaeofaunas

43

Sites in the Mediterranean climatic region of Iberia, on the other hand, have
lower site elevation variance (Table 4.1), suggesting less predictable resource distribution than in the high-elevation regions. As with the Euro-Siberian Iberian sites,
mean site elevation does not change significantly through time; site elevation variance does seem to increase through time (Fig. 4.5) but this trend is not statistically
significant (Table 4.2).
But the most dramatic difference is in southern France. In this region, site
elevation variance is quite low compared to that on the Iberian Peninsula (Table 4.1).
As in Mediterranean Iberia, site elevation variance seems to increase through time
(Fig. 4.3)and in this case the trend is statistically significant (Table 4.3). Site
elevation variance more than doubles in the Epipaleolithic over previous time periods, suggesting Epipaleolithic huntergatherers subsistence went through a major
reorganization concurrent with the climate variability associated with the eventual
transition to the Holocene (i.e., the Blling/Allerd warming and the Younger Dryas
cold event).
Site location analyses thus concur with the environmental data. The three regions
identified in Chap. 3 were home to huntergatherers with distinct mobility and subsistence patterns. In addition, with the exception of southern France in the Epipaleolithic,
the changing climate of the Late Paleolithic does not seem to have impacted hunter
gatherer mobility.

4.3

Regionalization and Archaeofaunas

Site location analyses can suggest differences in subsistence (as they do here), but
useful as they are, these are indirect methods. A more direct record of changing
human diets comes from the faunal and floral record from archaeological sites. While
some interesting archaeobotanical analyses have been conducted for the Late
Paleolithic (e.g., Allu et al. 2010; Gonzlez-Sampriz et al. 2009; Jones 2009), I
focus my analyses on the archaeofaunal record for two reasons: first, there is significantly more zooarchaeological data available for Paleolithic Europe than archaeobotanical; and second, methods of analysis for these two material types necessarily
differ, and so the data they produce can be very difficult to compare (though see
Bicho and Haws 2012).
I obtained faunal records from 85 sites containing 141 discrete deposits and compiled these into a database (Table 4.3). Archaeofaunal abundances for all datasets
were recorded as number of identified specimens (NISP) (Lyman 2008); I excluded
datasets for which only minimum numbers of individuals (MNI) were available due
to the problems with comparing these two types of data.
The database also excludes many taxa that likely were hunted in the Late
Paleolithic. I only included ungulate (hooved mammals) and lagomorph (rabbits and
hares) taxa in this analysis (Table 4.4). This was in part a matter of comparability:
for some of the assemblages, identification had been limited to these taxa. However,
excluding micromammals (for instance, rodents), birds, fish and shellfish also

Region
Southern
France

Site
La Garenne
Bois Ragot
Pont de Longues
Fourneau du Diable
Pont dAmbon
Combe Saunire
Blot
La Faurlie II
Rond du Barry
Ta
Roc de Marcamps
Beraud
La Madeleine
La Ferrassie
Cap Blanc
Laugerie-Haute
Saint Germain la Rivire
Moulin du Roc
Les Fieux
Combe Cullier
Le Flageolet
Gare de Couze
Moulin Neuf
Grotte XVI

Latitude
46.61
46.37
45.68
45.32
45.32
45.23
45.17
45.09
45.06
45.05
45.04
44.98
44.97
44.97
44.95
44.95
44.95
44.87
44.85
44.85
44.85
44.83
44.82
44.81

Longitude
1.50
0.68
3.20
0.63
0.60
0.88
3.48
0.92
3.86
4.50
0.49
3.68
1.02
0.93
1.15
1.00
0.33
0.92
1.70
1.57
1.08
0.73
0.29
1.16

Chronological periods
Mag
Epi
Mag
Grav
Epi
Sol
Mag
Epi
Mag, Epi
Mag, Epi
Mag
Epi
Mag, Epi
Grav
Mag
Sol, Mag
Mag
Epi
Grav
Mag
Grav, Mag
Epi
Mag, Epi
Grav, Sol, Mag

Source
Bayle et al. (2009)
Griggo (1995), Cochard (2004)
Fontana (2000)
Fontana (2001)
Jones (2006), Delpech (1983)
Castel et al. (2006)
Fontana (2000)
Cochard (2004)
Costamagno (1999)
Louchart and Soave (2002)
Delpech (1983)
Surmely et al. (2001)
Delpech (1983)
Delpech (1983)
Castel and Chadelle (2000)
Delpech (1983)
Costamagno (1999)
Jones (2006)
Gerbe (2010)
Delpech (1983)
Delpech (1983)
Delpech (1983)
Costamagno (1999)
Grayson et al. (2001)

Table 4.3 Archaeofaunas included in the database; Grav Gravettian, Sol Solutrean, Mag Magdalenian, Epi Epipaleolithic
Abbreviation
GAR
BR
PdL
DIA
PdA
CS
BLO
FAU
RdB
TAI
RdM
BER
MAD
FER
CB
LAU
GER
MdR
FIE
CUL
FLA
GdC
MNE
GXVI

44
4 Human Subsistence and the Archaeofaunal Record of Late Paleolithic Southwest

Region

Site
Murat II
Le Morin
Roc de Combe
Pgouri
Sainte Eulalie
Peyrugues
Gare de Conduch
Le Cuzoul
Saut du Loup
Baume dOullins
Abri de Gandil
Les Battuts
Salptrire
Le Plaisir
Fontaine du Pila
Duruthy
Dufaure
Canecaude
Gazel
Belvis
Moulin
Mas dAzil
Rhodes II
La Vache
Eglises

Latitude
44.80
44.80
44.77
44.62
44.58
44.52
44.48
44.45
44.37
44.34
44.05
44.00
43.94
43.62
43.61
43.52
43.50
43.34
43.32
43.14
42.98
42.92
42.86
42.82
42.76

Longitude
1.62
0.10
1.33
1.63
1.88
1.68
1.63
1.53
4.53
4.46
1.67
1.72
4.56
3.97
3.88
1.06
1.38
2.52
2.44
2.23
0.58
1.25
1.59
1.58
1.78

Chronological periods
Epi
Epi
Grav
Graettian, Mag, Epi
Sol, Mag, Epi
Grav, Sol, Mag
Mag
Sol, Mag
Epi
Epi
Mag
Grav
Grav, Sol, Mag, Epi
Epi
Epi
Mag, Epi
Mag, Epi
Mag
Mag
Epi
Epi
Mag
Epi
Epi
Epi

Source
Fat Cheung et al. (2014)
Delpech (1983)
Delpech (1983)
Sronie-Vivien (1995)
Delpech (1983)
Juillard (2009)
Castel et al. (2013)
Castel (2003)
Rillardon (2010)
Rillardon (2010)
Griggo (1997)
Delpech (1983)
Rillardon (2010)
Rillardon (2010)
Rillardon (2010)
Costamagno (2006)
Altuna and Mariezkurrena (1995)
Costamagno et al. (2009)
Costamagno et al. (2009)
Fontana (2000)
Costamagno et al. (2009)
Patou (1984)
Delpech (1983)
Pailhaugue (1998)
Delpech (1983)
(continued)

Abbreviation
MUR
MOR
RdC
PEG
SE
PEY
CON
CUZ
SdL
OUL
ABG
BAT
SAL
PLA
FdP
DUR
DUF
CAN
GAZ
BEL
MOU
MdA
RHO
VAC
EGL

4.3
Regionalization and Archaeofaunas
45

Region
EuroSiberian
Iberia

Site
Tito Bustillo
La Fragua
La Riera
Cueva Morin
Santimamine
Coimbre
Urtiaga
Ermittia
Aitzbitarte
Marizulo
El Miron
Ekain
Amalda
Erralla
Bolinkoba
Lezetxiki
Abauntz
Zatoya

Table 4.3 (continued)


Latitude
43.46
43.45
43.42
43.36
43.35
43.33
43.28
43.28
43.27
43.25
43.25
43.24
43.23
43.21
43.17
43.08
43.03
42.90

Longitude
5.07
3.42
4.86
3.86
2.64
4.68
2.32
2.36
1.89
1.98
3.45
2.26
2.20
2.18
2.61
2.53
2.04
1.21

Chronological periods
Mag
Epi
Grav, Sol, Mag, Epi
Grav, Sol, Mag
Mag, Epi
Mag
Mag, Epi
Sol, Mag, Epi
Sol, Mag, Epi
Epi
Mag, Epi
Mag, Epi
Grav, Sol
Mag
Grav, Sol, Mag
Sol, Epi
Sol, Mag
Epi

Source
Altuna (1976)
Marn Arroyo (2004)
Altuna (1986)
Yravedra and Gmez (2011)
Rofes et al. (2014), Quintana (2011)
lvarez Alonso et al. (2011)
Altuna (1972)
Altuna (1972)
Altuna (1963)
Altuna (1967, 1972)
Marn Arroyo (2010)
Altuna and Mariezkurrena (1984)
Yravedra Sainz de los Terreros (2005)
Altuna and Mariezkurrena (1985)
Castaos Ugarte (1983)
Altuna (1972)
Altuna et al. (2002)
Mariezkurrena and Altuna (1989)

Abbreviation
TB
FRA
RIE
CMO
SAN
COI
URT
ERM
AIT
MAR
MIR
EKA
AMA
ERR
BOL
LEZ
ABA
ZAT

46
4 Human Subsistence and the Archaeofaunal Record of Late Paleolithic Southwest

Region
Mediterranean
Iberia

Site
La Pea de las Forcas
lArbreda
Parco
Pea de Estebanvela
Cova des Blaus
Caldeirao
Lapa do Picareiro
Suo
Mallaetes
Cova Beneito
Tossal de la Roca
Santa Maira
Cendres
Cueva Ambrosio
Vale Boi
Zafarraya
Nerja

Latitude
42.17
42.16
41.87
41.42
39.84
39.65
39.53
39.30
39.01
38.80
38.79
38.73
38.69
37.82
37.09
36.95
36.76

Longitude
0.33
2.75
1.23
3.37
0.20
8.42
8.65
9.20
0.30
0.47
0.28
0.22
0.15
2.10
8.82
4.13
3.85

Chronological periods
Mag, Epi
Grav, Sol, Mag
Mag, Epi
Epi
Mag, Epi
Grav, Sol, Mag
Mag, Epi
Mag, Epi
Grav, Sol
Grav, Sol
Mag, Epi
Epi
Grav, Mag
Sol
Grav, Sol, Mag
Sol
Grav, Sol, Mag, Epi

Source
Utrilla (2014)
Estvez (1987)
Mangado et al. (2002)
Cacho (2013)
Villaverde and Martinez Valle (1995)
Davis et al. (2007)
Bicho and Haws (2012)
Bicho and Haws (2012)
Davidson (1989)
Villaverde and Martinez Valle (1995)
Cacho et al. (1995)
Morales Prez (2013)
Villaverde et al. (1997)
Yravedra Sainz de los Terreros (2005)
Manne et al. (2012)
Barroso Ruiz (2003)
Corts-Snchez et al. (2008)

Abbreviation
FOR
ARB
PCO
PdE
BLA
CAL
PIC
SU
MAL
BEN
TR
SM
CEN
AMB
VB
ZAF
NER

4.3
Regionalization and Archaeofaunas
47

48

4 Human Subsistence and the Archaeofaunal Record of Late Paleolithic Southwest

Table 4.4 Taxa included in the archaeofaunal database


Ungulates

Lagomorphs

Scientific name
Mammuthus primigenius
Rangifer tarandus
Saiga tatarica
Equus spp.
Bovinae
Cervus elaphus
Capra spp.
Rupicapra rupicapra
Capreolus capreolus
Sus scrofa
Oryctolagus cuniculus
Lepus spp.

Common name
Wooly mammoth
Reindeer
Saiga antelope
Horse/ass
Bison and aurochs
Red deer
Ibex
Chamois
Roe deer
Wild boar
Wild European rabbit
Hares

allowed me to be reasonably certain that the assemblages included represent human


subsistence activity rather than acquisition for non-subsistence activities (for instance,
manufacture of jewelry) or the subsistence activity of carnivores or raptors. While
lagomorph assemblages in particular are often the result of raptor activity, I excluded
all assemblages containing lagomorphs that have not been subject to a thorough
taphonomic analysis.

4.3.1

Similarity Indices and Cluster Analysis

Biogeographical studies often use analyses of compositional similaritytechniques


such as cluster analysis and non-metric multidimensional scaling (NMMDS)as a
means of testing for regionalization (Holt et al. 2013; Jones 2015; Kreft and Jetz
2010). These techniques are based on a quantititative assessment of how similar or
dissimilar the assemblages under consideration are to each other; such measures are
called similarity indices (or sometimes, dissimilarity indices). There are many different similarity indices available, all with particular advantages and disadvantages
(Jost et al. 2011), but the biggest distinction between them is whether they are based
on incidence (i.e., presence/absence data) or abundance. Abundance indices will be
more sensitive to differences in assemblage composition in time-averaged archaeofaunas (Jost et al. 2011), and so in this analysis I use the abundance-based Morisita
Horn index.
Similarity values, once calculated, can be used as the basis for analyses of compositional similarity. Cluster analysis, a multivariate statistical technique, groups sets of
objects (in this case zooarchaeological assemblages) based on their similarity values.
Typically results are presented in the form of a dendrogram, or relationship tree, which
can be difficult to read with large datasets. NMMDS, a technique which projects similarity data into three-dimensional space, can be used alongside cluster analyses to help

4.3

Regionalization and Archaeofaunas

49

visualize patterns. I use these two techniques together here to assess whether the
regions identified in Chap. 3 and supported by the elevation analysis are also present
in the archaeofaunas.

4.3.2

Archaeofaunal Similarity in the Upper Paleolithic

The cluster analysis of Upper Paleolithic archaeofaunas produced three distinct


groups (Fig. 4.4) which correspond with the paleoenvironmental regions identified
in Chap. 3. Cluster A contains the faunas from the Mediterranean bioclimatic region
of Iberia, Cluster B those from Euro-Siberian Iberia, and Cluster C those from
southern France. The cophenetic correlation coefficient, a measure of the degree of
fit of a classification to a set of data (Rohlf and Fisher 1968; Sarali et al. 2013), is
0.962, supporting the robustness of these groups. There is no sign of chronological
influence on these groups; assemblages do not sub-cluster by chronological period.
There are, however, two archaeofaunas from southern France which fall outside
cluster C: the Salptrire Magdalenian assemblage and the Les Battuts Gravettian
assemblage. Both are also anomalous in other ways. Les Battuts (which groups with
cluster B, the Euro-Siberian Iberia cluster) has a relatively small sample size
(NISP = 471), which may be influencing its relative abundance values. The
Salptrire Magdalenian (NISP = 1110), which groups with cluster A (the
Mediterranean Iberia cluster), is more complex. This assemblage is not formally
Magdalenian at all, although it dates to the same period as the Magdalenian faunas in this assemblage (around 19 kya BP; see Rillardon 2010); it has instead been
classified as Salptrien ancien, a classification known only from three sites in the
eastern Languedoc region.
The NMMDS for the Upper Paleolithic (Fig. 4.5a) confirms the regional clustering of archaeofaunas, but it also confirms Salptrire Magdalenian and Les Battuts
Gravettian assemblages as outliers. When NMMDS was conducted by sub-chronological period (i.e., Gravettian, Solutrean, Magdalenian, and Epipaleolithic), Les
Battuts did cluster with southern France. The Salptrire Magdalenian, however,
remained outside all regional clusters even when considered only with other
Magdalenian faunas.
Salptrire is the lone site from the French Mediterranean in the Upper Paleolithic
analysis, which makes it difficult to understand why this archaeofauna looks different. It may be that climate-driven environmental change began earlier in the French
Mediterranean than elsewhere in southern France. As discussed in Chap. 3, the
Mediterranean Sea may have decreased significantly in salinity after the LGM
(Gaspari et al. 2015). The Salptrire Magdalenian archaeofauna may simply be
reflecting wider environmental change. But it is also possible that the Salptrire
Magdalenian is anomalous for other reasons. Without additional Salptrien
ancien or Mediterranean Magdalenian-era assemblages in this analysis, it is impossible to answer this question.

Fig. 4.4 Cluster analysis of Upper Paleolithic archaeofaunas. (A) Contains archaeofaunas from
Mediterranean Iberia, (B) archaeofaunas from Euro-Siberian Iberia, and (C) archaeofaunas from
Southern France. Asterisk indicates extra-regional faunas

4.3

Regionalization and Archaeofaunas

51

Fig 4.5 NMMDS diagram for (a) Upper Paleolithic archaeofaunas and (b) Epipaleolithic
archaeofaunas

Overall, the correspondence of the archaeofaunal clusters for the Upper Paleolithic
with the environmental regions discussed in Chap. 3 is striking. This correspondence
is clearly visible when mapped (Fig. 4.6) as well as in the NMMDS. The similarity
analyses, like the elevation analysis, indicate human subsistence choices were
strongly regional throughout the Upper Paleolithic in Southwest Europe.
However, the similarity analyses also reveal important differences between regions
in archaeofaunal structure. The assemblages in cluster A (the Mediterranean Iberia
group) are remarkably similar to each other (Fig. 4.4). The first branch separates at a
similarity value of 0.70, and the second at 0.84. This is especially notable given the
broad geographic space over which these sites are scattered (Fig. 4.6); and it is in stark
contrast to both cluster B (the Euro-Siberian Iberia group; first branch separating at
0.30 and second at 0.40) and cluster C (the southern France group; first branch separating at 0.20 and second at 0.25). The archaeofaunas of Euro-Siberian Iberia and

52

4 Human Subsistence and the Archaeofaunal Record of Late Paleolithic Southwest

Fig. 4.6 Geographic location by cluster of (a) Upper Paleolithic archaeofaunas and (b) Epipaleolithic
archaeofaunas

4.3

Regionalization and Archaeofaunas

53

southern France are much more different from each other than are those of Mediterranean
Iberia.
In the context of the site elevation variance analysis, this finding makes sense.
The high site elevation variance found in the Upper Paleolithic of Euro-Siberian
Iberia may indicate a primarily residential mobility strategy, with people moving
frequently to exploit specific resources. Such a mobility strategy would necessarily
produce sites containing different taxa.
There is, however, even more dissimilarity between the faunas of Upper Paleolithic
southern Franceand this region had exceptionally low site elevation variance
(Fig. 4.3). How can this be? The answer may lie in the more restricted environments of
southern France. With a desert to the southwest and glaciers to the north, east, and
south, this region was strongly constrained by barriers, in contrast to the larger refugium of the Iberian Peninsula. Previous studies have suggested that the low site elevation variance in Upper Paleolithic southern France may reflect a tightly constrained
environmentand a similarly impoverished resource base (Jones 2007). In other
words, site elevation variance may be low in this region because people had nowhere to
go. At the same time, the distance between the archaeofaunas in cluster C indicates
little consistency in what animals were being accumulated from site to site in this
region. Consistency in site elevation combined with variance in the archaeofaunas suggests extreme patchiness in animal abundances in this regionfauna were probably not
evenly distributed over the landscape, but were rather found in resource patches.

4.3.3

Archaeofaunal Similarity in the Epipaleolithic

Although Epipaleolithic archaeofaunas do produce clear clusters (the cophenetic


correlation coefficient for this dataset is 0.94) for Euro-Siberian and Mediterranean
Iberia, the southern France cluster breaks down in this time period (Fig. 4.7), just as
the environments of southern France seem to have done. While the two Iberian
regions remain distinct (clusters B and C), seven of the southern France assemblages group on their own (cluster A); eight cluster with the Euro-Siberian Iberian
sites (cluster B), and nine with the Mediterranean Iberian group (cluster C). The
NMMDS also shows two tight Iberian clusters, but the southern France cluster is
large and encompasses both of the Iberian ones (Fig. 4.5b).
There are clear-cut regional patterns apparent in these data (Fig. 4.6b). All five of
the southern France assemblages in the Mediterranean Iberia group (cluster C) are
from the French Mediterranean. As today these two areas are part of the same
Mediterranean bioclimatic region (Rivas-Martnez et al. 2004), this likely indicates the expansion of Mediterranean biota out of the refugia of the central and
southern portions of the Iberian Peninsula to a distribution more similar to todays.
Similarly, the higher-elevation sites from the French Pyrenees and Massif Central
tend to cluster with the Euro-Siberian Iberian sites. However, the archaeofaunas
from the Atlantic southwest of Francethe Aquitaine regionare scattered across
all three groups (Fig. 4.7).

Fig. 4.7 Cluster analysis of Epipaleolithic archaeofaunas. (A) Contains seven of the archaeofaunas
from southern France; (B) comprises all the Euro-Siberian Iberian archaeofaunas as well as those
from high-elevation southern France; and (C) includes all the archaeofaunas from Mediterranean
Iberia and all the archaeofaunas from the French Mediterranean (indicated by asterisk) of southern
France as well as four faunas from the Dordogne region (indicated by double asterisk)

4.4 Persistent Regionalization: But with a Twist

55

The archaeofaunal data from southern France clearly support the hypothesis that
a major subsistence transition took place in this region at the end of the Paleolithic.
However, much of the variability in this group seems to be in sites from one area:
Aquitaine. Today, the Aquitaine region is marked by patchiness beyond the elevation and latitudinal gradient: vegetation is predominantly Atlantic deciduous forest,
but there are also pockets with a distinctly Mediterranean character. The distribution
of the Aquitaine archaeofaunas across all three clusters of Epipaleolithic faunas
suggests the modern patchy mosaic was present in this region from the Blling/
Allerd through the onset of the Holocenebut not earlier.

4.4

Persistent Regionalization: But with a Twist

I began this chapter with two questions: first, did Upper Paleolithic human subsistence track environmental regions, and second, how might these changes in human
hunting relate to the proposed Broad Spectrum Revolution for Southwest Europe?
While taxonomic abundance analyses of the archaeofaunas are needed to fully
answer the second question, the site location and archaeofaunal cluster analyses
show that indeed, hunting choices varied by environmental region in both the Upper
Paleolithic and the Epipaleolithic. In the Upper Paleolithic, when environmental
data suggest three discrete regions, there are three distinct patterns in both site location and in archaeofaunal composition. In the Epipaleolithic, when environmental
data indicate a major reorganization of the environments of southern France, the
two Iberian regions show distinct archaeofaunal composition but the southern
France region does not. Environmental change at the Pleistocene-Holocene transition has famously been described as a move from plaids (locally heterogeneous
environments) to stripes (Guthrie 1984). The change in southern French archaeofaunas suggests this was the case here.
In terms of a Broad Spectrum Revolution, the consistency through time in both
site location and archaeofaunal composition of the Iberian Peninsula does not seem
to support a revolution of any kind. Instead, the faunas of this region suggest
continuity, with some minor shifts in response to climate. Southern France, on the
other hand, is a different story. The major changes during the Epipaleolithic do suggest a significant, perhaps even revolutionary, transformation in hunting practice at
this time in this region. But was it a Broad Spectrum Revolution? To answer this
question, we must look at the composition of the faunas more closely.

References
Aldenderfer, M. S. (2008). High elevation foraging societies. In H. Silverman & W. H. Isbell
(Eds.), The handbook of South American archaeology (pp. 131143). New York: Springer.
Allu, E., Ibez, N., Saladi, P., & Vaquero, M. (2010). Small preys and plant exploitation by late
pleistocene huntergatherers. A case study from the Northeast of the Iberian Peninsula.
Archaeological and Anthropological Sciences, 2, 1124. doi:10.1007/s12520-010-0023-2.

56

4 Human Subsistence and the Archaeofaunal Record of Late Paleolithic Southwest

Altuna, J. (1963). Fauna de mamferos del yacimiento prehistrico de Aitzbitarte IV. Munibe, 15,
105124.
Altuna, J. (1967). Fauna de mamferos del yacimiento prehistrico de Marizulo (Urnieta),
Guipzcoa. Munibe, 3(4), 271298.
Altuna, J. (1972). Fauna de mamferos de los yacimientos prehistricos de Guipzcoa. Munibe, 24,
1465.
Altuna, J. (1976). Los mammiferos del yacimiento prehistorico de Tito Bustillo (Asturias). In J. A.
Moure Romanillo & M. Cano Herrera (Eds.), Excavaciones en la cueva de Tito Bustillo
(Asturias): Trabajos de 1975 (pp. 149189). Oviedo, Spain: Diputacin Provincial, Instituto de
Estudios Asturianos del Patronato Jos M.a Quadrado.
Altuna, J. (1986). The mammalian faunas from the prehistoric site of La Riera. In L. G. Straus &
G. A. Clark (Eds.), La Riera Cave: Stone age hunter-gatherer adaptations in northern Spain
(pp. 237274). Tempe, AZ: Arizona State University Anthropological Research Papers.
Altuna, J., & Mariezkurrena, K. (1984). Bases de subsistencia, de origen animal, de los pobladores
de Ekain. In J. Altuna & J. M. Merino (Eds.), El Yacimiento Prehistrico de la Cueva de Ekain.
San Sebastin, Spain : Sociedad de Estudios Vascos.
Altuna, J., & Mariezkurrena, K. (1985). Bases de subsistencia de los pobladores de Erralla:
Macromamferos. Munibe, 37, 87117.
Altuna, J., & Mariezkurrena, K. (1995). Les restes osseux de macromammifres. In L. G. Straus
(Ed.), Les derniers chasseurs de rennes du monde pyrnen: LAbri Dufaure: Un gisement
tardiglaciaire en Gascogne, fouilles 1980-1984 (pp. 181211). Paris: Mmoire de la Socit
Prhistorique Franaise.
Altuna, J., Mariezkurrena, K., & Elorza, M. (2002). Arquezoologa de los niveles paleolticos de
la Cueva de Abauntz (Arraiz Navarra). Saldvie: Estudios de prehistoria y arqueologa, 2, 126.
lvarez Alonso, D., Arrizabalaga Valbuena, ., Jord Pardo, J. F., & Yravedra Sainz de los
Terreros, J. (2011). La secuencia estratigrfica magdaleniense de la cueva de Combre
(Peamellera Alta, Asturias, Espaa). Frvedes, 7, 5764.
Banks, W. E., dErrico, F., Peterson, A. T., Vanhaeren, M., Kageyama, M., Sepulchre, P., et al.
(2008). Human ecological niches and ranges during the LGM in Europe derived from an application of eco-cultural niche modeling. Journal of Archaeological Science, 35, 481491.
doi:10.1016/j.jas.2007.05.011.
Banks, W. E., Zilhao, J., dErrico, F., Kageyama, M., Sima, A., & Ronchitelli, A. (2009). Investigating
links between ecology and bifacial tool types in Western Europe during the Last Glacial
Maximum. Journal of Archaeological Science, 36, 28532867. doi:10.1016/j.jas.2009.09.014.
Barroso Ruiz, C. (2003). El Pleistoceno superior de la cueva del Boquete de Zafarraya. Crdoba,
Spain: Junta de Andaluca, Consejera de Cultura.
Bayle, G., Crpin, L., & David, F. (2009). volution des comportements de subsistances des
Magdalniens navettes du Centre de la France: Les grands mammifres du site dhabitat de
La Garenne (Saint-Marcel, Indre) In J. Desprie, S. Tymula, & A. Rigaud (Eds.), Donnes
rcentes sur le Magdalnien de la Garenne (Saint-Marcel, Indre). La Place du Magdalnien
navettes en Europe (pp. 65104). Saint Marcel, France: ASSAAM.
Bicho, N., & Haws, J. (2012). The Magdalenian in central and southern Portugal: Human ecology
at the end of the Pleistocene. Quaternary International, 272273, 616. doi:http://dx.doi.
org/10.1016/j.quaint.2012.02.055.
Binford, L. R. (1980). Willow smoke and dogs tails: Hunter-gatherer settlement systems and
archaeological site formation. American Antiquity, 45, 420.
Cacho, C. (Ed.). (2013). Ocupaciones magdalenienses en el interior de la Pennsula Ibrica: La
Pea de Estevanvela (Aylln, Segovia). Madrid, Spain: CESIC, Junta de Castilla y Len.
Cacho, C., Fumanal, M. P., Lpez, P., Lpez, J. A., Prez Ripoll, M., Martnez Valle, R., et al.
(1995). El Tossal de la Roca (Vall dAlcal, Alicante). Reconstruccin paleoambiental y cultural
de la transicin del Tardiglaciar al Holoceno inicial. Recerques del Museu dAlcoi, IV, 11101.
Castaos Ugarte, P. M. (1983). Estudio de los macromamiferos del yacimiento prehistorico de
Bolinkoba (Abadiano-Vizcaya). KOBIE, Revista de Ciencias, 13, 261298.

References

57

Castel, J.-C. (2003). conomie de chasse et dexploitation de lanimal au Cuzoul de Vers (Lot) au
Solutren et au Badegoulien. Bulletin de la Socit Prhistorique Franaise, 100, 4165.
Castel, J.-C., & Chadelle, J.-P. (2000). Cap Blanc (Marquay, Dordogne), Lapport de la fouille de
1992 la connaissance des activits humaines et lattribution culturelle des sculptures/Cap
Blanc (Marquay, Dordogne), results of the 1992 excavation regarding human activities and
sculpture cultural attribution. Palo, 12, 6175.
Castel, J.-C., Kuntz, D., & Chauvire, F.-X. (2013). Lexploitation du monde animal au
Palolithique suprieur en Quercy: Un tat des connaissances. In M. Jarry, J.-P. Brugal, &
C. Ferrier (Eds.), Modalit doccupation et exploitation des milieux au Palolithique dans le
Sud-Ouest de la France: Lexemple du Quercy (pp. 395418). Les Eyzies de Tayac, France:
Socit du Muse National de Prhistoire et de la Recherche Archologique.
Castel, J.-C., Liolios, D., Laroulandie, V., Pike-Tay, A., Chauviere, F.-X., Chadele, J.-P., et al.
(2006). Animal resource exploitation by the Solutreans of Combe Saunire (Dordogne, France).
In M. Maltby (Ed.), Integrating Zooarchaeology (pp. 138152). Oxford, England: Oxbow.
Cochard, D. (2004). Les lporids dans la subsistance palolithique du Sud de la France. Thse de
doctorat, Universit de Bordeaux I, Talence.
Conard, N. J., Bolus, M., & Mnzel, S. C. (2012). Middle Paleolithic land use, spatial organization
and settlement intensity in the Swabian Jura, southwestern Germany. Quaternary International,
247, 236245. doi:10.1016/j.quaint.2011.05.043.
Corts-Snchez, M., Morales-Muiz, A., Simn-Vallejo, M. D., Bergad-Zapata, M. M., DelgadoHuertas, A., Lpez-Garca, P., et al. (2008). Palaeoenvironmental and cultural dynamics of the
coast of Mlaga (Andalusia, Spain) during the Upper Pleistocene and early Holocene. Quaternary
Science Reviews, 27, 21762193. doi:http://dx.doi.org/10.1016/j.quascirev.2008.03.010.
Costamagno, S. (1999). Stratgies de chasse et fonction des sites au Magdalnien dans le Sud de
la France. Thesis/dissertation (deg), Universit de Bordeaux I, , Talence.
Costamagno, S. (2006). Archozoologie desgrands mammifres des gisements de la falaise du
Pastou. In M. Dachary (Ed.), Les Magdalniens Duruthy: Qui taient-ils? Comment vivaientils? (pp. 2029). Mont-de-Marsan: Conseil gnral des Landes.
Costamagno, S., Laroulandie, V., Langlais, M., & Cochard, D. (2009). Exploitation du monde
animal sur le versant nord des Pyrnes durant le Tardiglaciaire. In J.-M. Fullola, N. Valdeyron,
& M. Langlais (Eds.), Els Pirineus i les rees circumdants durant el Tardiglacial: Mutacions i
filiacions tecnoculturals, evoluci paleoambiental (16000-10000 BP) (pp. 185209).
Puigcerd, Spain: Institut dEstudis Ceretans.
Council of Europe Directorate-General Enterprise and Industry (DG-ENTR) (2013).
EU-DEM. Retrieved May 3 2014, from http://www.eea.europa.eu/data-and-maps/data/ds_
resolveuid/HQ5ZD47A6I.
Davidson, I. (1989). La economa del final del Paleoltico en la Espaa oriental. (Vol. 85, Serie de
Trabajos Varios). Valencia: Servicio de investigacin prehistrica, Diputacin Provincial de
Valencia.
Davis, S., Robert, I., & Zilho, J. (2007). Caldeiro cave (Central Portugal)Whose home?
Hyaena, man, bearded vulture. Courier Forschungsinstitut Senckenberg, 259, 213226.
Delpech, F. (1983). La faune du Palolithique suprieur dans le Sud-Ouest de la France. Paris:
Cahier du Quaternaire.
Estvez, J. (1987). La fauna de lArbreda (sector alfa) en el conjunt de faunes del Pleistoc
Superior. Cypsela, VI, 7387.
Fat Cheung, C., Chevallier, A., Bonnet-Jacquement, P., Langlais, M., Ferri, J.-G., Costamagno,
S., et al. (2014). Comparaison des squences aziliennes entre Dordogne et Pyrnes: tat des
travaux en cours. In M. Langlais, N. Naudinot, & M. Peresani (Eds.), Les groupes culturels de
la transition Plistocne-Holocne entre Atlantique et Adriatique (pp. 1744). Paris: Socit
prhistorique franaise.
Fisher, L. E. (2002). Mobility, search modes, and food-getting technology. In B. Fitzhugh &
J. Habu (Eds.), Beyond foraging and collecting: Evolutionary change in hunter-gatherer settlement systems (pp. 157180). New York: Kluwer/Plenum.

58

4 Human Subsistence and the Archaeofaunal Record of Late Paleolithic Southwest

Fontana, L. (2000). Stratgies de subsistance au Badegoulien et au Magdalnien en Auvergne:


Nouvelles donnes. In G. Pion (Ed.), Le Palolithique suprieur rcent: Nouvelles donnes sur
le peuplement et lenvironnement (pp. 5965). Paris: Socit prhistorique franaise.
Fontana, L. (2001). tude archozoologique du Diable (Bourdeilles, Dordogne): Un exemple du
potntiel des faunes Palolithiques issues des fouilles anciennes. Palo, 13, 159182.
Gaspari, S., Scheinin, A., Holcer, D., Fortuna, C., Natali, C., Genov, T., et al. (2015). Drivers of
population structure of the bottlenose dolphin (Tursiops truncatus) in the Eastern Mediterranean
Sea. Evolutionary Biology, 42, 177190. doi:10.1007/s11692-015-9309-8.
Gerbe, M. (2010). Economie alimentaire et environnement en Quercy au Palolithique: tude des
associations fauniques de la squence des Fieux (Lot). Universit dAix-Marseille,
Gonzlez-Sampriz, P., Utrilla, P., Mazo, C., Valero-Garcs, B., Sopena, M. C., Morelln, M.,
et al. (2009). Patterns of human occupation during the early Holocene in the Central Ebro Basin
(NE Spain) in response to the 8.2 ka climatic event. Quaternary Research, 71, 121132.
doi:http://dx.doi.org/10.1016/j.yqres.2008.10.006.
Grayson, D. K., Delpech, F., Rigaud, J.-P., & Simek, J. F. (2001). Explaining the development of
dietary dominance by a single ungulate taxon at Grotte XVI, Dordogne, France. Journal of
Archaeological Science, 28, 115125.
Griggo, C. (1995). Significations paloenvironnementales des communauts animales Plistocne
reconues dans labri Suard et la Grotte de Bois-Ragot: Essai de quantification des variables
climatiques. Thse de doctorat, Universit de Bordeaux I, Talence.
Griggo, C. (1997). La faune magdalnienne de labri Gandil, Bruniquel (Tam-et-Garonne): tudes
palontologique, taphonomique et archozoologique. Palo, 9, 279294.
Grove, M. (2010). Logistical mobility reduces subsistence risk in hunting economies. Journal of
Archaeological Science, 37, 19131921. doi:10.1016/j.jas.2010.02.017.
Guthrie, R. D. (1984). Mosaics, allochemics, and nutrients: An ecological theory of Late
Pleistocene megafaunal extinctions. In P. S. Martin & R. G. Klein (Eds.), Quaternary extinctions: A prehistoric revolution (pp. 259298). Tucson, AZ: University of Arizona Press.
Hammer, ., Harper, D. A. T., & Ryan, P. D. (2001). PAST: Paleontological statistics software
package for education and data analysis. Palaeontologia Electronica, 4(1):9pp.
Henry, D. O. (1994). Prehistoric Cultural Ecology in Southern Jordan. Science, 265, 336341.
Holt, B. M. (2003). Mobility in Upper Paleolithic and Mesolithic Europe: Evidence from the lower
limb. American Journal of Physical Anthropology, 122, 200215. doi:10.1002/ajpa.10256.
Holt, B. G., Lessard, J.-P., Borregaard, M. K., Fritz, S. A., Arajo, M. B., Dimitrov, D., et al.
(2013). An update of Wallaces zoogeographic regions of the world. Science, 339, 7478.
doi:10.1126/science.1228282.
Jones, E. L. (2006). Prey choice, mass collecting, and the wild European rabbit (Oryctolagus
cuniculus). Journal of Anthropological Archaeology, 25, 275289.
Jones, E. L. (2007). Subsistence change, landscape use, and changing site elevation at the
Pleistocene-Holocene transition in the Dordogne of Southwestern France. Journal of
Archaeological Science, 34, 344353.
Jones, M. (2009). Moving north: Archaeobotanical evidence for plant diet in Middle and Upper
Paleolithic Europe. In J.-J. Hublin & M. P. Richards (Eds.), The evolution of Hominin diets
(Vertebrate paleobiology and paleoanthropology, pp. 171180). Dordrecht, Netherlands: Springer.
Jones, E. L. (2013). Mobility, settlement, and resource patchiness in Upper Paleolithic Iberia.
Quaternary International, 318, 4652. doi:http://dx.doi.org/10.1016/j.quaint.2013.05.027.
Jones, E. L. (2015). Archaeofaunal evidence of human adaptation to climate change in Upper
Paleolithic Iberia. Journal of Archaeological Science: Reports, 2, 257263. doi:http://dx.doi.
org/10.1016/j.jasrep.2015.02.008.
Jost, L., Chao, A., & Chazdon, R. L. (2011). Compositional similarity and (beta) diversity. In
A. E. Magurran & B. J. McGill (Eds.), Biological diversity: Frontiers in measurement and
assessment (pp. 6684). Oxford, England: Oxford University Press.
Juillard, F. (2009). La macro-faune de labri des Peyrugues: Biomtrie, odontomtrie. Prhistoire
du sud-ouest, 17, 155177.

References

59

Kelly, R. L. (1995). The foraging spectrum: Diversity in hunter-gatherer lifeways. Washington,


DC: Smithsonian Institution Press.
Kreft, H., & Jetz, W. (2010). A framework for delineating biogeographical regions based on species
distributions. Journal of Biogeography, 37, 20292053. doi:10.1111/j.1365-2699.2010.02375.x.
Louchart, A., & Soave, R. (2002). Changement dampleur de lexploitation des oiseaux entre le
Magdalnien et lAzilien: lexemple du Ta 2 (Brome). Quaternaire, 13, 297312.
Lyman, R. L. (2008). Quantitative paleozoology. Cambridge; New York: Cambridge University Press.
Mangado, J., Bartrol, R., Calvo, M. A., Nadal, J., Fullola, J.-M., & Petit, M.-. (2002). Evolucin
de los sistemas de captacin de recursos entre el Magdaleniense superior final y el Epipaleoltico
geomtrico de la Cueva del Parco (Als de Balaguer, La Noguera, Lleida). Zephyrus, 55,
143155.
Manne, T., Cascalheira, J., vora, M., Marreiros, J., & Bicho, N. (2012). Intensive subsistence
practices at Vale Boi, an Upper Paleolithic site in southwestern Portugal. Quaternary
International, 264, 8399. doi:http://dx.doi.org/10.1016/j.quaint.2012.02.026.
Mariezkurrena, K., & Altuna, J. (1989). Anlisis arqueozoolgico de los Macromamferos del
yacimiento de Zatoya (Abaurrea Alta, Navarra). Trabajos de Arqueologa Navarra, 8,
237266.
Marn Arroyo, A. B. (2004). Anlisis arqueozoolgico, tafonmico y de distribucin espacial de la
fauna de mamferos de la Cueva de la Fragua (Santoa, Cantabria). Munibe, 56, 1944.
Marn Arroyo, A. B. (2008). Mobility patterns and control of the eastern Cantabrian territory during late last glacial. Trabajos de Prehistoria, 65, 2945.
Marn Arroyo, A. B. (2009). Exploitation of the Montane Zone of Cantabrian Spain during the
Late Glacial: Faunal evidence from El Mirn Cave. Journal of Anthropological Research, 71,
69102.
Marn Arroyo, A. B. (2010). Arqueozoologa en el Cantbrica oriental durante la transicin
Plestoceno/Holoceno. Cantabria, Spain: Santander/PUbliCan, Ediciones Universidad de Cantabria.
Menndez Fernndez, M., Garca Snchez, E., & Quesada Lpez, J. M. (2005). Magdaleniense
inferior y territorialidad en la Cueva de La Gelga (Asturias). In N. F. Bicho (Ed.), O Paleoltico:
Actas do IV Congresso de Arqueologia Peninsular (pp. 6371). Faro, Portugal: Centro de
Estudos de Patrimnio, Departamento de Histria, Arqueologia e Patrimnio, Universidade do
Algarve.
Miller, R. (2012). Mapping the expansion of the Northwest Magdalenian. Quaternary International,
272273, 209230. doi:http://dx.doi.org/10.1016/j.quaint.2012.05.034.
Morales Prez, J. V. (2013). La transici del Paleoltic Superior Final/Epipaleoltic al Mesoltic
en el territori Valenci. Aportacions de lestudi zooarqueolgic del jaciment de Santa Maira
(Castell de Castells, Alacant). In A. Sanchis Serra & J. L. Pascual Benito (Eds.), Animals i
Arqueologia hui. I Jornades dArqueozoologia del Museu de Prehistria de Valncia
(pp. 181202). Museu de Prehistria: Valncia.
Morgan, C. (2008). Reconstructing prehistoric hunter-gatherer foraging radii: A case study from
Californias southern Sierra Nevada. Journal of Archaeological Science, 35, 247258.
Pailhaugue, N. (1998). Faune et saisons doccupation de la salle Monique au Magdalenien
Pyreneen, grotte de la Vache (Alliat, Ariege, France). Quaternaire, 9, 385400.
Patou, M. (1984). La faune de la galerie Rive Droite du Mas dAzil (Arige): Donnes paloclimatiques et palthnographiques. Bulletin de la Socit Prhistorique Franaise, 81, 311319.
Quintana, J. C. L. (2011). La cueva de Santimamie: Revisin y actualizacin (2004-2006).
Bilbao, Spain: Diputacin Foral de Bizkaia.
Rillardon, M. (2010). Environnement et subsistance des derniers chasseurs-cueilleurs dans la
basse valle du Rhne et ses marges du Plniglaciaire suprieur (20 ka BP) lOptimum climatique (8 Ka BP). Provence, France: Universit Aix-Marseille I.
Rivas-Martnez, S., Penas, A., & Daz, T. E. (2004). Bioclimatic map of Europe, bioclimates. Len,
Spain: Cartographic Service, University of Len, Spain.
Rivas-Martinez, S., & Rivas-Saenz, S. (2009). Worldwide bioclimatic classification system,
19962009. Retrieved from http://www.globalbioclimatics.org2015.

60

4 Human Subsistence and the Archaeofaunal Record of Late Paleolithic Southwest

Rofes, J., Murelaga, X., Martnez-Garca, B., Bailon, S., Lpez-Quintana, J. C., Guenaga-Lizasu,
A., et al. (2014). The long paleoenvironmental sequence of Santimamie (Bizkaia, Spain):
20,000 years of small mammal record from the latest Late Pleistocene to the middle Holocene.
Quaternary International, 339340, 6275. doi:http://dx.doi.org/10.1016/j.quaint.2013.05.048.
Rohlf, F. J., & Fisher, D. L. (1968). Test for hierarchical structure in random data sets. Systematic
Zoology, 17, 407412.
Sarali, S., Dogan, N., & Dogan, I. (2013). Comparison of hierarchical cluster analysis methods by
cophenetic correlation. Journal of Inequalities and Applications, 2013, 18.
Sronie-Vivien, M.-R. (1995). La grotte de Pgouri, Caniac-du-Causse (Lot). Prhistoire
Quercinoise, Supplment no., 2.
Straus, L. G. (1986). Late Wrm adaptive systems in Cantabrian Spain: The case of eastern
Asturias. Journal of Anthropological Archaeology, 5, 330368.
Straus, L. G., Bicho, N., & Winegardner, A. C. (2000). The Upper Paleolithic settlement of Iberia:
First-generation maps. Antiquity, 74, 553566.
Surmely, F., Quinqueton, A., & Virmont, J. (2001). Le gisement pipalolithique ancien de la
grotte Braud Saint-Privat-dAllier (Haute-Loire, France).
Yravedra Sainz de los Terreros, J. (2005). Patrones de aprovechamiento de recursos animales en
el Pleistoceno Superior de la Pennsula Ibrica: Estudio tafonmico y zooarqueolgico de los
yacimientos de la cueva del Esquilleu, la cueva de Amalda, la Pea de Estebanvela y Cueva
Ambrosio. Madrid, Spain: Universidad Nacional de Educacin a Distancia.
Yravedra Sainz de los Terreros, J., & Gmez Castanedo, A. (2011). Anlisis de los procesos
tafonmicos de cueva morn: Primeros resultados de un estudio necesario. Zephyrus, LXVII,
6990.
Utrilla, P. (Ed.). (2014). La Pea de las Forcas (Graus, Huesca): Un asentamiento estratgico en
la confluencia del sera y el Isbena. Zaragoza, Spain: Prensas Universitarias de Zaragoza.
Vermeersch, P. M. (2014). Radiocarbon Palaeolithic Europe Database, Version 17. Retrieved from
http://ees.kuleuven.be/geography/projects/14c-palaeolithic/index.html.
Villaverde, V., Martnez Valle, R., Guillem-Calatayud, P. M., Bada, E., Zalbidea, L., & Garca, R.
(1997). Els nivells magdalenians de la Cova de les Cendres (Teulada, Moraira). Resultats del
sondeig del quadre A-17. Aguaits, 1314, 77115.
Villaverde, V., Martinez-Valle, R., Guillem, P. M., & Fumanal, M. P. (1996). Mobility and the role
of small game in the Middle Paleolithic of the central region of the Spanish Mediterranean: A
comparison of Cova Negra with other Paleolithic deposits. In E. Carbonell & M. Vaquero
(Eds.), The last Neandertals, the first anatomically modern humans: A tale about the human
diversity: Cultural change and human evolution: The crisis at 40 ka BP (pp. 267288).
Tarragona, Spain: Universitat Rovira i Virgili.
Villaverde, V., & Martinez Valle, R. (1995). Caractersticas culturales y econmicas del final del
Paleoltico superior en el Mediterrneo espaol. In V. Villaverde (Ed.), Los Ultimos Cazadores
de los Tiempos Glaciares. Transformaciones culturales y econmicas durante el Tardiglaciar
y el inicio del Holoceno en el mbito mediterrneo (pp. 79117). Alicante: Instituto de Cultura
Juan Gil-Albert.
Wygal, B., & Heidenreich, S. (2014). Deglaciation and human colonization of northern Europe.
Journal of World Prehistory, 27, 111144. doi:10.1007/s10963-014-9075-z.

Chapter 5

Archaeofaunal Diversity and Broad Spectrum


Diets in Late Paleolithic Southwest Europe

5.1

Introduction

The Broad Spectrum Revolution is, at its heart, concerns changing dietary diversity.
As discussed in Chap. 2, Flannerys original model (1969) suggested human diets in
the Near East initially were tightly focused (i.e., had little diversity), broadened
when human demographic pressure forced widening of the dietary niche, and then
narrowed again with the widespread transition to agricultural economies. More
recent work rooted in the prey choice model has reframed this slightly, proposing
that early diets focused on high-return prey items. As human demographic pressure
increased, resource depression occurred and those high-return prey items became
scarce; therefore people were forced to expand their diets to include smaller and
lower-return prey items such as small game, plants, and seeds. By this reasoning, if
a Broad Spectrum Revolution took place in Southwest Europe at the end of the
Pleistocene, smaller prey should be added to the diet in the Epipaleolithic, but larger
prey should remain. Thus the total number of taxa in the diet should increase in a
gradual and linear fashion.
To test this hypothesis, ideally the full suite of paleodietary remainsincluding
fish, birds, shellfish and paleobotanical remainswould be analyzed. Unfortunately,
the challenges involved with such an endeavor are many. As discussed in Chap. 4,
plants and animals are deposited differently in the archaeological record and their
analysis requires specific and often incompatible methods of quantification. Shellfish,
too, are counted differently than vertebrate taxa, making analyses that involve a combination of plants, shellfish, and vertebrate fauna a challenge (though see Stiner and
Munro 2002 for one approach to this). For this reason many people who study the
development of broad spectrum diets focus their attention on vertebrate fauna (even
if they discuss invertebrates in a qualitative way), assuming that increasing diet
breadth in vertebrate fauna will reflect changes in the entire diet (e.g., Conrad 2015;
Corts-Snchez et al. 2008; Lupo 2007; Rillardon and Brugal 2014; Stutz et al. 2009;
Zhang et al. 2013).
The Author(s) 2016
E.L. Jones, In Search of the Broad Spectrum Revolution in Paleolithic Southwest
Europe, SpringerBriefs in Archaeology, DOI 10.1007/978-3-319-22351-3_5

61

62

5 Archaeofaunal Diversity and Broad Spectrum Diets in Late Paleolithic Southwest

However, the problem of comparing dissimilar data types is far from the only
challenge involved in testing the Broad Spectrum Revolution. A regional hypothesis
such as this one requires a meta-analysis, or an analysis involving multiple datasets.
Meta-analyses, while powerful and rigorous tools, also present a number of potential serious biases. Two relevant potential biases in this case are the many older
collections in this dataset and taxonomic specialization in identification.
While there are numerous issues with the use of older collections in metaanalyses (Jones and Gabe 2015), the one of most concern here is collection bias. In
the early years of archaeology, many archaeologists only collected the bones of
mammals assumed to have been part of the human diet and ignored other osteological material (Peres 2010). In the early years of Upper Paleolithic archaeology,
screening was variable at best, although since the 1960s fine-screening has become
standard (Lopez-Martinez 2008; Pokines 2000). Lack of screening, variable screening, or even screening with larger mesh can bias assemblages towards larger species
and away from birds, fish, and other small taxa (Cannon 1999; Nagaoka 2005;
Schaffer and Sanchez 1994; Stahl 1996).
This age-based differential recovery bias could be dealt with by excluding older
collections from the analysis (Jones and Gabe 2015). However, there is another
problem with this dataset that would render such an action futile, and that is taxonomic specialization in identification. Among zooarchaeologists, there are more
individuals who focus on ungulates and lagomorphs than those who focus on
rodents, birds, fish, or invertebrates. The result of this is that there are many zooarchaeological datasets containing information about the ungulate and lagomorph
component of archaeofaunas, but few with information about the rest of the zooarchaeological assemblage. While there are fully reported Late Paleolithic zooarchaeological assemblages, there are also many for which there is no information
about even the presence or absence of birds, fish, or shellfish in the assemblage,
let alone their taxonomic abundance. Even when these data are reported, it is
often without any taphonomic analysis, and thus whether these taxa entered the
archaeological record because they were hunted by people, collected by some
other predator, or introduced by a different depositional process is unclear. Given
these challenges, testing for changing diversity using the full suite of taxa would
likely tell us about differences in archaeological practice rather than about changing
Late Paleolithic diets.
Fortunately, there is a long tradition of zooarchaeological work on ungulate and
lagomorph archaeofaunas in Late Paleolithic Southwest Europe (e.g., Altuna 1972;
Delpech 1983), and because of this there are many robustly identified assemblages.
As discussed in Chap. 4, the lagomorph component of archaeofaunas can be influenced by nonhuman predators; it can also be impacted by archaeological practice (i.e.,
collection methods and screen-size variability). Because of the pioneering work of
Altuna, Delpech, and their students, assemblages that are problematic in this regard
have been identified as such and thus can be ruled out of analysis as discussed in Sect.
4.3. In addition, there are established methods for exploring changing dietary diversity even while focusing on the larger mammalian fauna (Grayson and Delpech 1998,
2002; Grayson et al. 2001; Jones 2015; Munro 2004).

5.2 Measuring Changing Archaeofaunal Diversity

63

I thus use the archaeofaunal database (Sect. 4.3; Table 4.3) as the basis for the
diversity analyses presented in this chapter. As in Chap. 4 I consider faunas by the
paleobioregions proposed in Chap. 3 (Mediterranean Iberia, Euro-Siberian Iberia,
and southern France.)

5.2

Measuring Changing Archaeofaunal Diversity

Although we often speak of diversity as a single concept, it in fact has many


components (Magurran and McGill 2011). Zooarchaeologists commonly focus on
two aspects of diversity, richness and evenness (Lyman 2008). I use measures of
both, as well as the proportion of lagomorphs, to assess changing diversity in the
Late Paleolithic Southwest Europe archaeofaunal dataset (i.e., Table 4.3).

5.2.1

Diversity 1: Richness

Richness is the number of types in a set; in a biological context, it is usually the


number of species or taxa present. In zooarchaeological analysis, richness is generally measured by the number of taxa (NTAXA) in an assemblage (Grayson and
Delpech 1998; Jones 2013b; Lyman 2008). While this seems an intuitive way to
measure diet breadth, in practice NTAXA is infrequently used, as this measure can
sometimes be misleading.
The largest problem with NTAXA as a measure of richness is that it is strongly
controlled by sample size (the number of identified specimens, or NISP). It has been
well-demonstrated that as NISP increases, so does NTAXA (Grayson 1984; Lyman
2008, 2015). In some studies of diet breadth, the authors have made use of
this, identifying different NISP/NTAXA relationships through time and/or space
(e.g., Grayson et al. 2001; Nagaoka 2002). In other instances, researchers have contextualized NTAXA with additional data (e.g., Jones 2013b; Morrison and Hunt
2007; Nagaoka 2001) and with attention to the potential role of sample size in structuring the data.
In the Late Paleolithic Southwest Europe faunal dataset, there is a significant
(p = 0.00) but not a strong (r2 = 0.12) relationship between NISP and NTAXA
(Fig 5.1); this same relationship holds when the data are analyzed by region and/or
by chronological period. Sample size thus seems to be a consistent influence across
space and timebut it not the only significant influence on these data.
NTAXA does pattern by region, particularly when time is taken into account
(Table 5.1). While the Mediterranean Iberia and southern France groups have similar mean NTAXA values across the Upper Paleolithic and Epipaleolithic, in EuroSiberian Iberia NTAXA decreases in the Epipaleolithic (Fig. 5.2). Statistical tests of
whether these samples originate from the same distribution support this observation.
Both ANOVA and KruskalWallis (a nonparametricand thus more appropriate for

64

5 Archaeofaunal Diversity and Broad Spectrum Diets in Late Paleolithic Southwest

Fig. 5.1 Relationship between NISP and NTAXA for the archaeofaunal dataset (r2 = 0.12, p = 0.00;
slope = 0.05, intercept = 0.75). Both values have been ln-transformed

Table 5.1 Changing richness (NTAXA) and evenness (1/D) through time and by region
Region
All regions

Mediterranean
Iberia

Euro-Siberian
Iberia

Southern France

Period
Gravettian
Solutrean
Magdalenian
Epipaleolithic
Gravettian
Solutrean
Magdalenian
Epipaleolithic
Gravettian
Solutrean
Magdalenian
Epipaleolithic
Gravettian
Solutrean
Magdalenian
Epipaleolithic

N
19
23
52
43
7
8
10
14
3
7
12
10
9
8
29
19

Mean
NTAXA
6.88
7.31
7.19
7.05
6.29
6.63
6.00
5.86
8.00
7.57
8.50
6.40
7.33
8.00
7.55
7.74

NTAXA
Variance
3.98
3.16
3.01
2.82
1.90
1.13
2.20
1.98
1.00
1.62
0.82
2.27
5.50
4.57
2.90
2.43

Mean
1/D
1.94
1.95
2.02
1.79
1.71
1.48
1.39
1.70
2.65
2.30
2.27
2.27
1.52
1.97
2.41
2.34

1/D
Variance
0.64
0.77
0.69
0.80
0.40
0.17
0.24
0.21
0.08
1.42
1.08
0.10
0.93
0.17
0.54
1.22

zooarchaeological abundance dataform of ANOVA; see Lyman 2008, Wolverton


et al. 2014, and Zar 1999) analyses identify a significant difference through time in
Euro-Siberian Iberia, but not in either of the other two regions (Table 5.2). While
sample size could be influencing this result, a Pearson correlation analysis suggests
otherwise (r = +0.02, p = 0.93). The decrease in NTAXA in Euro-Siberian Iberia in
the Epipaleolithic appears to be real.

5.2 Measuring Changing Archaeofaunal Diversity

65

12
10
8
6
4

Mediterranean Iberia Epipaleolithic

Mediterranean Iberia Upper Paleolithic

Euro-Siberian Iberia Epipaleolithic

Euro-Siberian Iberia Upper Paleolithic

Southern France Epipaleolithic

Southern France Upper Paleolithic

Fig. 5.2 Changing richness (NTAXA) through time by region

Table 5.2 Differences in NTAXA through time and by region; statistically significant values are
in bold
Region
All regions, through time
Mediterranean Iberia
Euro-Siberian Iberia
Southern France

ANOVA
F = 1.30, p = 0.28
F = 0.27, p = 0.84
F = 7.02, p = 0.00
F = 0.54, p = 0.67

KruskalWallis
H = 4.22, p = 0.23
H = 0.73, p = 0.85
H = 12.49, p = 0.00
H = 1.37, p = 0.71

Bartletts
2 = 0.40, p = 0.94
2 = 1.60, p = 0.66
2 = 3.82, p = 0.28
2 = 4.06, p = 0.26

Is this an indicator of increasing specialization in the Euro-Siberian paleobioclimatic region of Iberia? Likey not. This region lost a prey type in the Epipaleolithic;
the reindeer (Rangifer tarandus) became locally extinct here due to warming climates (lvarez-Lao and Garca 2010; Gmez-Olivencia et al. 2014; Sommer et al.
2013). While reindeer never comprise a large portion of the Euro-Siberian Iberian
Upper Paleolithic archaeofaunas in this set, they are consistently present (see Table
5.3). By contrast, this taxon is absent from all Epipaleolithic Euro-Siberian faunas.
Because NTAXA measures maximum diet breadth (Nagaoka 2001), the loss of the
reindeer can (and here does) appear as a significant change.
The extirpation of the reindeer in Euro-Siberian Iberia, then, caused a reduction
in the large mammal portion of hunter-gatherer diet breadth. It is possible that diets

66

5 Archaeofaunal Diversity and Broad Spectrum Diets in Late Paleolithic Southwest

Table 5.3 Mean proportion lagomorphs, reindeer, and warmer-adapted taxa (Capreolus capreolus,
Cervus elaphus, and Sus scrofa) through time and by region

Region
All regions

Mediterranean
Iberia

Euro-Siberian
Iberia

Southern France

Period
Gravettian
Solutrean
Magdalenian
Epipaleolithic
Gravettian
Solutrean
Magdalenian
Epipaleolithic
Gravettian
Solutrean
Magdalenian
Epipaleolithic
Gravettian
Solutrean
Magdalenian
Epipaleolithic

N
19
23
52
43
7
8
10
14
3
7
12
10
9
8
29
19

Mean
proportion
reindeer
0.30
0.29
0.29
0.12
0
<0.01
0
0
<0.01
0.01
0.02
0
0.60
0.80
0.57
0.27

Mean
proportion
lagomorphs
0.27
0.25
0.21
0.39
0.72
0.75
0.76
0.81
0.00
0.02
0.01
0.00
0.03
0.03
0.06
0.25

Mean
proportion
warmer taxa
0.17
0.15
0.17
0.25
0.14
0.07
0.07
0.04
0.31
0.39
0.49
0.69
0.14
0.01
0.03
0.18

widened to nonmammalian taxa to compensate (see, for example, Marn Arroyo


2013); while I cannot test this hypothesis with these data, I will examine whether
the impact of this event can be seen in other aspects of archaeofaunal structure later
in this chapter.
A final note on the NTAXA analysis: these numbers show no change in southern
France, the region in which (based on the analyses in Chap. 4) we would most
expect to see an increase in diet breadth. The reindeer was not extirpated in southern
France until the Holocene (Sommer et al. 2013). The overall proportion of reindeer
in the southern French archaeofaunas does decrease in the Epipaleolithic (Table 5.1),
suggesting a decline in local abundance prior to extirpation, but the fact that this
resource was still at least sporadically availableno matter how infrequentlyis
why NTAXA values remain constant here.

5.2.2

Diversity 2: Evenness

Evenness is the degree to which taxa are equally distributed within a setfor
instance, in an assemblage with 10 taxa, if each taxon had a NISP of 5 (for a total
NISP of 50), the assemblage would be completely even. By contrast, if one taxon
had a NISP of 41 and the rest had a NISP of 1 each (again, for a total NISP of 50),
the assemblage would be uneven.
While many find evenness to be a less intuitive measure of diversity, in zooarchaeological contexts it is extremely useful. Archaeological assemblages are time-

5.2 Measuring Changing Archaeofaunal Diversity

67

averaged; they contain faunas deposited over many years- sometimes thousands of
years. The NTAXA of an assemblage will count a taxon represented by one sample
equally as one represented by a thousand, even though that one sample may represent a single rare event. This may be the case with reindeer in Upper Paleolithic
Euro-Siberian Iberia, for instance. Evenness avoids this problem because it takes a
taxons relative abundance into account.
Generally, low evenness, like low NTAXA, is seen as an indicator of narrow diet
breadth. Low evenness can characterize narrow diets focused on a few higherranked prey types, and increasing evenness can indicate resource depression (e.g.,
Nagaoka 2001). However, differences in site type, deposition history, and the ways
the analyst constrains the data can impact evenness and produce contradictory
results (Jones 2004a). For example, in protohistoric assemblages from New Mexico,
USA, evenness remains constant before and after the introduction of domestic fauna
of Old World origins, even though a closer look at those assemblages shows that
small mammals drop out of the diet and are replaced by larger onesa classic
example of narrowing diet breadth (Jones 2016). In short, evenness must be contextualized, just as NTAXA must (Jones 2004a).
Evenness values for the Late Paleolithic Southwest Europe faunal dataset can be
seen in Table 5.1. I used the inverse of Simpsons dominance index (1/D) to measure
evenness as this index is independent of NTAXA and less prone to sample size issues
than other evenness indices (Jones 2004a; Magurran 2004). Sample size can still influence 1/D, but a Pearson correlation showed no significant relationships between sample
size and the values presented here (overall 1/D: r = 0.16, p = 0.35; 1/D with lagomorphs excluded; r = 0.16, p = 0.36; mean 1/D through time: r = +0.02, p = 0.92).
Evenness does seem to pattern by region in these data, though not through time
(Fig. 5.3). Mediterranean Iberia has low evenness overall, while evenness is higher
(and more variable) in Euro-Siberian Iberia and southern France. ANOVA and
KruskalWallis tests confirm the difference between regions (Table 5.3). The reason
for the low evenness in Mediterranean Iberia likely is the high proportion of the
wild European rabbit (Oryctolagus cuniculus) in these faunas (Table 5.1); the rabbit
is often present in Mediterranean Iberian archaeofaunas by the thousands. Indeed,
when lagomorphs (both rabbits and hares, Lepus spp.) are removed from the dataset, the difference in evenness between regions disappears (Table 5.3).
This is, again, not what we would expect based on the analyses in Chap. 4, which
identified profound differences between southern France and the Iberian regions
and suggested a dietary transition in southern France during the Epipaleolithic. The
comparison with protohistoric New Mexico mentioned earlier, however, provides a
possible explanation.
In the New Mexican case, significant changes in diet were obscured as a result of
the introduction of new taxa into the available set (Jones 2013b, 2016). While there is
no evidence for anthropogenic introductions in this case, there is evidence for environmentally driven change in available taxa (Sommer et al. 2013; Sommer et al. 2008).
Cold-adapted taxa such as reindeer and saiga antelope (Saiga tatarica) were in the
process of extirpation, while warmer-adapted taxa such as roe deer (Capreolus capreolus), wild boar (Sus scrofa), and red deer (Cervus elaphus) were increasing in
abundance (Delpech 1999).

5 Archaeofaunal Diversity and Broad Spectrum Diets in Late Paleolithic Southwest

68
5
4.5
4
3.5
3
2.5
2
1.5
1

Mediterranean Iberia Epipaleolithic

Mediterranean Iberia Upper Paleolithic

Euro-Siberian Iberia Epipaleolithic

Euro-Siberian Iberia Upper Paleolithic

Southern France Epipaleolithic

Southern France Upper Paleolithic

0.5

Fig 5.3 Changing evenness (1/D) through time by region

This turnover in taxonomic composition could keep evenness constant across the
Upper PaleolithicEpipaleolithic transition in these data. And indeed, this seems to
be what is happening (Fig. 5.4). Warmer-adapted taxa increase significantly in the
Epipaleolithic in southern France, and this offsets the concurrent decrease in coldadapted taxa (Table 5.3). While the proportion of these taxa in Euro-Siberian Iberian
archaeofaunas also appears to increase from the Upper Paleolithic to the
Epipaleolithic, the trend is not statistically significant; and they actually decrease
(though again, not significantly) in Mediterranean Iberia.

5.2.3

Lagomorphs, Diet Breadth, and Changing Diversity

As mentioned earlier, one of the most widely-observed differences between Upper


Paleolithic Mediterranean Iberian archaeofaunas and those from Euro-Siberian
Iberia and southern France is the high abundance of the wild European rabbit in the
former, and the virtual absence of this taxon in the latter (Hockett and Haws 2002;
Jones 2012, 2015; Rillardon and Brugal 2014; Villaverde et al. 1998). In the
Epipaleolithic, this changes, and the rabbit becomes common in a subset of southern French archaeofaunas (Jones 2006; Rillardon and Brugal 2014).

5.2 Measuring Changing Archaeofaunal Diversity

69

1
Upper Paleolithic
Epipaleolithic

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0

Southern France

Euro-Siberian Iberia

Mediterranean Iberia

Fig 5.4 Average proportion of warmer-adapted taxa (calculated as NISP Cervus elaphus + Capreolus capreolus + Sus scrofa/total mammalian NISP) in Upper Paleolithic and
Epipaleolithic archaeofaunas by region. A KruskalWallis test confirms a change in lagomorph
representation from the Gravettian through the Epipaleolithic in southern France (H = 14.19,
p = 0.00) but found no significant difference in either Euro-Siberian (H = 6.60, p = 0.09) or
Mediterranean (H = 4.37, p = 0.22) Iberia

The increasing representation of rabbits in Epipaleolithic France is often cited


as evidence for a broad spectrum revolution in this region (Jones 2004b). This is
consistent with a more global assumption that lagomorphs in general, and rabbits
(both Oryctolagus and its distant relative Sylvilagus spp., the American cottontail)
in particular, are lower-ranked prey. Rabbits are small; they are fast and so (in theory) require more energetic investment to hunt successfully; they are assumed to be
nutritionally deficient due to low fat content; and historically they have been seen as
starvation food (Callou 2003; Nelson and Schollmeyer 2003; Schollmeyer 2011;
Speth and Spielmann 1983; Stiner and Munro 2002; Ugan 2005). Problems associated with excessive protein intake are even colloquially known as rabbit starvation
syndrome (e.g., Bilsborough and Mann 2006). For this reason the proportion of
rabbits and other lagomorphs in the diet (calculated in a myriad of ways) has become
a widely-used proxy for resource depression among zooarchaeologists.
In recent years, however, a number of researchers have turned a critical eye to the
assumption that increasing representation of rabbits in zooarchaeological assemblages necessarily indicates resource stress (Dean 2007b; Driver and Woiderski
2008; Hockett and Bicho 2000; Hockett and Haws 2002, 2003; Jones 2006).
Increasing rabbit abundance can indicate landscape change, both anthropogenic and

70

5 Archaeofaunal Diversity and Broad Spectrum Diets in Late Paleolithic Southwest

non (Driver and Woiderski 2008; Jones 2012); it can reflect human interactions
with predator taxa such as coyotes; it can reflect changes in hunting technique
(Jones 2006); and it can reflect changes in diet breadth (Dean 2007a). In short, as
with richness and evenness, changing proportions of rabbits in zooarchaeological
assemblages must be contextualized to be meaningful.
A critical piece of context for the Southwest Europe case is climate and associated changing biogeography. As discussed in Chap. 3, the European rabbit seems to
have been only sporadically present north of the Pyrenees prior to the Blling/
Allerd warming at ca. 13 kya. Thus, any increase in representation of this taxon in
southern France may indicate the addition of a new potential diet resource to the
landscape rather than a resource stress-induced expansion of diet breadth. Hares,
however, were available in southern France throughout the Upper Paleolithic
(Cochard 2004), and (based on size and speed) could also be considered lowerranked prey. I therefore consider changing lagomorph representation, calculated as
(NISP Oryctolagus + NISP Lepus spp.)/total mammalian NISP, through time as a
potential measure of increasing diet breadth in the Late Paleolithic Southwest
Europe archaeofaunal dataset.
The lagomorph abundance analysis shows that in these assemblages, lagomorphs
are indeed differentially represented by region (Table 5.3, Fig. 5.5). They are

Upper Paleolithic
Epipaleolithic

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0

Southern France

Euro-Siberian Iberia

Mediterranean Iberia

Fig 5.5 Average proportion of lagomorphs (calculated as NISP Oryctolagus cuniculus + Lepus
spp./total mammalian NISP) in Upper Paleolithic and Epipaleolithic archaeofaunas by region.
A KruskalWallis test confirms a change in lagomorph representation from the Gravettian through
the Epipaleolithic in southern France (H = 13.50, p = 0.00) but found no significant difference in
either Euro-Siberian (H = 2.65, p = 0.09) or Mediterranean (H = 0.20, p = 0.67) Iberia

5.2 Measuring Changing Archaeofaunal Diversity

71

Table 5.4 Differences in 1/D through time and by region; statistically significant values are in bold

With
lagomorphs

Without
lagomorphs

Region
All regions,
through time
Mediterranean
Iberia
Euro-Siberian
Iberia
Southern
France
Between regions,
all times
Between
regions, UP
Between
regions, EP
All regions,
through time
Mediterranean
Iberia
Euro-Siberian
Iberia
Southern
France
Between
regions
Between
regions, UP
Between
regions, EP

ANOVA
F = 0.10,
p = 0.96
F = 1.47,
p = 0.24
F = 0.83,
p = 0.49
F = 2.41,
p = 0.08
F = 10.29,
p = 0.00
F = 5.49,
p = 0.01
F = 7.57,
p = 0.00
F = 1.64,
p = 0.18
F = 1.45,
p = 0.24
F = 0.79,
p = 0.51
F = 2.59;
p = 0.06
F = 1.96,
p = 0.14
F = 3.10,
p = 0.06
F = 0.00,
p = 0.99

Welchs test
F = 0.09,
p = 0.96
F = 1.32;
p = 0.30
F = 3.14;
p = 0.07
F = 4.17;
p = 0.02
F = 14.75,
p = 0.00
F = 5.83,
p = 0.01
F = 17.60,
p = 0.00
F = 1.61,
p = 0.19
F = 1.09;
p = 0.38
F = 2.88;
p = 0.09
F = 5.01;
p = 0.00
F = 1.83,
p = 0.16
F = 2.44,
p = 0.10
F = 0.0,
p = 0.99

Kruskal
Wallis
H = 0.25,
p = 0.97
H = 5.85,
p = 0.11
H = 4.87,
p = 0.18
H = 5.51,
p = 0.14
H = 22.69,
p = 0.00
H = 9.70,
p = 0.01
H = 14.16,
p = 0.00
H = 7.18,
p = 0.07
H = 3.80,
p = 0.28
H = 5.29,
p = 0.15
H = 7.75,
p = 0.05
H = 4.58,
p = 0.10
H = 5.17,
p = 0.08
H = 0.10,
p = 0.95

Bartletts
2 = 0.47,
p = 0.92
2 = 1.45,
p = 0.69
2 = 14.83,
p = 0.00
2 = 3.04,
p = 0.39
2 = 16.58,
p = 0.00
2 = 12.53,
p = 0.00
2 = 19.44,
p = 0.00
2 = 0.05,
p = 0.99
2 = 5.86,
p = 0.12
2 = 12.42,
p = 0.00
2 = 6.58,
p = 0.09
2 = 1.03,
p = 0.60
2 = 9.14,
p = 0.01
2 = 9.28,
p = 0.01

consistently present at high frequencies in Mediterranean Iberia and rare in


Euro-Siberian Iberia. In southern France, they are infrequent in the Upper Paleolithic
and increase significantly in the Epipaleolithic (Fig. 5.5). As discussed in Sect. 5.3.2,
the differential representation of lagomorphs does seem to be driving variation in
evenness between regions, but their increasing representation in southern France
through time does not impact evenness values (Table 5.4). In addition, the increasing representation of lagomorphs seems to be entirely due to an increase in
Oryctolagus in the Epipaleolithic; there is no statistically significant change in the
representation of Lepus in southern French archaeofaunas either between the Upper
Paleolithic and Epipaleolithic (the non-parametric KolmogorovSmirnov test for
equal distributions: D = 0.18, p = 0.74) or from the Gravettian through the
Epipaleolithic (KruskalWallis: H = 1.60, p = 0.65).

72

5 Archaeofaunal Diversity and Broad Spectrum Diets in Late Paleolithic Southwest

While the representation of lagomorphs increases significantly in southern France


in the Epipaleolithic, we cannot conclude that this is an indicator of increasing
resource stress. Not only are there known problems with assuming lagomorphs are
lower-ranked resources, but in these data the increase in lagomorphs seems to be
driven by the range expansion of Oryctolagus.

5.2.4

Diversity, Diet Breadth, and Faunal Turnover

The diversity measures all tell the same story: there were changes in available faunas in Southwest Europe in the Epipaleolithic, but these do not seem to have led to
an expansion in diet breadth. In Mediterranean Iberia, the archaeofaunas suggest
continuity. In Euro-Siberian Iberia, the loss of the reindeer seems to have resulted in
a slight restriction in mammalian diet breadth, but in all other waysas in
Mediterranean Iberiadiets seem to have remained relatively constant. In southern
France, there are significant changes in the composition of archaeofaunascoldadapted taxa decrease, warmer-adapted taxa increase, and Oryctolagus drives
increasing representation of lagomorphsbut these changes offset each other and
so are not apparent in either richness or evenness values. These analyses suggest
that changes in southern French archaeofaunas in the late Paleolithic reflect climate
change-driven faunal turnoverthe replacement of one set of faunas by another
rather than diet breadth expansion. Fortunately, there is a clear-cut method with
which to test this hypothesis: nestedness.

5.3

Nestedness and Faunal Turnover

Nestedness, or the degree to which assemblages are subsets of each other, is a


widely used measure of biotic structure (Atmar and Patterson 1993; Bascompte
et al. 2003; Dapporto et al. 2014; Husemann et al. 2014). A nested assemblage is
one in which all the taxa in less-rich assemblages are present in the most-rich
assemblages. To use a hypothetical example, a dataset in which some assemblages
contained reindeer, red deer, boar, and hare, while others contained red deer, boar,
and hare, and yet others contained only hare, would be nested (Table 5.4). Nestedness
today is commonly seen in islands (biotas on smaller islands are often subsets of
those found in larger islands) and in mutualistic networks (Bascompte et al. 2003;
Gilarranz et al. 2015).
An unnested assemblage, by contrast, is one which contains completely different
taxa. For instance, a dataset in which some assemblages contained reindeer and red
deer, and others boar and hare, would be unnested (Table 5.4). Unnested biotas today
are present in situations where there is faunal turnoveralong latitudinal gradients,
for instance (Dobrovolski et al. 2012; Mouillot et al. 2013; Svenning et al. 2011).

5.3

Nestedness and Faunal Turnover

5.3.1

73

Understanding Zooarchaeological Nestedness

Nestedness has been used in a variety of ways in zooarchaeological analyses


(Conrad 2015; Jones 2004a, 2015; Lyman 2008), but all are based on the understanding that in most cases, archaeofaunas should be nested: some will contain a
full set of species, and others, subsets. In many cases, this will reflect the sample
size-NTAXA relationship. If sample size is not driving nestedness, then nestedness
may indicate overall shifts in diet breadth (Jones 2004a). By contrast, unnested
archaeofaunas suggest environmentally-driven faunal turnover, or in some cases,
the presence of many distinct, specialized hunting strategies (Jones 2015).
If there was a climate-driven transformation in the faunal landscapes of southern
France in the Epipaleolithic, and this transformation was the driver of differences in
archaeofaunal composition identified in this chapter and in Chap. 4, we would
expect the archaeofaunas from this time and region to be unnested. All other
archaeofaunas in the dataset (when analyzed by region) should be nested.

5.3.2

Calculating Nestedness: The NODF

Like diversity, nestedness can be measured in a variety of ways (e.g., Almeida-Neto


et al. 2008; Guimares and Guimares 2006; Patterson and Atmar 1995). For this
analysis, I use a metric called the NODF (nestedness based on overlap and decreasing fills; Almeida-Neto and Ulrich 2011), which I calculated using NeD software
(Strona et al. 2014). The NODF produces values ranging from 0 to 100; these values
increase as nestedness increases (a perfectly nested set would score 100). I also used
NeD to calcluate standardized effect sizes (as z-values) based on 500 simulated null
matrices (using the proportional row and column total algorithm; see Bascompte
et al. 2003). Effect sizes with a p value lower than .05 were considered significantly
nested; those greater than .05 were classified as unnested.

5.3.3

Regional Nestedness in Southwest Europe

The archaeofaunas from Epipaleolithic southern France are not nested (Table 5.5),
supporting the hypothesis that the changes in composition observed in this region
reflect climate change-driven faunal turnover rather than resource stress-induced
Table 5.5 Hypothetical examples of nested and unnested archaeofaunas
Taxon
Reindeer
Red deer
Wild boar
Hare

Nested
Assemblage 1
Present
Present
Present
Present

Assemblage 2
Absent
Absent
Present
Present

Unnested
Assemblage 3
Present
Present
Absent
Absent

Assemblage 4
Absent
Absent
Present
Present

74

5 Archaeofaunal Diversity and Broad Spectrum Diets in Late Paleolithic Southwest

Table 5.6 Archaeofaunal nestedness by region and through time


Region
Mediterranean
Iberia
EuroSiberian Iberia
Southern France

Time period
Upper Paleolithic
Epipaleolithic
Upper Paleolithic
Epipaleolithic
Upper Paleolithic
Epipaleolithic

Matrix
size
28 10
15 10
26 12
10 9
45 12
18 12

Fill
64.3 %
60.7 %
68.3 %
71.1 %
63.0 %
64.8 %

NODF
77.21, z = 3.04, p < 0.01
73.16, z = 1.89, p < 0.05
68.50, z = -0.54, p > 0.05
86.17, z = 2.65, p < 0.01
77.81, z = 4.65, p < 0.001
73.10, z = 1.53, p > 0.05

Nested?
Yes
Yes
No
Yes
Yes
No

increasing diet breadth. However, while the majority of the other archaeofaunas are,
as predicted, nested, there is one exception: the data from Upper Paleolithic EuroSiberian Iberia are also unnested.
Does this mean that Euro-Siberian Iberia underwent faunal turnover in the Upper
Paleolithic? In this case, probably not. As mentioned earlier, unnested archaeofaunas can indicate faunal turnover, but sometimes they may reflect the presence of
many distinct, specialized hunting strategies (as in Jones 2015). Because the huntergatherers of Upper Paleolithic Euro-Siberian Iberia seem to have practiced a more
residential mobility pattern (Chap. 4), it is likely that in this case, these faunas are
unnested because of the diversity of site types from which these archaeofaunas originatean interpretation supported by other research as well (Jones 2013a, 2015). It
is interesting to note that Euro-Siberian Iberian Epipaleolithic archaeofaunas are
nested. It is possible that in the Epipaleolithic, mobility began to become more
logistical in this area (Jones 2013a), but more data from the early Holocene period
would be necessary to test this hypothesis (Table 5.6).

5.4

Landscape Transformations and Resource Stress

The mammalian diversity and nestedness data from Late Paleolithic Southwest
Europe tell a coherent tale: there is no evidence for increasing diet breadth here.
Instead, these data suggest a strongly regional landscape in which barriers and corridors shaped prehistoric peoples access to different resources. In the refugia of
Mediterranean Iberia, there is no evidence for any significant changes in archaeofaunal composition from the Gravettian through the later Upper Paleolithic. The
archaeofaunas from Euro-Siberian Iberia do provide some evidence for change,
with a significant reduction in NTAXA and an increase in nestedness in the
Epipaleolithic. Although these assemblages also suggest continuity, as evenness
and relative abundance of major prey types remain constant, the results presented
here are not incompatible with the hypothesis that diet breadth expanded in this
region in the early Holocene (as suggested by Marn Arroyo 2013).
Change does seem to have taken place in southern France. While evenness and
richness values remain constant across the Pleistocene-Holocene transition in this

References

75

region, compositional analysis reveals this is an artifact of climate-driven faunal


turnover. There are significant differences in the taxonomic structure of Upper
Paleolithic and Epipaleolithic archaeofaunas from southern France, revealed by
shifts in relative abundance of different prey types and by a change from nested to
nonnested archaeofaunas as well as by the site location and archaeofaunal similarity
data in Chap. 4. However, these differences seem to reflect climate-driven changes
in environment rather than resource stress.

References
Almeida-Neto, M., Guimares, P., Guimares, P. R., Loyola, R. D., & Ulrich, W. (2008). A consistent metric for nestedness analysis in ecological systems: Reconciling concept and measurement. Oikos, 117, 12271239. doi:10.1111/j.0030-1299.2008.16644.x.
Almeida-Neto, M., & Ulrich, W. (2011). A straightforward computational approach for measuring
nestedness using quantitative matrices. Environmental Modelling & Software, 26, 173178.
doi:http://dx.doi.org/10.1016/j.envsoft.2010.08.003.
Altuna, J. (1972). Fauna de mamferos de los yacimientos prehistricos de Guipzcoa. Munibe, 24,
1465.
lvarez-Lao, D. J., & Garca, N. (2010). Chronological distribution of Pleistocene cold-adapted
large mammal faunas in the Iberian Peninsula. Quaternary International, 212, 120128.
Atmar, W., & Patterson, B. (1993). The measure of order and disorder in the distribution of species
in fragmented habitat. Oecologia, 96, 373382. doi:10.1007/bf00317508.
Bascompte, J., Jordano, P., Melin, C. J., & Olesen, J. M. (2003). The nested assembly of plant
animal mutualistic networks. Proceedings of the National Academy of Sciences, 100, 9383
9387. doi:10.1073/pnas.1633576100.
Bilsborough, S., & Mann, N. (2006). A review of issues of dietary protein intake in humans.
International Journal of Sport Nutrition & Exercise Metabolism, 16, 129152.
Callou, C. (2003). De la garenne au clapier. Histoire et archologie du lapin europen. Paris:
Publications scientifiques du Musum.
Cannon, M. D. (1999). A mathematical model of the effects of screen size on zooarchaeological
relative abundance measures. Journal of Archaeological Science, 26, 205214.
Cochard, D. (2004). Les lporids dans la subsistance palolithique du Sud de la France. Talence:
Universit de Bordeaux I. Thse de doctorat.
Conrad, C. (2015). Archaeozoology in Mainland Southeast Asia: Changing Methodology and
Pleistocene to Holocene Forager Subsistence Patterns in Thailand and Peninsular Malaysia.
Open Quaternary, 1, 123. doi:http://dx.doi.org/10.5334/oq.af.
Corts-Snchez, M., Morales-Muiz, A., Simn-Vallejo, M. D., Bergad-Zapata, M. M., DelgadoHuertas, A., Lpez-Garca, P., et al. (2008). Palaeoenvironmental and cultural dynamics of the
coast of Mlaga (Andalusia, Spain) during the Upper Pleistocene and early Holocene. Quaternary
Science Reviews, 27, 21762193. doi:http://dx.doi.org/10.1016/j.quascirev.2008.03.010.
Dapporto, L., Fattorini, S., Vod, R., Dinc, V., & Vila, R. (2014). Biogeography of western
Mediterranean butterflies: combining turnover and nestedness components of faunal dissimilarity. Journal of Biogeography, 41, 16391650. doi:10.1111/jbi.12315.
Dean, R. M. (2007a). Hunting intensification and the Hohokam collapse. Journal of
Anthropological Archaeology, 26, 109132. doi:10.1016/j.jaa.2006.03.010.
Dean, R. M. (2007b). The lagomorph index: Rethinking rabbit bones in Hohokam sites. Kiva, 73, 730.
Delpech, F. (1983). La faune du Palolithique suprieur dans le Sud-Ouest de la France. Paris:
Cahier du Quaternaire.
Delpech, F. (1999). Biomasse donguls au Palolithique et infrences sur la dmographie. Palo,
11, 1942.

76

5 Archaeofaunal Diversity and Broad Spectrum Diets in Late Paleolithic Southwest

Dobrovolski, R., Melo, A. S., Cassemiro, F. A. S., & Diniz-Filho, J. A. F. (2012). Climatic history and
dispersal ability explain the relative importance of turnover and nestedness components of beta
diversity. Global Ecology and Biogeography, 21, 191197. doi:10.1111/j.1466-8238.2011.00671.x.
Driver, J. C., & Woiderski, J. R. (2008). Interpretation of the lagomorph index in the American
Southwest. Quaternary International, 185, 311. doi:http://dx.doi.org/10.1016/j.quaint.2007.09.022.
Flannery, K. V. (1969). Origins and ecological effects of early domestication in Iran and the Near
East. In P. J. Ucko & G. W. Dimbleby (Eds.), The domestication and exploitation of plants and
animals (pp. 73100). Chicago: Aldine.
Gilarranz, L. J., Sabatino, M., Aizen, M. A., & Bascompte, J. (2015). Hot spots of mutualistic
networks. Journal of Animal Ecology, 84, 407413. doi:10.1111/1365-2656.12304.
Gmez-Olivencia, A., Arceredillo, D., lvarez-Lao, D. J., Garate, D., San Pedro, Z., Castaos, P.,
et al. (2014). New evidence for the presence of reindeer (Rangifer tarandus) on the Iberian
Peninsula in the Pleistocene: An archaeopalaeontological and chronological reassessment.
Boreas, 43, 286308. doi:10.1111/bor.12037.
Grayson, D. K. (1984). Quantitative zooarchaeology. New York: Academic Press.
Grayson, D. K., & Delpech, F. (1998). Changing diet breadth in the Early Upper Paleolithic of
southwestern France. Journal of Archaeological Science, 25, 11191130.
Grayson, D. K., & Delpech, F. (2002). Specialized early Upper Paleolithic hunters in Southwestern
France? Journal of Archaeological Science, 29, 14391449.
Grayson, D. K., Delpech, F., Rigaud, J.-P., & Simek, J. F. (2001). Explaining the development of
dietary dominance by a single ungulate taxon at Grotte XVI, Dordogne, France. Journal of
Archaeological Science, 28, 115125.
Guimares, P., & Guimares, P. R. (2006). Improving the analyses of nestedness for large sets of
matrices. Environmental Modelling and Software, 21, 15121513.
Hockett, B. S., & Bicho, N. F. (2000). The rabbits of Picareiro Cave: small mammal hunting during
the Late Upper Palaeolithic in the Portuguese Estremadura. Journal of Archaeological Science,
27, 715723.
Hockett, B. S., & Haws, J. A. (2002). Taphonomic and methodological perspectives of leporid
hunting during the Upper Paleolithicof the Western Mediterranean Basin. Journal of
Archaeological Method and Theory, 9, 269302.
Hockett, B. S., & Haws, J. A. (2003). Nutritional ecology and diachronic trends in Paleolithic diet
and health. Evolutionary Anthropology, 12, 211216.
Husemann, M., Schmitt, T., Zachos, F. E., Ulrich, W., & Habel, J. C. (2014). Palaearctic biogeography revisited: Evidence for the existence of a North African refugium for Western Palaearctic
biota. Journal of Biogeography, 41, 8194. doi:10.1111/jbi.12180.
Jones, E. L. (2004a). Dietary evenness, prey choice, and human-environment interactions. Journal
of Archaeological Science, 31, 307317.
Jones, E. L. (2004b). The European Rabbit (Oryctolagus cuniculus) and the development of broad
spectrum diets in southwestern France: Data from the Dordogne Valley. In J.-P. Brugal &
J. Desse (Eds.), Petits animaux et socits humaines: du complment alimentaire aux ressources utilitaires (pp. 223234). Antibes: ditions APDCA.
Jones, E. L. (2006). Prey choice, mass collecting, and the wild European rabbit (Oryctolagus
cuniculus). Journal of Anthropological Archaeology, 25, 275289.
Jones, E. L. (2012). Upper Paleolithic rabbit exploitation and landscape patchiness: The Dordogne
vs. MediterraneanSpain. Quaternary International, 264, 5260.
Jones, E. L. (2013a). Mobility, settlement, and resource patchiness in Upper Paleolithic Iberia.
Quaternary International, 318, 4652. doi:http://dx.doi.org/10.1016/j.quaint.2013.05.027.
Jones, E. L. (2013b). Subsistence change among the 17th-century Din? A reanalysis of the faunas
from the Fruitland Data Recovery Project. Journal of Ethnobiology, 33, 148166.
doi:10.2993/0278-0771-33.1.148.
Jones, E. L. (2015). Archaeofaunal evidence of human adaptation to climate change in Upper
Paleolithic Iberia. Journal of Archaeological Science: Reports, 2, 257263. doi:http://dx.doi.
org/10.1016/j.jasrep.2015.02.008.

References

77

Jones, E. L. (2016). Changing landscapes of early colonial New Mexico: Demography, rebound,
and zooarchaeology. In C. Herhahn, & A. F. Ramenofsky (Eds.), How, why, and beyond:
exploring cause and explanation in historical ecology, demography, and movement. Boulder,
CO: University of Colorado Press.
Jones, E. L., & Gabe, C. (2015). The promise and peril of older collections: meta-analyses in the
American Southwest. Open Quaternary, 1, 113. doi: http://doi.org/10.5334/oq.ag.
Lopez-Martinez, N. (2008). The lagomorph fossil record and the origin of the European rabbit. In
P. C. Alves, N. Ferrand, & K. Hacklnder (Eds.), Lagomorph biology (pp. 2746). Berlin,
Germany: Springer.
Lupo, K. D. (2007). Evolutionary foraging models in zooarchaeological analysis: recent applications and future challenges. Journal of Archaeological Research, 15, 143189. doi:10.1007/
s10814-007-9011-1.
Lyman, R. L. (2008). Quantitative paleozoology. Cambridge, MA: Cambridge University Press.
Lyman, R. L. (2015). On the variable relationship between NISP and NTAXA in bird remains and
in mammal remains. Journal of Archaeological Science, 53, 291296. doi: http://dx.doi.
org/10.1016/j.jas.2014.10.027.
Magurran, A. E. (2004). Measuring biological diversity. Malden, MA: Blackwell.
Magurran, A. E., & McGill, B. J. (2011). Biological diversity: frontiers in measurement and
assessment. New York: Oxford University Press.
Marn Arroyo, A. B. (2013). Human response to Holocene warming on the Cantabrian Coast
(northern Spain): an unexpected outcome. Quaternary Science Reviews, 81, 111. doi:http://
dx.doi.org/10.1016/j.quascirev.2013.09.006.
Morrison, A. E., & Hunt, T. L. (2007). Human Impacts on the Nearshore Environment: An archaeological case study from Kauai, Hawaiian Islands. University of Hawaii Press Pacific Science,
61, 325345.
Mouillot, D., De Bortoli, J., Leprieur, F., Parravicini, V., Kulbicki, M., & Bellwood, D. R. (2013).
The challenge of delineating biogeographical regions: nestedness matters for Indo-Pacific coral
reef fishes. Journal of Biogeography, 40, 22282237. doi:10.1111/jbi.12194.
Munro, N. D. (2004). Zooarchaeological measures of hunting pressure and occupation intensity in
the Natufian. Current Anthropology, 45, S5S33.
Nagaoka, L. (2001). Using diversity indices to measure changes in prey choice at the Shag River
Mouth Site, Southern New Zealand. International Journal of Osteoarchaeology, 11,
101111.
Nagaoka, L. (2002). The effects of resource depression on foraging efficiency, diet breadth and
patch use in southern New Zealand. Journal of Anthropological Archaeology, 21, 419442.
Nagaoka, L. (2005). Differential recovery of Pacific Island fish remains. Journal of Archaeological
Science, 32, 941955. doi:10.1016/j.jas.2004.12.011.
Nelson, M. C., & Schollmeyer, K. G. (2003). Game resources, social interaction, and the ecological footprint in southwest New Mexico. Journal of Archaeological Method and Theory, 10,
69110.
Patterson, B. D., & Atmar, W. (1995). The nestedness calculator: a visual basic program, including 294 presence-absence matrices. Chicago: AICS Research and the Field Museum.
Peres, T. M. (2010). Methodological Issues in Zooarchaeology. In A. M. VanDerwarker & T. M.
Peres (Eds.), Integrating zooarchaeology and paleoethnobotany: A consideration of issues,
methods, and cases (pp. 1536). New York: Springer.
Pokines, J. T. (2000). Microfaunal research design in the Cantabrian Spanish Paleolithic. Journal
of Anthropological Research, 56, 95112.
Rillardon, M., & Brugal, J.-P. (2014). What about the Broad Spectrum RevolutionSubsistence
strategy of huntergatherers in Southeast France between 20 and 8 ka BP. Quaternary
International, 337, 129153. doi:http://dx.doi.org/10.1016/j.quaint.2014.01.020.
Schaffer, B. S., & Sanchez, J. L. J. (1994). Comparison of 1/8- and 1/4-mesh recovery of
controlled samples of small-to-medium-sized mammals. American Antiquity, 59, 525530.

78

5 Archaeofaunal Diversity and Broad Spectrum Diets in Late Paleolithic Southwest

Schollmeyer, K. G. (2011). Large game, agricultural land, and settlement pattern change in the
eastern Mimbres area, southwest New Mexico. Journal of Anthropological Archaeology, 30,
402415. doi:10.1016/j.jaa.2011.04.004.
Sommer, R. S., Kalbe, J., Ekstrm, J., Benecke, N., & Liljegren, R. (2013). Range dynamics of the
reindeer in Europe during the last 25,000 years. Journal of Biogeography, 41, 298306.
doi:10.1111/jbi.12193.
Sommer, R. S., Zachos, F. E., Street, M., Jris, O., Skog, A., & Benecke, N. (2008). Late Quaternary
distribution dynamics and phylogeography of the red deer (Cervus elaphus) in Europe.
Quaternary Science Reviews, 27, 714733.
Speth, J. D., & Spielmann, K. A. (1983). Energy source, protein metabolism, and hunter-gatherer
subsistence strategies. Journal of Anthropological Archaeology, 2, 131.
Stahl, P. W. (1996). The recovery and interpretation of microvertebrate bone assemblages from
archaeological contexts. Journal of Archaeological Method and Theory, 3, 3175.
Stiner, M. C., & Munro, N. D. (2002). Approaches to prehistoric diet breadth, demography, and
prey ranking systems in time and space. Journal of Archaeological Method and Theory, 9,
181214.
Strona, G., Nappo, D., Boccacci, F., Fattorini, S., & San-Miguel-Ayanz, J. (2014). A fast and unbiased procedure to randomize ecological binary matrices with fixed row and column totals.
Nature Communications, 5, 4414. doi:10.1038/ncomms5114.
Stutz, A. J., Munro, N. D., & Bar-Oz, G. (2009). Increasing the resolution of the Broad Spectrum
Revolution in the Southern Levantine Epipaleolithic (1912 ka). Journal of Human Evolution,
56, 294306. doi:http://dx.doi.org/10.1016/j.jhevol.2008.10.004.
Svenning, J.-C., Fljgaard, C., & Baselga, A. (2011). Climate, history and neutrality as drivers of
mammal beta diversity in Europe: insights from multiscale deconstruction. Journal of Animal
Ecology, 80, 393402. doi:10.1111/j.1365-2656.2010.01771.x.
Ugan, A. (2005). Climate, bone density, and resource depression: What is driving variation in large
and small game in Fremont archaeofaunas? Journal of Anthropological Archaeology, 24,
227251. doi:10.1016/j.jaa.2005.05.002.
Villaverde, V., Aura, J. E., & Barton, C. M. (1998). The Upper Paleolithic in Mediterranean Spain:
A review of current evidence. Journal of World Prehistory, 12, 121198.
Wolverton, S., Dombrosky, J., & Lyman, R. L. (2014). Practical Significance: Ordinal Scale Data
and Effect Size in Zooarchaeology. International Journal of Osteoarchaeology. doi:10.1002/
oa.2416.
Zar, J. H. (1999). Biostatistical analysis. Upper Saddle River, NJ: Prentice Hall.
Zhang, Y., Zhang, S., Xu, X., Liu, D., Wang, C., Pei, S., et al. (2013). Zooarchaeological perspective on the Broad Spectrum Revolution in the Pleistocene-Holocene transitional period, with
evidence from Shuidonggou Locality 12, China. Science China Earth Sciences, 56, 14871492.
doi:10.1007/s11430-013-4584-7.

Chapter 6

Was There a Broad Spectrum Revolution


in Southwest Europe?

6.1

Introduction

As of this writing, Kent Flannerys Broad Spectrum Revolution hypothesis is nearly


50 years old (1969), but research inspired by this model shows no sign of slowing
(e.g., Conrad 2015; Rillardon and Brugal 2014; Stiner and Munro 2011; Zeder
2012). This has, without a doubt, been an extremely productive concept for archeology. Two general findings have emerged in recent years. First, while there is evidence for increasing diet breadth in many locations, it does not always correspond
with the PleistoceneHolocene transition. Sometimes diet breadth increases well
before the 10 kya mark (Manne et al. 2012; Stiner and Munro 2011; Straus and
Clark 1986); sometimes, the change comes later (Conrad 2015). Second, in many
cases broad spectrum diets seem to have emerged through a quite complex set of
humanenvironment interactions (e.g., this volume; Jones 2006; Rillardon and
Brugal 2014; Zaatari and Hublin 2014; Zeder 2012). This second seems to have
been the case for Southwest Europe.

6.2

What the Archeofaunas Say

In this volume, I used analyses of prehistoric human subsistenceincluding site


location (Sect. 4.2), archeofaunal similarity (cluster and NMMDS analyses; Sect.
4.3), archeofaunal diversity (Sect. 5.2), and archeofaunal nestedness (Sect. 5.3)to
test for evidence of changing diets during the period immediately preceding the
Pleistocene/Holocene transition. While these tests did produce some evidence for
change, the results varied by region (Table 6.1).

The Author(s) 2016


E.L. Jones, In Search of the Broad Spectrum Revolution in Paleolithic Southwest
Europe, SpringerBriefs in Archaeology, DOI 10.1007/978-3-319-22351-3_6

79

80

6 Was There a Broad Spectrum Revolution in Southwest Europe?

Table 6.1 Summary of subsistence-related changes in Late Paleolithic Southwest Europe


Region
Mediterranean
Iberia

Time period
Upper Paleolithic
Epipaleolithic

Euro-Siberian
Iberia

Upper Paleolithic
Epipaleolithic

Southern France

Upper Paleolithic

Epipaleolithic

6.2.1

Effect
Some local variation in subsistence but overall
consistency. Logistical mobility
No significant change in faunal composition from the
Upper Paleolithic; a possible increase in residential
mobility?
Overall consistency in faunal composition; evidence
of residential mobility and targeted hunting
Extirpation of reindeer causes decrease in average
NTAXA; mobility may become less residential
Logistical mobility. Overall consistency but much
variation in faunal composition, suggesting an
unpredictable, patchy, non-analogous environment
across the region
Faunal composition inconsistent, suggesting the
presence of numerous distinct environments within the
region. Significant increase in warmer-adapted taxa and
concurrent decrease in cold-adapted taxa; the wild
European rabbit enters the diet on a consistent basis in
some areas. Site elevation variance increases suggesting
increasing diversity in hunter-gatherer mobility

Mediterranean Iberia

The Mediterranean Iberian region produced the least evidence for dietary change in
the Late Paleolithic, perhaps not surprising given that a this environment was likely
comparatively stable in the Pleistocene (Chap. 3). Taxonomic richness (Sect. 5.2.1)
and evenness (Sect. 5.2.2) remain constant and analyses found no evidence for significant change in taxonomic representation (Sects. 4.3 and 5.2). This is not to say
that subsistence and hunting practice did not change at all for the Late Paleolithic
huntergatherers of Mediterranean Iberiasite-specific evidence (e.g., CortsSnchez et al. 2008; Manne et al. 2012) demonstrates this to be false, and the shifting distribution of archeological sites in this region in response to the LGM (Jones
2013; Straus et al. 2000) likewise suggests otherwise.
However, there does not seem to have been a regionwide directional change in
Mediterranean Iberia, as would be expected in the case of a Broad Spectrum
Revolution. Rather, huntergatherers seem to have practiced a consistent logistical
foraging strategy throughout the Late Paleolithic. Increasing site elevation variance
during the Epipaleolithic (Sect. 4.2) may suggest huntergatherer mobility became
more residential during this period; at the very least, sites are located across a wider
elevation range in this period. But otherwise, the record from Mediterranean Iberia
suggests that on average, at least the mammalian portion of human diets remained
consistent into the Epipaleolithic in this region.

6.2

What the Archeofaunas Say

6.2.2

81

Euro-Siberian Iberia

Some evidence of dietary change is present in archeofaunas from the Euro-Siberian


regions of Iberia, although again, the data suggest more continuity than otherwise
(see also Straus 2009). The effects of Blling/Allerd warming and the concomitant
extirpation of reindeer are apparent in declining NTAXA values in the Epipalolithic
(Sect. 5.2.1), and the site elevation (Sect. 4.2) and nestedness analyses (Sect. 5.3.3)
suggest a decline in residential mobility in this period.
However, the similarity (Sect. 4.3) and evenness (Sect. 5.2.2) analyses show consistency in subsistence adaptations across the Late Paleolithic. While climate change
certainly impacted the resources available in this region, no revolution is evident
in the mammalian taxamerely a replacement of cold-adapted taxa with warmeradapted ones. Lagomorphs (here primarily hares, Lepus spp.) remain a constant but
low-frequency presence in both Upper Paleolithic and Epipaleolithic archeofaunas.
These data, of course, do not preclude potential widening of the diet among nonmammalian taxa. Shellfish are present in at least some Euro-Siberian Iberian sites
as early as the Solutrean (Straus and Clark 1986) and their use may have intensified
in the early Holocene (i.e., Marn Arroyo 2013; although see Gutirrez-Zugasti
2011). Nor do these analyses preclude earlier episodes of intensification (as in
Freeman 1973, Straus 1977, Straus 2009) However, the mammalian faunas from the
Pleistocene-Holocene transition provide more evidence of consistency than change.

6.2.3

Southern France

The majority of the evidence for change in these data comes from southern France.
While portions of this regionparticularly the Dordogne River Valleydo seem to
have served as a refuge for some warmer-adapted taxa during all but the coldest periods (e.g., Sommer et al. 2011; Sommer and Nadachowski 2006; Sommer and Zachos
2009), the refugia of southern France were tightly constrained by a number of physiogeographic barriers, including the Landes periglacial desert, the Pyrenees, and the
Massif Central. As a whole, southern France may have been relatively resource-poor,
at least in some places and some of the time. Low site elevation variance (Sect. 4.2),
archeofaunas that group together and contain numerous different clusters (Sect. 4.3)
but nonetheless are highly nested (Sect. 5.3) suggest Upper Paleolithic environments
that were (albeit consistently) both patchy and unpredictable.
Epipaleolithic archeofaunas, by contrast, contain strong evidence for change (Sects.
4.3 and 5.3.3), although not in the manner predicted for a Broad Spectrum Revolution.
These data provide no evidence of widening diet breadth. The proportion of lagomorphs in archeofaunas undoubtedly increases at this time, but this is due to the environmentally driven rise in abundance of the wild European rabbit; there is no significant
change in hare representation (Sect. 5.2.3). In addition, changing lagomorph abundance does not impact diversity values; both NTAXA and evenness stay constant
across the Upper PaleolithicEpipaleolithic boundary (Sects. 5.2.1 and 5.2.2).

82

6 Was There a Broad Spectrum Revolution in Southwest Europe?

The consistency in diversity seems to reflect the simultaneous extirpation of


cold-adapted species and colonization/increasing abundance of warmer-adapted
taxa. Epipaleolithic faunas are unnested, indicating significant turnover in the mammalian taxa (Sect. 5.3.3) and similarity data show a collapse of the archeofaunal
patterns seen in the Upper Paleolithic (Sect. 4.3). Whereas in the Upper Paleolithic
Southern French archeofaunas showed consistency, in the Epipaleolithic the faunas
suggest a move towards the bioregions present in this area today. The French
Mediterranean archeofaunas from this period group with the Iberian Mediterranean
ones; the higher-elevation French sites group with higher-elevation sites from EuroSiberian Iberia; and the faunas from the Dordogne show heterogeneity, just as the
flora and fauna in this river valley do today.
How did people respond to this environmental change? Site location analyses
suggest a significant change in mobility in the Epipaleolithic of southern France.
Site elevation variance increases, indicating more variability in site location than
was present in the Upper Paleolithic (Sect. 4.2). And finally, the changes in both the
archeofaunas and in human settlement seem to have been nonlinear (in contrast
with the predictions of the Broad Spectrum Revolution) and spatially variable (in
agreement with what we know about environmental change). This suggests environment, not human population pressure, is the cause of the dietary transformations
associated with the PleistoceneHolocene transition.

6.3

Environmental Change and Resource Stress

The subsistence changes in Euro-Siberian Iberia and southern France seem to have
been driven primarily by environmental change rather than by independent human
demographic pressure. In this sense, these analyses do not support the presence of a
Broad Spectrum Revolution in Southwest Europe at the end of the Paleolithic.
However, population pressure is not so easy to separate from environmental
change; often, the two work together. When there is a significant change in the environmentbe it physiogeographic, floral, or faunalsuch a change can decrease (or
increase) the number of people that environment is able to support (e.g., Binford
1968). Binford specifically highlighted the loss of land to sea-level rise, but any
environmental change may impact the resources available to people (as well as to
other taxa)in either a positive or a negative sense.
Therefore the fact that in Late Paleolithic Southwest Europe environmental change
played a large role in the shifting abundance of faunas in the human diet does not
negate the possibility that environmentally driven resource stress caused a broadening
of diets. There is strong evidence in many Southwest Europe archeofaunas for environmentally-driven stress in response to climatically challenging periods (Aura et al. 2015;
Jones 2009; Manne et al. 2012; Straus and Clark 1986), and the data presented here do
indicate significant change in southern France and some change in Euro-Siberian
Iberia. Was warming a challenge for Late Paleolithic huntergatherers?
While this question remains open, the faunal turnover and increasing variance in
site location in southern France in fact suggests the oppositethat there were more

References

83

resources available to huntergatherers in southern France in the Epipaleolithic.


The archeofaunas from this region indicate constrained, patchy, and unpredictable
environments in the Upper Paleolithica landscape constrained by barriers and
with fluctuating environmental conditions. This likely would have been extremely
challenging for huntergatherers (Jones 2007). The low site elevation variance,
combined with the physical barriers discussed in Chap. 3, suggests a highly constrained environment in which people had few options to move.
This seems to have changed in the Epipaleolithicfor the better. Variabilityin site
location as well as in archeofaunal compositionincreased, and this may suggest people had more options when climates warmed. The addition of the wild European rabbit
to the diet may be one symptom of better conditions in the Epipaleolithic (Jones 2007).
The data from Euro-Siberian Iberia are more equivocal. Though diet breadth
decreased, rather than increased, in this region across the Upper Paleolithic boundary,
this likely reflects the focus of these analyses on mammalian taxa. Decreasing site
elevation variance and an increase in nestedness in the Epipaleolithic could reflect the
beginnings of resource stress here; but such change, if it indeed occurred, likely took
place primarily in the early Holocene, a period outside the scope of this work.

6.4

Conclusions

The Late Paleolithic hunter-gatherers of Southwest Europe do not appear to have been
suffering from resource stress due to environmental change. While stress may have
come in the early Holocene, warming prior to this period seems (contrary to widelyheld assumptions) to have no impact on the hunting of large mammals in Mediterranean
Iberia, to have had limited impact in Euro-Siberian Iberia, and to have lessened resource
stress in southern France. In short, these data show no evidence for either the classically
conceived Broad Spectrum Revolution or for negative impact of warming climate.
This does not mean, of course, that climate change was unimportant in the Late
Paleolithic. On the contrary, the environmental changes indicated by these data
clearly had profound impacts on the hunter-gatherers who lived through them. But
the archaeological data I have considered here suggest these hunter-gatherers
responded to these changes regionally, developing a mosaic of adaptations to
climate variability rather than following a unilinear trajectory. The landscapes of
Late Paleolithic Southwest Europe, though substantially different from modern
ones, seem to have been similar to modern ones in their diversity.

References
Aura, J. E., Jord, J. F., Montes, L., & Utrilla, P. (2015). Human responses to Younger Dryas in the Ebro
Valley and Mediterranean Watershed (Eastern Spain). Quaternary International, 242, 348359.
Binford, L. R. (1968). Post-Pleistocene adaptations. In S. R. Binford & L. R. Binford (Eds.), New
Perspectives in Archaeology (pp. 313341). Chicago: Aldine.
Conrad, C. (2015). Archaeozoology in Mainland Southeast Asia: Changing Methodology and
Pleistocene to Holocene Forager Subsistence Patterns in Thailand and Peninsular Malaysia.
Open Quaternary, 1, 123. doi:http://dx.doi.org/10.5334/oq.af.

84

6 Was There a Broad Spectrum Revolution in Southwest Europe?

Corts-Snchez, M., Morales-Muiz, A., Simn-Vallejo, M. D., Bergad-Zapata, M. M., DelgadoHuertas, A., Lpez-Garca, P., et al. (2008). Palaeoenvironmental and cultural dynamics of the
coast of Mlaga (Andalusia, Spain) during the Upper Pleistocene and early Holocene. Quaternary
Science Reviews, 27, 21762193. doi:http://dx.doi.org/10.1016/j.quascirev.2008.03.010.
Flannery, K. V. (1969). Origins and ecological effects of early domestication in Iran and the Near
East. In P. J. Ucko & G. W. Dimbleby (Eds.), The domestication and exploitation of plants and
animals (pp. 73100). Chicago: Aldine.
Freeman, L. G. (1973). The significance of mammalian faunas from Paleolithic occupations in
Cantabrian Spain. American Antiquity, 38, 344. doi:10.2307/279309.
Gutirrez-Zugasti, I. (2011). Coastal resource intensification across the PleistoceneHolocene
transition in Northern Spain: Evidence from shell size and age distributions of marine gastropods. Quaternary International, 244, 5466. doi:http://dx.doi.org/10.1016/j.quaint.2011.04.040.
Jones, E. L. (2006). Prey choice, mass collecting, and the wild European rabbit (Oryctolagus
cuniculus). Journal of Anthropological Archaeology, 25, 275289.
Jones, E. L. (2007). Subsistence change, landscape use, and changing site elevation at the
Pleistocene-Holocene transition in the Dordogne of Southwestern France. Journal of
Archaeological Science, 34, 344353.
Jones, E. L. (2009). Climate change, patch choice, and intensification at Pont d'Ambon (Dordogne,
France) during the Younger Dryas. Quaternary Research, 72, 371376.
Jones, E. L. (2013). Mobility, settlement, and resource patchiness in Upper Paleolithic Iberia.
Quaternary International, 318, 4652. doi:http://dx.doi.org/10.1016/j.quaint.2013.05.027.
Manne, T., Cascalheira, J., vora, M., Marreiros, J., & Bicho, N. (2012). Intensive subsistence
practices at Vale Boi, an Upper Paleolithic site in southwestern Portugal. Quaternary
International, 264, 8399. doi:http://dx.doi.org/10.1016/j.quaint.2012.02.026.
Marn Arroyo, A. B. (2013). Human response to Holocene warming on the Cantabrian Coast
(northern Spain): An unexpected outcome. Quaternary Science Reviews, 81, 111. doi:http://
dx.doi.org/10.1016/j.quascirev.2013.09.006.
Rillardon, M., & Brugal, J.-P. (2014). What about the Broad Spectrum Revolution? Subsistence
strategy of huntergatherers in Southeast France between 20 and 8 ka BP. Quaternary
International, 337, 129153. doi:http://dx.doi.org/10.1016/j.quaint.2014.01.020.
Sommer, R. S., Fritz, U. W. E., Sepp, H., Ekstrm, J., Persson, A., & Liljegren, R. (2011). When
the pond turtle followed the reindeer: Effect of the last extreme global warming event on the
timing of faunal change in Northern Europe. Global Change Biology, 17, 20492053.
doi:10.1111/j.1365-2486.2011.02388.x.
Sommer, R. S., & Nadachowski, A. (2006). Glacial refugia of mammals in Europe: Evidence from
fossil records. Mammal Review, 36, 251265. doi:10.1111/j.1365-2907.2006.00093.x.
Sommer, R. S., & Zachos, F. E. (2009). Fossil evidence and phylogeography of temperate species:
'glacial refugia' and post-glacial recolonization. Journal of Biogeography, 36, 20132020.
Stiner, M. C., & Munro, N. D. (2011). On the evolution of diet and landscape during the Upper
Paleolithic through Mesolithic at Franchthi Cave (Peloponnese, Greece). Journal of Human
Evolution, 60, 618636. doi:10.1016/j.jhevol.2010.12.005.
Straus, L. G. (1977). Of deerslayers and mountain men: Paleolithic faunal exploitation in
Cantabrian Spain. In L. R. Binford (Ed.), For theory building in archaeology (pp. 4176).
New York: Academic Press.
Straus, L. G. (2009). The Late Upper Paleolithic-Mesolithic-Neolithic transitions in Cantabrian
Spain. Journal of Anthropological Research, 65, 287298.
Straus, L. G., Bicho, N., & Winegardner, A. C. (2000). The Upper Paleolithic settlement of Iberia:
First-generation maps. Antiquity, 74, 553566.
Straus, L. G., & Clark, G. A. (Eds.). (1986). La Riera Cave: Stone age hunter-gatherer adaptations
in northern Spain. Tempe, AZ: Arizona State University Anthropological Research Papers.
Zaatari, S., & Hublin, J.-J. (2014). Diet of Upper Paleolithic modern humans: Evidence from
microwear texture analysis. AJPA American Journal of Physical Anthropology, 153, 570581.
Zeder, M. A. (2012). The Broad Spectrum Revolution at 40: Resource diversity, intensification,
and an alternative to optimal foraging explanations. Journal of Anthropological Archaeology,
31, 241264. doi:10.1016/j.jaa.2012.03.003.

Glossary

ANOVA or analysis of variance A statistic which tests for differences between


group means
Aquitaine The southwesternmost region of France, bordered by the Atlantic to
the west and the Massif Central to the east
Bioclimatic region A geographic region in which climates are generally similar. In Southwest Europe there are two macrobioclimatic regions: the temperate/
Euro-Siberian region and the Mediterranean region
Biogeographic region A geographic region in which biotas (that is, flora and
fauna) are generally similar. In Southwest Europe, there are five (see Fig. 1.1)
Body-size proxy In optimal foraging theory, the use of prey body size as a proxy
for caloric return
Blling-Allerd An interstadial, or warming period, that began ca. 13 kya (uncalibrated) and ended with the onset of the Younger Dryas at ca. 11 kya (uncalibrated)
Broad Spectrum Revolution A transition from diets focused on a small number
of high-return taxa to a wider range of taxa, hypothesized to be a precursor to
agriculture
Cluster analysis A set of multivariate analytical techniques which group sets of
objects in such a way that objects in the same group (or cluster) are more similar
to each other than to those in other groups
Diet breadth How wide, or how many taxa, the diet contains. In the prey choice
model, increasing diet breadth may indicate resource depression. In zooarchaeology diet breadth is typically measured by NTAXA
Diversity The degree of heterogeneity in some group or area. In zooarchaeology,
diversity is often divided into two measurements: richness and evenness
Epipaleolithic The transitional period between the Upper Paleolithic and
Mesolithic; in this work, it comprises the periods known as Final Magdalenian
and Azilian and the time period between 13 and 10 kya. See Table 1.1 for more
details

The Author(s) 2016


E.L. Jones, In Search of the Broad Spectrum Revolution in Paleolithic Southwest
Europe, SpringerBriefs in Archaeology, DOI 10.1007/978-3-319-22351-3

85

86

Glossary

Euro-Siberian Iberia In this work, a paleo-bioregion comprising most of the


northern portion of the Iberian Peninsula; corresponds with the temperate/EuroSiberian bioclimatic region of today (see Jones 2015; Rivas-Martnez et al. 2004)
Evenness A measure of diversity that assesses how evenly taxa are represented; in
this work, I use the reciprocal of Simpsons (1/D) to measure evenness
Faunal turnover The replacement of one set of faunas by another; common along
latitudinal and elevation gradients
Gravettian In this work, the Upper Paleolithic culturehistorical period between
ca. 29 and 21 kya
Handling time In the prey choice model of optimal foraging theory, the time it
takes to successfully pursue, capture, and process a prey item after encounter
Holocene The current geological epoch; began ca. 10 kya
Ice core A core sample removed from an ice sheet, often used in paleoclimate
research
Kya Thousands of years ago
KruskalWallis A nonparametric method for testing whether samples come from
the same distribution; sometimes called the nonparametric ANOVA
Lagomorph A member of the taxonomic order Lagomorpha; includes hares, rabbits, and pikas
Landes desert A portion of Aquitaine that for much of the Late Paleolithic was
comprised of sand dunes and loess accumulations, making it a formidable barrier
for prehistoric huntergatherers
Last Glacial Maximum or LGM When glaciers were at their maximum extents
during the last glacial period, around 20 kya
Late Paleolithic In this work, the period from 29 through 10 kya; comprises the
Upper Paleolithic and Epipaleolithic
Magdalenian In this work, the Upper Paleolithic culturehistorical period
between ca. 18 and 13 kya
Marine core A core sample removed from the sea or ocean floor, often used in
both paleoclimate and paleovegetation research
Mass collecting or mass harvest A technique such as net-hunting which allows
a number of individual prey items to be taken at one time; may disrupt the bodysize proxy
Massif Central An elevated region in south-central France, containing both
mountains and plateaus. This region served as a biological barrier between
southwestern France and more eastern parts of Europe
Mediterranean Iberia In this work, a paleo-bioregion corresponding with the
Mediterranean bioclimatic region of today
MorisitaHorn index An abundance-based similarity index, used here as the
basis for archaeofaunal similarity analyses (i.e., cluster and NMMDS analyses)
Nestedness A measure of biotic structure which assesses the degree to which
assemblages are subsets of each other

Glossary

87

Niche construction theory A body of theory which sees human environmental


engineering as the key to understanding dietary change
NISP Number of identified specimens, a basic zooarchaeological quantitative
unit
Non-analogous environments Environments for which there are no present-day
equivalent; argued to have been frequent in the Pleistocene
Non-metric multidimensional scaling or NMMDS A non-metric (and nonparametric) means of visualizing the similarity of cases in a dataset
NTAXA The number of taxa. Used in zooarchaeology as a measure of richness
Palynology The analysis of fossil pollen
Patchiness In this work, a patchy environment is one in which resources are distributed nonuniformly across a landscape
PleistoceneHolocene transition The transition from the last glacial period to
more modern climates at ca. 10 kya
Prey choice model An optimal foraging theory model that predicts which
resources a forager will pursue, given certain assumptions and parameters
Prey mobility The speed and manner in which a prey moves; can contribute to
increased handling time for hunters
Pyrenees The mountain chain separating France from the Iberian Peninsula;
serves as an ecological barrier today and in the past
Resource depression In optimal foraging theory, reductions in prey capture rates
by foragers due to the foragers own subsistence activities
Richness The component of diversity that comprises the number of types in a
group. In zooarchaeology, usually measured as NTAXA
Similarity index Any number of measures that provides a quantitative estimate of
how similar (or dissimilar) the members of a group are in composition. Most are
based on either incidence or abundance. In this work, I use the MorisitaHorn
index
Solutrean In this work, the culturehistorical period extending from ca. 2118 kya
and coinciding with the LGM
Southern France In this work, the paleoenvironmental region including
Aquitaine, the French Pyrenees, the French Mediterranean, and the western edge
of the Massif Central
Taxon A taxonomic category or group, such as a phylum, order, family, genus, or
species. Taxa is the plural
Time-averaging The mixing of assemblages deposited over a span of time
together, resulting in one assemblage from a range of time periods
Upper Paleolithic In this work, comprises the Gravettian, Solutrean, and
Magdalenian periods, between ca. 29 and 13 kya
Younger Dryas A cold climate phase occurring from ca. 1110 kya, in which the
Blling-Allerd warming abruptly reversed

Index

A
Age, 7, 1517
Aquitaine, 6, 53, 55
Archaeofaunal, 7, 10, 17, 23, 3755, 6175
Archaeofaunas, 2, 3, 7, 10, 17, 4355, 62, 6575
Archaeologists, 2, 5, 1014, 17, 18, 38, 62

B
Barriers, 4, 29, 30, 53, 74, 81, 83
Bio-climatic regions, 3, 27, 29, 30, 37, 49, 53
Bioregions, 4, 27, 29, 82
Body-size proxy, 1315, 18
Blling/Allerd, 27, 28, 31, 37, 43, 55, 70, 81
Broad spectrum revolution, 7, 1214, 18, 23,
37, 55, 61, 62, 69, 7983

C
Cervus elaphus, 9, 29, 37, 48, 66, 67, 69
Change, 2, 3, 7, 12, 18, 38, 41, 43, 55, 61,
6670, 7275, 7983
Changing environments, 12, 38
Children, 17
Climate, 3, 7, 2331, 38, 43, 55, 65, 70, 83
Climate change, 2, 11, 12, 23, 43, 81, 83
Cluster, 49, 5153, 55, 79, 81
Cluster analysis, 4850, 54, 55, 79
Corridors, 29, 30, 74

D
Decreasing, 28, 73, 83
Desert, 4, 23, 28, 31, 37, 53, 81
Diet breadth, 15, 61, 63, 6574, 79, 81, 83

E
Ebro basin, 30
Environmental changes, 23, 24, 37, 49, 55,
8283
Environmental conditions, 83
Environmentally driven resource stress, 82
Environmental stress, 8283
Environments, 14, 7, 9, 10, 12, 18, 2331, 37,
38, 41, 53, 55, 75, 8083
Epipaleolithic, 6, 7, 11, 25, 28, 37, 38, 4043,
49, 5155, 61, 63, 6575, 8083
Euro-Siberian, 27, 29, 30, 37, 65, 69, 81
Euro-Siberian Iberia, 28, 31, 37, 3942, 46,
4951, 53, 6365, 6772, 74, 8083
Evenness, 63, 64, 6668, 7072, 74, 80, 81

F
Faunal turnover, 7275, 82
Fossil pollen, 2426
France, 4, 53, 69
French, 49, 5355, 66, 68, 71, 72, 82

G
Gender, 1517
Glorious days of mammoth hunting, 12

H
Handling time, 14, 15
Hares, 15, 29, 43, 48, 67, 70, 72, 73, 81
Holocene, 11, 27, 42, 43, 55, 66, 74, 81, 83
Human demographic pressure, 61, 82
Human selection, 23

The Author(s) 2016


E.L. Jones, In Search of the Broad Spectrum Revolution in Paleolithic Southwest
Europe, SpringerBriefs in Archaeology, DOI 10.1007/978-3-319-22351-3

89

90

Index

Huntergatherer mobility, 13, 7, 9, 11, 12,


15, 24, 30, 31, 37, 38, 4143, 65, 74,
80, 83
Hunting, 10, 12, 15, 17, 31, 55, 70, 73, 74,
80, 83

O
Oldest Dryas, 2628
Optimal foraging theory, 13, 14, 18
Oryctolagus, 6972
O. cuniculus, 16, 17, 31, 48, 67, 70

I
Iberia, 4, 25, 27, 37, 43, 49, 65, 81
Iberian Mediterranean, 53, 67, 80, 82
Iberian Peninsula, 14, 27, 2930, 37, 43, 53, 55
Ice cores, 2426
Increase in forager population, 14

P
Paleoenvironmental reconstruction, 27
Paleolithic, 17, 912, 2331, 3755, 6175,
8083
Points, 11, 14, 17
Population pressure, 12, 82
Prey choice model, 1318, 61
Prey type, 13, 14, 65, 67, 74, 75
Proxy, 10, 24, 26, 27, 69
Pyrenees, 4, 9, 27, 3031, 37, 53, 70, 81

L
Landes, 31, 37, 81
Landes desert, 28, 31
Landscape, 17, 15, 23, 24, 2731, 38, 53,
6970, 7375, 83
Last glacial maximum (LGM), 5, 26, 28, 30,
37, 38, 42, 49, 80
Lepus, 16, 48, 67, 70, 71, 81
Location, 3, 4, 7, 13, 16, 23, 26, 27, 3743, 52,
55, 75, 79, 82, 83
Logistical, 38, 41, 74, 80
Low, 12, 24, 38, 41, 43, 53, 67, 69, 81, 83

M
Mammalian, 62, 69, 70, 72, 74, 8083
Marine cores, 25, 26
Mass capture, 15
Mediterranean Iberia, 28, 31, 3743, 47,
4951, 53, 54, 6368, 71, 72, 74, 80, 83
Mediterranean Sea, 29, 49
Men, 15, 17
Mesolithic, 1112
Mobility, 13, 7, 9, 11, 12, 14, 15, 24, 30, 31, 37,
38, 4143, 53, 65, 74, 8083.
See also Huntergatherer mobility
logistical mobility, 38, 41, 74, 80
prey mobility, 1415
residential mobility, 38, 41, 42, 53, 74, 80, 81
Model, 7, 13, 79

N
Neolithic, 11, 12
Nestedness, 7274, 79, 81, 83
Niche construction theory, 18
Non-analogous environment, 31, 37, 80
Non-metric multidimensional scaling
(NMMDS), 48, 49, 51, 53, 79

R
Rabbits, 16, 17, 28, 31, 43, 48, 6770, 80,
81, 83
Rangifer tarandus, 9, 28, 48, 65
Ranking, 13, 15, 17
Reconstructions, 2427, 29
Red deer, 9, 29, 37, 48, 67, 72, 75
Regions, 3, 4, 7, 11, 15, 2631, 3739,
4147, 49, 51, 5355, 6371, 7375,
7981, 83
Reindeer, 2, 9, 11, 28, 48, 6567, 72, 73,
80, 81
Residential, 38, 41, 80
Residential mobility, 38, 42, 53, 74, 81
Resource depression, 14, 17, 18, 61, 67, 69
Resource stress, 7, 69, 70, 7275, 8283
Richness, 910, 6366, 70, 72, 74, 80

S
Significant change, 65, 67, 71, 72, 74, 8082
Site elevation variance, 38, 41, 43, 53, 8083
Southern France, 1, 3, 2531, 3744, 4951,
5355, 6375, 8083
Southern Francec, 25
Stress, 7, 910, 69, 70, 72, 7475, 8283
Studies, 25, 9, 14, 15, 25, 28, 48, 53, 61, 63
Subsistence, 23, 9, 11, 12, 14, 31, 3755,
7982

U
Ungulates, 43, 48, 62
Upper Paleolithic, 4, 6, 7, 9, 11, 37, 42, 4953,
55, 62, 63, 65, 6771, 74, 75, 8083

91

Index
V
View, 12

Y
Younger Dryas, 2729, 37, 43

W
Wild European rabbit, 16, 28, 31, 48, 67, 68,
80, 81, 83
Women, 15, 17

Z
Zooarchaeology, 2, 3

You might also like