People and the Land through Time: Linking Ecology and History, Second Edition [Second Edition] 9780300249590

A revised and updated edition of a classic book that defines the field of historical ecology People and the Land throug

222 11 6MB

English Pages 336 [330] Year 2019

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

People and the Land through Time: Linking Ecology and History, Second Edition [Second Edition]
 9780300249590

Citation preview

People and the Land through Time

This page intentionally left blank

EMILY W. B . (RU SSELL) SO UTHGATE

People and the Land through Time LINKING ECOLOGY AND HISTORY SECOND EDITION

New Haven & London

Published with assistance from the foundation established in memory of James Wesley Cooper of the Class of 1865, Yale College. Copyright © 2019 by Emily W. B. (Russell) Southgate. All rights reserved. This book may not be reproduced, in whole or in part, including illustrations, in any form (beyond that copying permitted by Sections 107 and 108 of the U.S. Copyright Law and except by reviewers for the public press), without written permission from the publishers. Yale University Press books may be purchased in quantity for educational, business, or promotional use. For information, please e-mail [email protected] (U.S. office) or [email protected] (U.K. office). Set in Postscript Sabon type by Integrated Publishing Solutions. Printed in the United States of America. Library of Congress Control Number: 2019932084 ISBN 978-0-300-22580-8 (paper : alk. paper) A catalogue record for this book is available from the British Library. This paper meets the requirements of ANSI/NISO Z39.48-1992 (Permanence of Paper). 10 9 8 7 6 5 4 3 2 1

Contents



Preface to the First Edition  vii



Preface to the Second Edition  xi



Acknowledgments xv Part I  Questions and Clues  1

1 History Hidden in the Landscape  3 2 Historical Records and Collections  18 3 Field Studies: Bringing Historical Records Down to Earth  34 4 The Sedimentary Record  48 Part II  The Diversity of Human Interactions with the Natural World  61 5 Fire: Mimicking Nature  63 6 Extending Species’ Ranges  83 7 Harvesting Natural Resources  104

vi  Contents

8 Agriculture and Its Residual Effects  126 9 Patterns of Human Settlement and Industrialization  150 Part III  Contributions of Historical Ecology to Understanding Ecological Issues  173 10 Diversity and Species Extinctions  175 11 Biospheric Sustainability in a Changing World  197 Conclusion Toward the Future: Research and Applications  219 Notes 233 References 251 Index 299

Preface to the First Edition

When I started studying ecology, I hoped to be able to explain the composition of plant communities by understanding the interactions of species’ physiology and population dynamics with microenvironments. Reading and research, however, have convinced me that while these interactions are important for determining what species can grow somewhere, the history of a site and region plays a major role in determining what species actually do grow there. This idea is not new, even within the discipline of ecology, but until recently ecologists have downplayed it in their efforts to discover general laws that govern species distributions, ecosystem properties, and other ecological processes, regardless of time or space. On the other hand, historians are realizing that the environment in which people live has influenced human history, so that they too must be sensitive to changing environmental conditions. Concern about a deteriorating environment caused by human activities pervades our current view of the world. Many people are of the opinion that unless we mend our ways, we risk disaster. Many also see scientific research, especially ecological research, as the potential source of solutions to environmental problems. In dealing with scientific research related to the environment as altered by people, however, scientists are faced with an overlay of causation that has varied over time and space with changing human culture.

vii

viii  Preface to the First Edition

I have written this book to help point to different aspects of current environments that bear the imprint of various past human activities, which must be considered in order to understand the current processes. The emphasis is on remnant effects on current communities, ecosystems, and landscapes and on how factoring these effects into ecological studies can help elucidate processes. Along the way, it should become clear how differently people have viewed, understood, and used the nonhuman environment and how these differences contribute to impacts as well as, in complex webs of feedback, to changing activities and attitudes. In conducting historical ecological research, I have been convinced of the importance of distinguishing between time as a measurable dimension of duration, such as one day or one year, and historical time as a specific duration, such as 6 June 1952 or 1735. The partitioning of the processes that we observe between those that are based on the nonspecific unit of time and those that are historically constrained will help us tremendously in relating theoretical studies to actual responses of real ecosystems. I have two main goals in this book, one related to research and the other to environmental management. I hope to stimulate further research on the role that history, specifically human history, has played in shaping communities, ecosystems, and landscapes and conversely, the role that changing environments have played in human history. I have tried to do this by pointing to the ubiquity of residual as well as current human interactions with the environment and by demonstrating that these impacts have changed over time, up to and continuing in the present. Second, those who plan and manage natural areas should learn that their systems are never static and that the present conditions are merely stages in a continually changing mosaic. They cannot be frozen in time. My examples are drawn from all over the world, from a wide variety of biomes, though emphasis is placed on temperate systems, especially in the eastern United States and in western Europe, as these are the ones with which I am most familiar. They are discussed as illustrations; references are given for readers who would like more definitive discussions of the individual examples. The concepts apply, however, anywhere. I expect this approach to be useful both as an introduction to historical ecology for professional ecologists, environmental historians, historical geographers, and historical anthropologists and as an advanced undergraduate and graduate textbook for such courses as historical ecology and environmental issues. I start with an exposition of the importance of considering the past of ecosystems and then introduce techniques that can be used for reconstructing this past. I then discuss a variety of ways in which people have

Preface to the First Edition  ix

affected the environment over time, from using fire to laying out property boundaries. I conclude by discussing how a historical ecological approach contributes to an understanding of some issues of current concern: changes to lakes, biodiversity, and sustainability. I hope that readers will carry away an excitement for including human history in ecological studies and ecology in historical studies. This integration of the disciplines has great potential for both and presents challenges that must be met if we are to deal responsibly with our role in the biosphere.

This page intentionally left blank

Preface to the Second Edition

In the two decades since publication of the first edition of this book, there has been an explosion in research in historical ecology. Google Scholar sends me notification of one or more new publications that use the term historical ecology every day, and the four hundred or so new references in this edition are a small sampling of recent research. New technologies and a recognition that history is an important consideration in conservation decisions have contributed both to interest in the field and to advances in our understanding of the importance of history for influencing current—and future— landscapes and ecosystems. Recent academic meetings have included sessions on historical ecology, and the Frontiers in Historical Ecology Symposium in Switzerland in 2011 featured worldwide research in historical ecology that went beyond case studies, but included these as well. Accelerated change in the global environment over the last century has also stimulated appreciation for the value of looking at the past to help understand how we have gotten to where we are today. There continues to be a dichotomy between those who view historical ecology from an anthropological point of view and those who view it from the science of ecology. The most recent book with an anthropological perspective on historical ecology describes it as a research framework for studying the “human-environment relationship,” important for designing land management

xi

xii  Preface to the Second Edition

decisions that are “effective and equitable,” with a focus on specific locations.1 Environmental history also focuses on the human aspect, especially how interactions with the natural world have affected human history. Most ecologists who do historical ecology, on the other hand, focus more on ecological processes, and how incorporating changing human impacts over time can help explain current processes and patterns. Many ecologists, however, especially in the United States, find it difficult to accept the importance of history to the systems that they study. I hope that the wealth of studies that I discuss in this book will show that historical ecology provides critical insights into the structure and function of ecosystems as well as conservation decisions. With some regret, I have eliminated the chapter on lakes from this edition. Since the first edition, there have been at least two books devoted to paleolimnology, written by limnologists who understand lake systems far better than I do (two editions of J. P. Smol, Pollution of Lakes and Rivers: A Palaeoenvironmental Perspective, 2002, 2008; and A. S. Cohen, Paleolimnology: The History and Evolution of Lake Systems, 2003). The Journal of Paleolimnology has been publishing specifically historical studies of lakes since 1988. I have incorporated some lake studies in the other parts of the book, but have not delved into detail on lake ecology as I did in that chapter of the first edition. Most of my research and field experience has been in northeastern North America, where the dominant vegetation is deciduous forest. This has led to much of the illustrative material in the book coming from this biome, though I include many examples of historical ecological research from other continents and other biomes. I hope that the abundance of examples will provide sufficient evidence of the importance of research in these other regions as well as an entrée into the historical ecological literature of these areas. I have focused in the chapter on sediment more on pollen than on other records in the sediment, also because that is where my experience lies. In my defense, I think that pollen is a good character to emphasize, as it gives an apparently simple representation of vegetation surrounding the sedimentary basin, while actually having a very complex relationship with that feature of the environment. The process of translating pollen data into meaningful interpretations of the contributing vegetation highlights the kind of analysis needed for using any sedimentary proxy for the environment. I originally thought that that revising the first edition required merely updating some references in cases where work had been done on a topic since I originally referred to it. It became abundantly clear to me, though, as I considered changes in the field since the late 1990s, that the fields of historical ecology and environmental history had moved ahead to such an extent that

Preface to the Second Edition  xiii

I would need to do a more extensive rewrite. I have learned a lot in doing the research for the second edition, and I am more convinced than ever that studying ecology without reference to history misses vital insights into the functioning of ecological systems at all scales, from population to global. Similarly, conservation decisions taken without regard to the history of a region may be unsustainable. Based on the outdated concept of “climax vegetation,” they often ignore change in the past.2 This book is organized the same as the first edition, with an introduction to the field followed by chapters on methodology. Subsequent chapters deal with specific factors of human impact and the importance of history for understanding some important current ecological issues. The first four chapters set the stage for the rest of the book; the subsequent chapters to some extent can stand alone, though they do progress from the most long-running human impacts to more recent ones to give a feeling for the pervasive role of people in the environment since the evolution of Homo sapiens. I hope that others will accept the challenge to add depth to their understanding by finding out more about the history of the systems that they study.

This page intentionally left blank

Acknowledgments

My interest in historical ecology began when I was a child digging up old horseshoes in our garden and finding old stone walls in the woods where I played. I owe a great debt of gratitude to my parents, who were always enthusiastic supporters of my interest in science. Many people have contributed to the development of my ideas on historical ecology in addition to those who have more recently commented on various versions of this book and have contributed technical expertise. My teachers at the Baldwin School and at Denison University required me to write and to study the humanities as well as the sciences, preventing too narrow a focus. A year’s study at the University of Paris introduced me to the residual impacts of people on the forests of France and to the idea that all forests have experienced some human impact and that these impacts have changed over time. Further studies at Duke and Rutgers Universities continued this emphasis on the interplay between people and their environments. This book grew directly from a joint biology and history graduate seminar I taught at Duke University in 1990, while I was supported by a National Science Foundation Visiting Professorship for Women. The lively discussions among students with different intellectual backgrounds inspired me to begin the long process of writing a text that would build on and disseminate this enthusiasm for interdisciplinary interactions.

xv

xvi  Acknowledgments

Over the years, discussions with many individuals have contributed to the development of the ideas expressed here. I would like especially to acknowledge W. Dwight Billings, Michael Binford, Grace S. Brush, Norman L. Christensen, Harold L. Cousminer, William Cronon, Ronald B. Davis, Edward S. Deevey, Richard T. T. Forman, David R. Foster, Steven P. Hamburg, Warren Hofstra, Sally P. Horn, Daniel A. Livingstone, Peter L. Marks, Mark J. McDonnell, David Mladenoff, Steward T. A. Pickett, Frederick H. Russell, Robert L. Sanford Jr., Péter Szabó, John C. F. Tedrow, George Theokritoff, Peter L. Tobiessen, Charles Watkins, David Wigston, the attendees at the fourth Cary Conference, “Humans as Components of Ecosystems,” and those at the “Frontiers in Historical Ecology” international conference. Faculty and students participating in seminars in Rutgers University’s Quaternary Studies Graduate Program reinforced the importance and exciting potential of interdisciplinary research. Thompson Webb III read and provided invaluable commentary on the first edition, and Matthias Bürgi did the same with the second. Lars Östlund and several anonymous reviewers provided especially helpful comments on the manuscript. My editor, Jean E. Thomson Black, has been a wonderful source of support, encouragement, and gentle pressure to keep moving on this work. For assistance with the illustrations I thank Susan Hochgraf, F. Mason Barnett, James Gasprich, Joshua Schnalke, and William Nelson. I greatly appreciate the generosity of all those who have allowed me to use data, photographs, or illustrations from their publications; these are individually acknowledged in the appropriate figure captions. Sarah Butler, Olivia Peterson, and Drew Ferrier at Hood College helped with miscellaneous typing, organizing, and technical assistance, and the Beneficial-Hodson Library of Hood College provided invaluable support for my review of the current literature. For financial support during various phases of this project, I thank Yale University Press, the National Park Service, the National Science Foundation, the Mellon Foundation, the Nature Conservancy, the Koven Foundation, the New Jersey Department of Environmental Protection, and the Loewy-Mohonk Preserve Liaison Fellowship.

People and the Land through Time

This page intentionally left blank

PART

I

Questions and Clues

Reconstructing the past to evaluate its effects on present ecosystems is like working on a puzzle for which the researcher knows the outcome but not the rules. It requires the integration of vast amounts of diverse information. Written records and oral histories, both observations of the natural world by people, document past human activities. Some past human activities have left traces on the landscape that persist to the present. And organisms themselves have left a record by their remains in sedimentary deposits. Each kind of record provides a unique type of information, each with its own spatial and temporal scale. By integrating these sources, a historical ecologist can piece together a picture of past activities and communities in order to formulate and test hypotheses about causes of past changes and the contributions of past processes to present ecosystems and landscapes. This process of integration requires careful attention to the compatibility of the different kinds of information in terms of scale and biases, and an awareness of the appropriate kinds of critical analysis used to interpret the various categories of data. In the next four chapters, I introduce the three major categories of sources that inform historical ecological studies: written records, field studies, and sedimentary records. I suggest ways in which each makes unique contributions to these studies as well as offer caveats relative to biases and limitations. These

2  Questions and Clues

chapters do not provide detailed instructions on how to use the different sources; such detailed how-to information may be found in the various texts I mention, and these should be supplemented by courses and consultation with practitioners in the corresponding fields of study. But first I will raise some of the questions that historical ecologists ask.

1

History Hidden in the Landscape

From the vast boreal forests of Russia and Canada to the deserts of Africa and North America, human impact on the earth is ubiquitous and apparent. The prevalence and intensity of recent human impacts are so great that designation of a new geologic epoch, the Anthropocene, has been proposed to distinguish the current level of human impact from that of the past. The impacts that are apparent today, however, are superimposed on millennia of previous human activities and climate changes that have left legacies, some subtle, some not. Reminders of these past human activities persist even in such apparently pristine environments as the tropical forests of Africa and South America and many regions designated as wilderness in North America (fig. 1.1).1 Historical ecology seeks to explain many enigmatic features of present ecosystems and landscapes by deciphering the legacies of past human activities. It also yields insights into basic ecological patterns and processes that cannot be understood by studying only the present or very recent past.2 Apparent forest age, for example, is often not correlated with the diversity of native herbaceous ground flora even when soils, topography, and other factors are held constant. However, when one considers the kind of disturbances in the past, such as plowing, logging, or grazing, even those that occurred more than a century ago, the patterns become much clearer. Many highly diverse ecosystems of Europe depend on human-driven processes for their

3

4  Questions and Clues

Figure 1.1. Buildings of the Maya city of Ek Balam in Mexico, engulfed by regenerated forest. (Photo by Mason Barnett, 2017.)

very existence, having developed over many millennia under the influence of agriculture or wood harvesting. These human-imposed drivers interact with climate to produce unique landscapes. To understand the processes that maintain their diversity and stability, one must also understand the historical processes that have created them—in other words, the underlying causes.3 Past human influences on ecosystems and landscapes are cumulative and superimposed on one another and on changes in climate. This superposition is often referred to as a palimpsest, but this concept does not capture the full import of the relationship between the past of a landscape and the present. With a palimpsest, traces of previous use of a canvas may be discernable, but they do not usually form an integral part of subsequent work, while the past often has a strong influence on the subsequent conditions of an ecosystem. A typical suite of activities in many forested areas of eastern North America, for example, was logging in the nineteenth century, followed by fires, often followed by grazing, and finally culminating in apparent recovery of the forest. Logging favors some species such as chestnuts (Castanea dentata), which sprout prolifically from cut stumps, or birches (Betula spp.), which reproduce by seed in open sites. Fires eliminate the fire-sensitive hemlock (Tsuga canadensis) and beech (Fagus grandifolia). Grazing eliminates palatable species. No one of these activities taken alone can explain characteristics of the resulting forest, nor can several of them taken out of order. Similarly, in the oceans, exploitation of marine resources in the past changed food webs and population structure, which in turn influenced subsequent marine life

History Hidden in the Landscape  5

and patterns of human use. One cannot understand the current communities without considering the suite of past human-influenced alterations.4 The ubiquity of human impact means that it is very difficult, if not impossible, to find systems devoid of human influence. Two sites selected to compare the effects of different species composition on nutrient cycling will differ in past land use as well as in species composition, so a simple comparative study cannot yield results that are attributable simply to differing species composition. Apparently “ancient” forests may conceal even more ancient intense human activities that may still influence forest structure and function.5 These consequences of human activities superimposed on past climate change open up an exciting arena of integrative research. By interpreting the historical record, we can infer past human activities and climate, including spatial and temporal patterns at various scales, and through comparative studies we can establish how they have worked to shape the present. The study of past human impact is an entry point into the great diversity of possible interactions of species with their environments, which can assist us in making predictions that are relevant to a world permeated by human influences.6

Humans as Geographical Agents The idea that people have had an impact on the natural world, and that that impact is not transient, is not new. In 1853, Mary Somerville observed in Physical Geography that humans had exerted a major geographical influence on the earth. Several years later, George Perkins Marsh elaborated on this topic in Man and Nature, noting, “Not all the winds, and storms, and earthquakes, and seas, and seasons of the world, have done so much to revolutionize the earth as Man, the power of an endless life, has done since the day he came forth upon it, and received dominion over it.”7 Until that time, the assumption had generally been that although people had local impacts on the potential of land, the loss of agricultural potential in large areas of the world was due to natural causes such as climate.8 Marsh correlated the activities of people with loss of soil, silting of harbors, and other major geographical problems. It was even possible that the actions of people had changed local climates. In many U.S. states around the mid-1800s, concern about such problems led to the establishment of state geological surveys to catalogue natural resources and to make recommendations for their preservation. Official recognition of problems such as erosion indicates that they were taken seriously. The focus was on preservation of natural resources for human use, however,

6  Questions and Clues

rather than on understanding natural processes and conservation of natural systems. Changes in the intensity of human activities over space increasingly confront us today. Regionally, extensive forests alternate with towns and cultivated fields. For example, in the Appalachian Mountains of eastern North America, apparently wild forests contrast with farm fields. In the deserts of the American Southwest, isometric irrigated agricultural fields contrast with the surrounding desert matrix. In Hawai’i, verdant rain forests differ markedly from orderly, low-diversity pineapple plantations (fig. 1.2). The natural, pristine appearances of the Appalachian forests, southwestern deserts, and Hawai’ian forests are, however, to a large extent mirages; all bear distinctive imprints of past human activities. Logging and agriculture have leveled almost all the forests of the Appalachian Mountains sometime during the past two hundred years. In the Southwest, overgrazing and the pumping of water for irrigation have reduced many dry grasslands to desert shrublands. Non-native species have decimated native populations of plants and animals even deep in the forests of Hawai’i.9 Such past human activities have left both obvious and subtle imprints on many aspects of these systems. Which species did logging favor, and which did it eliminate? What have been the processes of extinction and introductions in the past, and how have they affected current community composition? How have these processes changed over time, and which characteristics of the present communities and landscapes reflect them? How have these human-caused changes interacted with fluctuations in climate? Historical ecology addresses such questions to understand the processes that control ecological systems. Some major categories of ecosystems appear to the casual observer to be undisturbed by humans. Most marshes in the northeastern United States, from the pocosins of North Carolina to tidal marshes along the Chesapeake Bay, have been ditched and drained in the past, though the ditches are often no longer apparent. These abandoned ditches, however, have left lasting impacts on the ecology of the extant ecosystems: a changed water table, fire impacts in dried organic soils, and changed species composition from mowing and grazing. On the other hand, over the past two centuries, human activities have also indirectly created new marshes in areas such as the Piedmont of Georgia in North America, by upland erosion and subsequent sedimentation along streams.10 An apparently natural acid fen in Pennsylvania owes its existence to logging less than a century ago, which changed the hydrology, possibly permanently (fig 1.3). Traces of abandoned human settlements both reveal and conceal past uses

a

b Figure 1.2. Island of Maui, Hawai’i. Contrast (a) the lush vegetation in a valley along the northeast coast with (b) pineapple plantations in the interior plains. Non-native species, however, are common in the unmanaged as well as the planted landscape. (Photos by the author, 1992.)

8  Questions and Clues

Figure 1.3. Spruce Flats Bog, Forbes State Forest, Normalville, Pennsylvania. The cut stems in the foreground indicate logging in the past. (Photo by the author, 2016.)

of the land. “Lost villages” abound worldwide, revealed by place-names on old maps, by physical remains, and by pictures (fig. 1.4). They are now, however, often hidden by pastures, plowed fields, or regenerated vegetation. To historians and archaeologists these places are of interest for what they reveal about human behavior in the past. Why did they come into existence and why did they disappear? More geographically oriented questions consider the relation of village locations to spatial features of the landscape. Deserted structures raise ecological questions as well. A village, for example, represents a concentration of people who used natural resources for fuel and food, disturbed the fauna and soil by hunting, plowing, or just trampling, and produced wastes. The influence of these on the regenerated vegetation may persist in patterns that are quite obscure when compared with the current physical environment.

Historical Ecology Informing Conservation Decisions about conservation involve both setting priorities and understanding processes. Both the priorities and the processes depend not only on current values and conditions but also on the past. Much riparian conservation today involves planting trees along streambanks. Where these were not forested in the past, such activities may actually destroy natural communities. For example, eighteenth-century historical sur-

Figure 1.4. Site of Quartz Mill near Virginia City, Nevada. The top picture was taken in 1868, the bottom one at the same place in 1979. Only a few stones remain to mark the site of the structure a hundred years later. (Mark Klett for the Rephotographic Survey Project.)

10  Questions and Clues

Figure 1.5. Survey of property on Abrams Creek in Winchester, Virginia, purchased by Isaac Perkins, 1734. Survey lines superimposed on USGS 7.5 minute topographic map by Galtjo J. Geertsma. Descriptions at points 1, 2, and 3 as given in the survey notes: 1, hickory; 2, “cross a small meadow”; 3, black oak.

veys describe “meadows” along many streams in a limestone area of Virginia (fig. 1.5). At least one of the surviving meadows, Abrams Creek Wetlands Preserve, contains several plant species that are rare in the state, including two that do not grow elsewhere in Virginia.11 Is it better to plant trees in this wetland to improve water quality in this urbanizing watershed or to leave it open to protect the rare plant community? Historical evidence bolsters the floristic study to indicate that protecting the marsh in its open character is a good strategy for conservation. This decision may have even wider implications. Currently there are many grassland birds in northeastern North America, such as the loggerhead shrike (Lanius ludovicianus) and the grasshopper sparrow (Ammodramus savannarum), which have flourished since the extensive forest cover was converted to farmland over the last few centuries. The numbers of almost all of these species are, however, declining rapidly because of both intensification of agriculture and reforestation. Preserving grassland habitats requires intensive management, most likely because they are not located where natural features prevented tree growth in precolonial landscapes. Study of native plant species that prefer open habitats indicates that most of these grew in scattered, at least semi-permanent open areas of various sizes where conditions are either too wet or too dry for trees to grow, for example, Abrams Creek Wetlands.12 Conservation of these habitats, both large and small, will support populations not only of these birds and often rare plants, but also the whole

History Hidden in the Landscape  11

panoply of species that live there. Without the historical reconstructions, the importance of these special sites as compared with more transient ones would not be as well established. Ancient woods are often singled out for conservation value, with the assumption that their diversity and structure have persisted for millennia, or at least many centuries. Management would thus emphasize maintaining the forest continuity. Detailed studies of woods in England, Belgium, and Sweden, however, have documented disturbances, including clear-cutting, in some of these, in one case as recently as a hundred years ago. An ancient forest in northern France regenerated after Roman farming. The species diversity has, however, rebounded from these events, especially the diversity of easily dispersed species such as fungi, lichens, and insects. These examples indicate that the diversity of many of the organisms in a forest is perhaps increased by a history of disturbances. Historical study has elucidated processes that are critical to maintaining the conservation value of these woodlands.13 In much of central Europe, a century of vegetation studies has led to the conclusion that the dominant ancient forest type of this region is deciduous, mainly consisting of oak (Quercus spp.) and European beech (Fagus ­sylvatica). Almost all spruce (primarily Picea abies) forests are considered to be the result of plantations. Thus conservation is focused on deciduous forests. However, historical ecological studies have found that spruce forests were dominant in a highland area of the Czech Republic throughout the Holocene epoch, suggesting that conservation agencies need to reconsider their priorities for reforestation.14 These few examples illustrate the fallacy of assuming that the conditions that have prevailed over the last century or so are characteristic of the longterm past. Reconstructing historical baseline conditions is an important component of evaluating processes leading to vulnerability of ecosystems, the precursor to “ecosystem collapse.”15 The changing pace of technology, human population growth, and attitudes toward the natural world have affected recent conditions, from modifying the physical environment for the production of crops to preserving so-called wilderness. Historical ecologists, environmental historians, historical geographers, and archeologists can put these recent changes into broader temporal context in which we can see larger patterns.

Interdisciplinary Considerations “The causes of the earth’s big environmental problems—deforestation, loss of biodiversity, pollution, climate change, and so forth—are all rooted

12  Questions and Clues

in human behavior.”16 This sentence introduced a news brief in the journal Science referring to a National Research Council report calling for interdisciplinary research on the “human dimensions of global change.”17 The report emphasized the importance of the social sciences, but did not emphasize incorporating historical changes in ecological processes caused by people as critical to understanding today’s ecosystems. Environmental historians and historical geographers consider the conceptual, political, and economic frameworks within which people have acted and how these frameworks have influenced interactions with the natural environment. Historians concentrate on explaining human nature and contemporary human affairs, at the same time revealing to ecologists the underlying human drivers of landscape patterns. Conversely, ecologists can point to the environmental consequences of certain activities, which in turn affect human responses to the environment.18 For example, in Illinois, soil quality and topography are often not suited for the use to which land has been put. Such cultural forces as land surveys, overriding more resource-oriented concerns, influenced the distribution of these land uses.19 Archeologists and anthropologists study past human cultures by analyzing physical remains. From these they can ascertain tremendous detail about resource use, population sizes and locations, and relationships over space and time between different cultures. All of these findings are embedded in and cause alterations to the surrounding nonhuman environment and in turn were influenced by these alterations. They may suggest to historical ecologists past factors of importance for understanding the current environment. There is a wide gulf between stating the need for interdisciplinary work and engaging in it; different disciplines ask different questions, look for answers in different places, and use different criteria to judge the quality of the answers. To cooperate fruitfully, researchers must recognize their differences and respect alternate points of view. Answering an ecologist’s, a historian’s, and a geographer’s questions as they relate to one subject may be of interest, but for us to gain the full value of interdisciplinary research the answers must amplify each other. Cooperation may even lead to questions that themselves are interdisciplinary and would not have been posed in any of the separate disciplines. Thus it is important that the interdisciplinary team work on a project from its inception.20 As an example, in the second half of the nineteenth century in the northeastern United States, the tanning industry cut many thousands of hemlock trees for their bark, an ingredient in the tanning process. Ecologists are interested in this phenomenon because of its effect on the ecosystem dominated by hemlock. Environmental historians might be interested in why the indus-

History Hidden in the Landscape  13

try grew when and where it did: what factors of industrialization, capitalism, and international trade determined the demand for tanbark at that time, rather than earlier? Historical geographers might question the distribution of tanbark sources with relation to transportation and energy sources as well as demand.21 Answers to all of these hypothetical questions fit together to create a story that no one could tell alone. Until the mid-nineteenth century, extract of oak bark was used for tanning hides in small tanneries. By the 1850s, however, farmland had replaced much oak forest, in part because good soil and climate for many oak species corresponded with good agricultural land. Large stands of hemlocks persisted in steep, remote areas like north-central Pennsylvania and south-central New York. The cattle industry by that time was producing large numbers of hides from cattle slaughtered in central stockyards, near railroad and steamboat transportation. Industrialists, using steam power and capital intensification, were able to build tanneries to process these hides only where there were large stands of bark-bearing trees, as it was cheaper to take the hides to the bark rather than vice versa. So the hemlocks were ruthlessly destroyed, not because their bark was favored or because they grew in especially convenient locations, but from a combination of ecological and cultural factors that came together at a certain time in history. Attitudes toward the resource were as important in determining its destruction as was its inherent usefulness.22 The attitudes of historians toward research that includes the natural environment have changed considerably over the past few decades, allowing more fruitful collaborations between historians and ecologists. In 1967, W. A. Walsh, a historian writing on the philosophy of history, asserted that “the historian is not concerned, at any point of his work, with nature for its own sake; only with nature as a background to human activities.” By 1985, some historians no longer saw the environment as merely a backdrop for history but rather as a legitimate focus for research. In that year, the historian K. E. Bailes described the newly emerging field of environmental history as “the study of what impact economics, politics, social structure, technologies, and value systems have had on the natural environment and the use of natural resources.” In the words of A. W. Crosby, “Environmental historians have discovered that the physical and life sciences can provide quantities of information and theory useful, even vital, to historical investigations.” Since that time the field of environmental history has blossomed, with at least two journals devoted to environmental history, Environment and History and Environmental History, established near the end of the twentieth century.23 Historical geographers, on the other hand, have a long tradition of study-

14  Questions and Clues

ing the interactions between past human activities and natural environments. In introducing the major themes of historical geography in 1941, Sauer regarded humans as agents of physical geography, while in 1966, Eyre and Jones observed that “all Geography is historical geography, in that nothing in the present can be explained without reference to the past.” In addition, they considered all geography to be human ecology. The emphasis, however, as with environmental history, remains on the human element, with a special emphasis on spatial relationships as displayed in maps.24 Archeologists have unearthed remains of ancient (and not-so-ancient) civilizations, often buried under apparently “pristine” vegetation. They have elucidated uses of natural resources, from forest products to soil and minerals, by people in the past that have left legacies on soil properties, species composition, and many other facets of the environment. While archeologists do not focus on these legacies, their research provides a backdrop against which to evaluate current ecosystems.25 Historical ecology shifts the emphasis to the natural environment itself. Turning Walsh’s statement inside out, human activity is mainly of interest to ecologists only as it affects nature. That ecologists recognize the importance of a historical approach today may be indicated by special issues of journals, such as Journal of Biogeography, Landscape Ecology, and Ecological Applications, dedicated to historical ecology. The International Association of Landscape Ecologists has established a “working group” in historical landscape ecology to alert members of current research. One fruitful approach to studying historical ecology is to see humans as geographical drivers of change, similar to climate and soils, with the added twist that human impact changes on different scales and in surprising, and usually unpredictable, ways. The development of the science of ecology in North America, Great Britain, and continental Europe exhibited an ambivalence toward human impact, in part conditioned by local conditions. Early twentieth-century ecologists in North America, led by Frederic Clements, predicated a stable, natural type of vegetation for a region, determined above all by the climate, called the climax vegetation. This appealing simplification of the bewildering variety of current vegetation won wide acceptance in North America. Maps of the vegetation of the United States, for example, generally indicate the “potential natural vegetation,” that is, the hypothetical climax, rather than the actual plant cover. A major goal of ecology was, then, to understand the climatic determinants of the climax vegetation of a region, and by comparing these factors over space to arrive at general organizing principles to explain the characteristics and dynamics of different plant communities. Ecologists in North America thus sought “natural” stands to study or test their theories.

History Hidden in the Landscape  15

There has also been the implicit assumption that once freed of active human management, vegetation would develop along lines dictated by natural forces, trending toward the hypothetical climax composition; the past human element could essentially be ignored. In other words, there is a teleological tendency in nature that is unaltered and unalterable by human actions, assuming that climate is constant. This attitude continues to influence ecology in the United States. For example, the United States National Vegetation Classification system has distinct categories for “natural” and “cultural” vegetation, not acknowledging that historical, cultural activities may be critical for determining even the vegetation classified as “natural” and that there is abundant evidence that climate and other forcing factors do not remain stable long enough for extensive climax vegetation to develop.26 On the other hand, in western Europe and other regions with a more obviously pervasive human impact, the idea of climax vegetation has had a rather different history, at least in part because there were few, if any, so-called virgin stands left. Maps of the vegetation of France, for example, made according to the Zürich-Montpellier school of phytosociology, describe and classify existing, not potential, land cover. The European Union’s Natura 2000 program for natural area protection is, however, based mainly on the concept of climatic climax vegetation, although “natural habitats” includes habitats that are both “entirely natural” and “seminatural.”27 The focus, however, is on natural habitat, controlled by “geographic, abiotic, and biotic features.” This has led to conflict where some of the very biodiverse ecosystems have a long and critical history of human-caused impacts, such as grazing. In England, the oldest woodlands date at least from the Middle Ages, having been actively managed for various timber products over hundreds of years. Some of the most interesting ecological studies there focus on species distribution under different kinds of human management on different substrates. The British ecologist Arthur Tansley exemplified the ambivalence toward the role of human impact when he stated, “It is true that ecological problems are complicated by man’s activity. . . . But the plants themselves are working in the same way, tending towards the same effects, whether man is at work or not.”28 This idea is captured by the dual concepts of the fundamental and realized niches of species. Most simply, the fundamental niche refers to the interaction between the physiology of a species and the limiting environmental variables that allow the species potentially to survive indefinitely. This corresponds to Tansley’s idea of the plants “working in the same way, tending towards the same effects.” The realized niche is the subset of conditions in which the species is actually found. This includes Tansley’s complications arising from

16  Questions and Clues

human activity as well as migrations and competitive interactions with other species. Historical studies have, however, indicated that species’ niches may be at least to some extent malleable, referred to as “niche shift.” For example, it appears that some species that have as a fundamental niche open, disturbed habitats flourished where agriculture was established thousands of years ago and have most likely over the millennia accumulated traits that favored success in these open, stable habitats. It is a mistake to assume either that the current habitats of species represent the totality of their climate ranges or that species are not locally adapted.29 Anthropological evidence suggests that humans, including those in hunting and gathering cultures, have long exerted an influence on the land. Therefore, even before the advent of intensive agriculture eight thousand years ago, it is unlikely that all aspects of landscapes were ascribable only to nonhuman causes. In 1956, W. L. Thomas, Jr.’s interdisciplinary Man’s Role in Changing the Face of the Earth brought together the thoughts of outstanding scholars on these impacts, but it had little direct or immediate influence on academic disciplines, especially ecology. By the 1970s, however, the topic was attracting more attention from ecologists, especially paleoecologists, leading G. E. Likens and M. B. Davis to conclude that “[it] is clear that the magnitude of changes caused by man in decades are equivalent to those occurring over thousands of years without him.”30 Although many ecologists today still ignore human impact in attempting to explain the natural world according to scientific principles, ignoring the history of their field sites is apt to lead to erroneous conclusions. Emphasizing the acceleration of change in recent centuries, B. L. Turner et al.’s The Earth as Transformed by Human Action, published in 1990, focused on only the past three hundred years.31 Historical ecological research includes several stages. The first is simply discovering what conditions were like sometime in the past by finding and interpreting information about a specific place and time. Once the various sources paint a similar picture, the next challenge is to determine the drivers and processes that were effective over time. This stage often includes comparing present patterns and the processes that drive them to past patterns and modeling the various possible relationships. Testing the models requires comparisons across time and space as well as collecting more information to corroborate or disprove the putative processes. Especially in the twenty-first century, historical ecologists have moved into this second phase of research. A third stage is to look for general patterns that will allow prediction. This is just beginning. N. L. Christensen attributed ecologists’ interest in history to four factors: (1) predictive models structured only on the natural world often account for

History Hidden in the Landscape  17

little of the variation in ecosystems; (2) historical impacts have a much longer lifespan than ecologists previously thought; (3) landscape-scale processes are nearly always influenced by past human activities; and (4) one can expect the influence of human impacts to increase in the future, so prediction will necessarily have to include human factors. In addition, the whole system is dynamic, varying on several scales, so that the effect of an action at one time may differ from that same action taken at another time. For example, abandoning an agricultural field during a period of dry years may have very different effects on future forest regeneration than abandoning the same kind of field in the same area during a period of wet years. Historical uniqueness, for example, the specific conditions at a given time, may modify the impact of the more easily observed variable of time.32 In addition, recent paleoecological research suggests that ecosystems, even those with minimal human impact, constantly change at a variety of scales, from decadal stand-level responses to such local disturbances as windstorms to centennial regional responses to changing climates. The search for a typical “climax” type of vegetation for a region is bound to fail because of the constant reshuffling of species under the influence of changing climates, species migrations, and microevolution.33 The human element serves to magnify the impacts of these influences. To be rigorous, historical ecology must establish clearly stated hypotheses to explain the impacts of past human activities and their consequences. In a way, the experiments have already been done, but, in contrast to controlled experiments in the field or laboratory, we know the consequences but not the experimental treatment or controls.34 Possible “treatments,” such as climate change or specific human activities, are posited, with the likely consequences of each. The full panoply of historical techniques can then be applied to either refute or support the hypothetical causes. A critical component of these studies is using multiple lines of evidence to test the proposed hypotheses.35 The next three chapters introduce the major sources of information used by historical ecologists, including some of the basic assumptions behind data use in various fields. I shall then show how ecologists, historians, and geographers have used these techniques to elucidate various aspects of human impact on the environment. I conclude with a discussion of how the historical perspective can help analyze current ecological problems as well as contribute to conservation.

2

Historical Records and Collections

The abundance and diversity of written accounts of the natural world attest to people’s abiding interest in the earth that sustains them. Travelers write letters describing the new lands they visit. Farmers keep track of their daily activities in diaries. Governments collect statistics about the land they govern. Such disparate sources as restaurant menus and the amount of wool needed for Viking ship sails may inadvertently reveal changes in fish stocks or land clearing. Natural history collections serve as extensive samples of the organisms that lived in the past, and published datasets used for scientific studies may provide specific details about past ecosystems and landscapes. All of these and more constitute evidence of past environments and the human actions and attitudes that affected them. Many historical documents and natural history collections have been digitized, and are available online, and many more are becoming available every day, decreasing the effort necessary to find many documents.1 For local insights into the past, however, visits to local libraries, archives, and historical societies still reward the effort. In any historical study of a literate period, these sources present unique information on human activities as well as provide the basic temporal framework for analyzing other kinds of data. Like any observations, however, each represents only a small portion of reality, seen through necessarily biased

18

Historical Records and Collections  19

eyes. They cannot, therefore, be used as direct statements of reality in the past but must be interpreted with awareness of potential bias.

Unique Values Written documents and collections offer uniquely specific temporal and spatial contexts for studying the history of human impact on the landscape.2 For example, many coastal salt marshes are completely obliterated today by structures and waste deposits. Dated historical maps can show how long they continued to function as salt marshes as well as their original sizes and locations.3 The proliferation of railroads in the mid- to late nineteenth century led not only to deforestation, as trees were cut for railroad ties and fuel, but also to more fires ignited by sparks from steam engines, affecting vegetation near the railroad lines. The locations of the lines and the politics of allocating public land to railroad companies to supply wood were important drivers of their impacts on the natural vegetation.4 Historical documents allow precise dating of events, which contributes to inferring cause and effect. For example, the arrival of rinderpest in Africa in 1887, a devastating disease of ruminant animals such as cattle and wildebeests, and its eradication from cattle in the 1960s explain massive changes in vegetation and elephant populations in the twentieth century.5 Historical records document economic and social systems that determine structures of cultural landscapes and their interactions with the natural environment. Using this information, the historical ecologist can frame hypotheses to interpret the environment in the context of past human drivers as well as changes in climate. These documents as well as other kinds of evidence can be used to test these hypotheses, not by experimentation, but rather by corroborating or refuting the hypotheses.

Evaluating the Sources Any historical document, whether a map, published data, or any other written source, must be subjected to a series of tests to determine its reliability as data. A first classification is whether a document is primary or secondary. A primary document is one written more or less contemporaneously by the person who made the observation or collected the data.6 Secondary sources are those that use primary sources to interpret a past event or condition of which the author has no firsthand experience; these are usually not as useful as evidence. This distinction is not always easy to make, however.

20  Questions and Clues

For example, an often-cited description of the forest composition in the New Netherlands (New Jersey and eastern New York) in the eighteenth century is repeated almost verbatim in several places as a description of local vegetation in sites many hundreds of kilometers apart. The description of each site is given, however, as if it were local; it was, rather, a set piece used to describe a forest.7 There is no doubt that the author had firsthand experience of the vegetation, but his descriptions are not useful for determining local forest composition. Often the rarity of such descriptions makes it tempting to regard them as valid without adequate analysis. It is perhaps better in these cases to have nothing. Scientific articles reporting results of research may be considered primary sources and are often used in meta-analyses. In order for these to be used as evidence of ecological conditions in the past, they must be treated in the same way as other, more literary primary sources. In comparing published data on past vegetation in several parks in New York City with current vegetation, Loeb found errors in the earlier data because they were grossly inconsistent with the current locations of trees. Although it is not usually possible to make such detailed comparisons, approaching datasets for meta-analyses as historical sources rather than as straightforward data is helpful in using them to best effect. One must ask why the studies were done, which affects coverage and specificity of the data collected and reported, and what biases may have influenced the results, in addition to establishing clear criteria for inclusion and exclusion of data.8 Primary sources appear on the surface to be accurate, usually objective original records of contemporary conditions, and within limits many are. The key to using them effectively is determining these limits and deciding what questions to put to them.9 At one extreme, one can be very critical of the sources, treating them in effect as a lawyer treats hostile witnesses.10 The other extreme is to treat them as accurate accounts of the truth. If the criteria used for evaluating them are made clear, a reader is better able to assess the interpretation. A series of observations such as those made by a census survey may reveal changes in the environment over time, from which one can infer causation. Changes in definitions, techniques, and even census takers can, however, make comparability difficult. In the United States Federal Census agricultural records, for example, the definition of a farm and of categories of land use within each farm changed over time. Even the reported area of political jurisdictions can change. In reports on forest area of New Jersey, for example, the area of the state was reported as 2,578,700 acres in 1899 and 2,553,100

Historical Records and Collections  21

Figure 2.1. Yield of wheat, 1815–1859 Liverpool Merchants’ Results. (Data from Healy and Jones 1962.)

acres in 1972, a 4 percent difference. Such adjustments can make assessing changes from one census period to the next problematical.11 Another example of complications in a series of records is the apparent decline in silica in the water of Lake Michigan at Chicago from 1926 to 1962, which has been often cited and even incorporated into textbooks to illustrate eutrophication of lake water. The decline appears when the data are analyzed over the entire period, but when they are partitioned into two sets, based on a change in analytical technique between 1948 and 1949, the major difference appears to be between the two sets of data. A change in technique that was designed to improve the quality of the data rendered them less useful for long-term studies. In another series of statistics, data for wheat yield in part of England from 1823 to 1857 suggest that yield increased from thirty to fifty bushels per acre. Survey techniques apparently inflated the yields, making them difficult to compare with modern estimates. In addition, it appears that most of the change occurred between 1838 and 1842, which corresponded with a change of estimator, so this may account for the apparent increase in yield (fig. 2.1).12 Outright fraud is also possible. For example, in 1755, the governor of North Carolina bemoaned the fact that some surveyors made up corner trees (trees found and marked at the corners of surveyed lots) without actually doing the surveying, taking advantage of the rectilinear design of the lots.13

22  Questions and Clues

Using trees noted in these surveys as a sample of contemporary vegetation would obviously not result in a reliable picture.

The Importance of Perceptions It is a tautology to say that written records are the products of the mental processes of the writer and are thus not simple, direct records of a past situation, but it bears emphasizing, especially for nonhistorians. Attitudes and perceptions color what is written. For example, ecologists and environmentalists know George Perkins Marsh primarily for his exposition of the depredations effected by humans on their environment over the centuries. They may be surprised to learn that in 1882 he wrote, “I am more than ever impressed with the superiority of the artificial [managed] forest, both in quantity and quality, as compared with that of the natural and spontaneous growth,” a statement consistent, however, with observations he made earlier in Man and Nature, including, “The sooner a natural wood is brought into the state of an artificially regulated one, the better it is for all the multiplied interests which depend on the wise administration of this branch of public economy.”14 As much as Marsh was a perceptive interpreter of human impact on the environment, he was also a man of his age who saw the natural world as a wealth of resources for the wise use of humankind. Similarly, we are inescapably of our own age, and our attitudes are deeply imbued with current values and perceptions. A few examples of the current diversity of perceptions, depending on professional or cultural biases, emphasize the role that attitudes play not only in our reactions to our environment but even in what we see when we look at it and record our impressions. On a field trip at a workshop on fire management of barrier islands, I visited an area of impenetrable palmetto scrub at Canaveral National Seashore (fig. 2.2). If I had visited this location on my own, I would have regarded the thick scrub primarily as a hindrance to walking off the trails. After I had attended a day and a half of lectures and discussions on fire ecology and management, however, this vegetation became “fuel load,” poised to carry highly destructive fires that would threaten human habitations. The crucial transformation was that I not only grasped the significance intellectually, I actually perceived the vegetation as fuel, a culturally conditioned response. A cultural bias may be illustrated by the observations of an English medievalist and his wife about a preserved forest in Princeton, New Jersey. Being accustomed to meticulously managed woods in England, my friends were quite distressed by how messy the woods were; they needed “tidying up” (fig.

Historical Records and Collections  23

Figure 2.2. Palmetto scrub at Canaveral National Seashore, Florida. (Photo by the author, 1984.)

2.3). They were referring not to human litter, which was not in evidence, but to natural litter, dead branches and fallen trees. To an American the site looks the way a preserved forest should look, with forest plant litter in varying states of decay forming a patchy cover on the forest floor. The combination of lack of neatness and waste of wood carries quite different connotations to others. Neither attitude is inherently right or wrong; they are merely culturally conditioned points of view, although the different methods of managing the forest would have varying impacts on nutrient cycling. Throughout history, however, attitudes have had consequences for how people have treated the environment, what they have studied, and how they have expressed their observations. A historical ecologist who relies on written sources is usually searching for different information than is a historian using the same sources, and he or she sees them from a different set of paradigms. Ecologists are usually interested in what the document reveals about the external world, what kinds of plants and animals there were, the weather, and other physical descriptions. For example, on 1 May 1732 a certain John Hayward in New Jersey stated that “he was approached . . . to become part of a band rounding up the many horses roaming loose in the woods at that time of year to put on weight for plowing to sell them in Maryland or elsewhere.”15 The economic historian may find evidence here of illegal trade in horses among the colonies. There is evidence of the lawlessness of this early frontier between plowed land and uncut forest. For the ecologist, there is evidence both of the progress of clear-

a

b Figure 2.3. Contrast the tangle of dead fallen branches at the Institute Woods, Princeton, New Jersey (a), with the neat, clear understory of a beech woods in Juniperhill Woods, England (b). (Photos by the author, spring 1995 and spring 1985.)

Historical Records and Collections  25

ing and of early grazing within the uncleared forest. What we cannot tell is whether this was an isolated incident. Without further corroboration, this can serve only as a tantalizing suggestion of conditions in the past.16 Both general and local contexts color what is written and how it is expressed, even in public documents. Censuses and other land records, for example, catalogue areas in different crops but generally cite no details of forest composition because the records were not ecological descriptions but estimates of agricultural potential. Probably the most complete inventory of land resources a millennium ago was the Domesday Book, an inventory the Norman conquerors made in Great Britain in 1086 CE. Assessors described woodlots, for example, by length and breadth, area in acres (though it is unclear how much an acre was in 1086), or even by how many swine they could support.17 This enumeration of property offers a glimpse into land cover but includes only major categories, such as fields, ponds, and woods, because its purpose was to regularize land tenure, not to describe land cover.18 The most detailed environmental information from the past describes economically or politically valuable aspects of the environment but only indirectly includes the kind of detail ecologists seek. Librarians are often unaware that their manuscript collections contain environmentally interesting information; I have often been told that some diary would contain nothing of interest, only to find that inadvertent or indirect comments were quite valuable. For example, an eighteenth-century diary commented on the presence of smoke in the air from “many meadows afire underground.”19 Combined with other evidence, this suggested that draining meadows (marshes) had allowed them to dry out so that the highly organic soil would burn; furthermore, it is a clue as to why the vegetation that regenerated after the drainage ditches were abandoned is apparently different from what was there before draining, as fires had changed the soil. By relating an understanding of ecological and physical processes to historical evidence one can often posit novel explanations for current systems.

Maps and Photographs Maps and photographs furnish graphic evidence of past land cover. Maps represent an interpretation of landscape features, often including topography as well as human constructions like roads, towns, and political boundaries. While a map may appear to be a simple rendering of the surface of the land in two dimensions, cartographers have a wide range of choices to make in drawing a map, which will emphasize some features while obscuring others. Some features may be out of scale with others, for example,

26  Questions and Clues

built structures may appear larger than life because they would appear too small to notice if they were drawn at the same scale as the base map. Definitions of land cover affect apparent patterns, such as contrast between woodland and prairie. Thus, in using maps, one must often apply even more careful critical analysis than would be used for a written document. Some useful kinds of maps for establishing land cover in the past include cadastral (relating to real estate boundaries) and military maps that were drawn to define properties or help plan military campaigns. Comparing historical maps with more recent aerial photographs or satellite imagery often gives insight into the processes that have shaped current landscape features, especially with careful attention to differences of scale. Aerial photographs and satellite imagery often distort topographic features and ground distances. Although they can be “rectified” to deal with this to some extent, the problem of displaying a three-dimensional surface on a two-dimensional map always remains. Some military maps that are several centuries old are highly accurate in describing topography, depending on how carefully the mapmaker could survey the area. A map made in 1777 by the British surveyor of the Saratoga battlefield in New York, for example, can be used to trace stream locations and the locations of open fields near the British headquarters. The fields even have depictions indicating whether they were recently cleared, using an overlay of cut stumps. However, the map is decreasingly useful the greater the distance from the headquarters—that is, as the danger of colonial snipers increased (fig. 2.4).20 Sixteenth- and seventeenth-century maps of North America depict topography, rivers, and villages near the coast but are often quite vague inland. Mountains described by Native Americans may be placed haphazardly, and mythical lakes and rivers abound. In the late nineteenth century, however, carefully surveyed maps include forest cover that corresponds remarkably well with current vegetation. As with other documents, however, it is essential to pay attention to definitions of terms: for example, the term forest may mean uncultivated brushlands in the United States but serves as a legal rather than vegetational description in Great Britain. As long as one bears in mind that maps include details relevant to their purpose and that accuracy is often not uniform, they can be used effectively to reconstruct a past landscape.21 Photographs taken at one location at two or more different times often record dramatic and detailed changes in plant cover. When the Mexican and American governments constructed monuments along their common border in the late nineteenth century, for example, surveyors took an archival photograph of each monument. Photographs of these same markers taken almost a

Historical Records and Collections  27

Figure 2.4. Military map of 1777, Saratoga battlefield, Saratoga County, New York. Note that the ravines and hills are delimited by hatch marks to show steep slopes. Cleared fields and forest cover are indicated by other symbols. (Wilkinson 1777, Geography and Map Division, Library of Congress.)

century later document vegetational change in their vicinity as well as differences across the border caused by different land use. A comparison of other photographs of the American West taken in the nineteenth century with the same places in the late twentieth century sheds light not only on land use and vegetational changes but also on the intentions of the earlier photographers. Many nineteenth-century photographic sites were carefully chosen for their dramatic landscapes or were photographed in ways that enhanced dramatic effects to impress people of the eastern United States with the western landscape.22 Thus comparisons of photographs over time can indicate changes in vegetation, but they, like other historical documents, must be interpreted in context, not just as objective data. A major rephotographic project in China, comparing pictures taken around 1900 CE with those from the early 2000s, documents apparently warming climate, indicated by earlier planting dates for rice and flowering of herbaceous plants; recovery after earthquake-triggered landslides; deterioration of grasslands due to overgrazing; and many other changes. Rephotographic studies can also be used to show the impacts of urbanization and previous

28  Questions and Clues

undocumented land use.23 There are many such collections of carefully documented photographs around the world that await further study. Aerial photography, particularly vertical aerial photography, lends itself even more readily to interpretation of landscape change. Such photographs, which date at the earliest from the 1920s, use technology developed during the First World War. Comparisons over time indicate patterns that suggest further ground-based study of land cover. In figure 2.5, most field patterns and successional conifer stands visible in 1937 (dark in the figure) have disappeared in the deciduous forest by 2016, and the eighteenth-century property boundary is also no longer apparent. Since the late 1960s, satellite imagery in a variety of wavelengths has also been available, so changes in land cover can be charted in even more detail.24 These images supply ample data to interpret short-term changes in landscape patterns at a variety of spatial scales. New analytical techniques are providing ways to improve the interpretation of change over time using historical maps and photographs. In Sweden, Cousins used transition matrices to adjust historical maps so that the features on them could be compared with those on more recent aerial photographs. This allowed her to conclude that while the relative proportions of arable fields and seminatural grassland in her study area have stayed about the same since the seventeenth century, there have been dynamic changes in where the semi­ natural grasslands have been over time, so that only 8 percent of them are more than three hundred years old.25 In Australia, Lunt and Spooner integrated historical cadastral maps and current aerial photographs to highlight pattern and process in fragmented terrestrial landscapes, testing the application of theories of island biogeography to terrestrial patterns. They found that the processes that have shaped the terrestrial systems are quite different from those that shape island systems, at least in part because the locations of the fragments are not random, and historical patterns of land use have affected structure and composition of the fragments.26

Other Kinds of Documents Governments collect a wide range of descriptive statistics.27 Many of these document the kinds and amounts of natural and agricultural resources available for exploitation. Historical land surveys in North America often include natural features and trees marking corners, referred to as “witness trees,” as well as the locations of roads, waterways, and buildings. Wills and estate inventories also describe properties. Tax rolls contain valuation of

Historical Records and Collections  29

property and crops. All of these kinds of documents provide what a scientist using historical documentation is most comfortable with: numbers, categories, and locations. For reconstructing land cover and use at specific times and changes over time and space they are invaluable. The data may record, albeit indirectly, locations and rates of forest clearance and use of other natural resources as well as field abandonment. For example, an estate inventory in New York made on 21 February 1823 included “15 Sheep with 2 lambs,” indicating that lambing had just started by 21 February, “$25 worth of Grain in the Ground,” “$4 worth of Cyder and Casks,” evidence of an orchard, and “1 Lumber Sleigh” and “1 lumber Chain and ax,” indicating the existence of uncleared woods, probably on the property.28 A plethora of other kinds of documents adds context and detail to interpretations. Lists of flora and fauna document occurrences of species in the past. Correspondence and diaries record individual perceptions of landscape features, often explaining more cryptic information available from formal documents. Whereas an agricultural census return may record the number of cords of wood cut on a farm, a farm diary may include details of how the farmer used the wood and when and how he cut it. For example, cutting small trees for “hoop-poles” (small, straight stems used for making barrels) had a different effect on forest vegetation than cutting larger trees for fencing.29 A preference for twigs as winter fodder similarly implies different impact and management than would leaf collecting.30 Some descriptions were written for blatantly commercial reasons. Advertisements of property for sale or other descriptions designed to attract people are obvious examples, although some sound quite convincing. Landscape paintings and other artwork also contain clues to past landscapes, though these may be distorted by the artistic perspective. For all of these sources, multiple lines of evidence are critical to using them to reconstruct past landscapes. For example, a book published in 1833 states, “Much of the greater part of the country [Frederick County, Virginia] . . . was one vast prairie,” giving as evidence that “there are several aged individuals now living, who recollect when there were large bodies of land . . . barren of timber. The barren land is now covered with the best of forest trees.”31 Is this evidence that much of this region, the Shenandoah Valley, was treeless in the early eighteenth century? I tested this secondary source by considering a compilation of eighteenth-century land surveys. There was no indication in the surveyors’ reports of extensive grasslands, and trees marked almost all property corners. Meadows were found only along streams. Combining this with the natural growth of trees in the valley, I concluded that the oral history here was most likely an inaccurate description.32 On the other hand,

a Figure 2.5. 1937 vertical aerial photograph (a) and 2016 satellite image (b) of the same location in northern Virginia. Note the loss of fields to the east of the creek and replacement of conifers (very dark trees, probably red cedar or Virginia pine) to the right of the star by deciduous forest. (2.5a: Photo from the Carto-

b graphic and Architectural Branch of the National Archives; 2.5b: Photo from the Loudoun County, VA Office of Mapping and Geographic Information, flown by the Sanborn Mapping Company, Inc., Colorado Springs, Colorado.)

32  Questions and Clues

a variety of oral history sources have confirmed both an earthquake and a tsunami off the Pacific coast of North America and in the Pacific Ocean in 1700, the date also established in Japanese records of the tsunami. The oral histories also indicate that earthquakes and tsunamis were recurrent events.33 Laws may also suggest land cover, but they, too, can be difficult to interpret. They may regulate activities to protect resources and/or property rights, so it is important to be aware of why a law was written. It may state the opposite of the reality, that is, how conditions ought to be. Thus, regulations on cutting white pine or other timber in various American colonies reflected not a scarcity of timber but rather the desire on the part of the British government to prevent individuals from poaching timber. Closed hunting seasons for some species in twelve of the thirteen American colonies by 1776, however, reflected both a diminished herd and a perception that deer hunting should be protected.34 Closed seasons did not indicate opposition by governments to hunting, but quite the contrary, a commitment to revive the resource for further hunting. In addition, just because an activity was not regulated does not mean that it was permitted; it may mean that not doing it was taken for granted.35 One cannot always rely on laws to state the obvious.

Natural Science Collections For centuries, people have collected specimens of living organisms as curios, for scientific documentation, and for exhibits. I have included them under historical documents because they must have written documentation to identify time and place of collection as well as collector. This documentation and the reasons for the collection should be treated as written documents in terms of potential biases. For example, a collector planning on exhibiting the collection would select only ideal specimens. The collection would not reveal anything about possible deformities, so all perfect specimens would not indicate that there was no disease or insect infestation in the population. Current digitization of herbaria and other collections is making them ever more available for comparative study, and techniques for extracting DNA and doing other chemical analyses open a large range of options for using historical collections in ecological studies.36 Ample evidence indicates that the climate has warmed from at least 1850 CE to the present, at varying rates in different places. Whether these changes have had impacts on species distributions and phenology can be tested using natural history collections such as herbarium specimens. By comparing spe-

Historical Records and Collections  33

cies distributions in a county in Pennsylvania, Bertin showed that during the second half of the twentieth century, more northern species had been lost at the southern end of their ranges than southern species had been gained, providing evidence that species losses caused by climate change may be more rapid than species gains. These kinds of detailed records can help strengthen statistical tests for the effects of climate change, which otherwise rely on short-term datasets.37 While useful for documenting species presence and distribution, the specimens themselves offer additional information for documenting change over time. Increased methyl mercury in museum specimens of ivory gulls (­Pagophila eburnea) combined with stable isotope analyses from 1877 to the present showed that while diet was constant, mercury increased forty-five times. Stable isotope analysis of preserved fish from the Rio Grande River from 1930 to 2000 documented significant loss of trophic complexity in the river over this time period, which would not have been apparent from current observations of the biota of the river.38 Observations such as these document the importance of maintaining and curating these collections.39 These written records afford abundant glimpses into the past. Because the questions that historical ecologists ask today are rarely those asked by people in the past, there is no one compendium of sources. Historians are always faced with fragmentary evidence and must assess the representativeness of the evidence they find. Some records have been deemed important enough to be edited, printed, or scanned, and new methods are being developed to search and use these efficiently.40 Unpublished records still survive in a variety of collections: libraries, historical societies, clerk’s offices and other depositories of public documents, business archives, and sometimes in attics and old trunks. These documents exhibit a wide variety of detail, reliability, and availability. The information gleaned from them must, however, be placed in the context of a larger question in order to be anything more than a chronicle. In such a context they provide explanations for some changes in the landscape, changing use over time, and a chronological framework for other research findings.41 With a little imagination one can usually mine more information from them than at first appears possible, and there remains a wealth of information that has yet to be explored for ecological insights into past environmental conditions and processes.

3

Field Studies: Bringing Historical Records Down to Earth

The evidence gleaned from written sources comes to life for ecologists when they apply it to specific landscapes and their component ecosystems. Documents guide historical ecologists to sites where they can search for evidence of past activities described in the documents or collect data to test hypotheses about the impacts of the past on the present. Often, however, there are no written records of a human activity, or the written documents include only vague locations or descriptions. With field evidence, ecologists can directly tie the past to the present structure and functioning of landscapes and ecosystems. Because many written records are not spatially precise or detailed, making the connection between documents and landscapes often involves painstaking fieldwork to decipher the relationship between the physical remains of past land use and historically documented activities such as logging or farming. The use of Global Positioning System (GPS) data in the field can guide an ecologist to the sites of these activities.1 Climate, microclimate, geology, soils, and the species pool determine the possible ecosystems of an area, including species composition and structure. In addition, these factors affect human activities. For example, there are most likely no remaining examples of undisturbed vegetation on excellent agricul-

34

Field Studies  35

tural soils. Farms have taken the place of forest and prairies in most places, and even woodlots have served for cordwood, lumber, fencing, and casual grazing. On the other hand, ecosystems in remote mountainous parts of the world may be relatively untouched by people, especially if they have no valuable mineral deposits.2 Abiotic features and location strongly influence both potential natural vegetation and human activities; the two are inextricably linked.

Relating Documentary to Field Evidence Although some patterns of vegetation, such as medieval coppice woodlands (stands of multiple-stemmed trees that have originated by sprouting after repeated cutting), obviously relate to past human activities, more subtle patterns may also owe their existence to past human activity. For example, in 1987, one roughly rectangular patch of oak-tulip poplar (Liriodendron tulipifera) forest in eastern Pennsylvania, about thirteen hectares, had more tulip poplar and black birch (Betula lenta) than the surrounding forest. In addition, there were many dead red cedars (Juniperus virginiana) in this area. Because red cedar rarely grows under a forest canopy in this region, it is a strong indicator of forest regeneration after field abandonment.3 Written records describe an active iron furnace in this area from 1771 to 1883. In an aerial photograph from 1937, most of the current tulip poplar forests were open fields. None of the current oak forests was an open field in 1937, but remains of old logging roads and the leveled circles of old charcoal hearths indicated that these forests had been heavily cut to make charcoal. This suggested that the stand had originated as an abandoned field well before 1937, while the nearby forest was still being cut for charcoal. An abandoned road and foundation supported this inference.4 Any study of the effect of cutting for charcoal on subsequent forest development must therefore exclude this stand. Documentary evidence thus supplied a composite picture of the history of the forest in general, but a field survey considering historical land use was necessary to differentiate this stand with a singular history. Some historical land surveys contain sufficient information to infer probable vegetation and even quantitative species composition of sites in the past. Comparison of these with the present, coupled with indications of intervening activities, suggests the consequences of the intervening land use and climate change on current plant and animal communities as well as on nutrient cycling and other ecosystem properties. Field studies search for evidence of such past features and for their resultant effects through a variety of techniques.

36  Questions and Clues

Figure 3.1. Medieval ridge and furrow landscape that has been preserved by pasturing, Leicestershire, England. The reverse S-curve of some of the pattern is characteristic of much ridge and furrow in the Midlands of England. (Cambridge University Collection of Aerial Photography, © copyright reserved.)

Field Techniques for Historical Reconstruction There are many field techniques that take research beyond what can be found in written documents. Clues to the past are hidden in such subtle (and not so subtle) features as topographic modifications, soils, and tree trunks. Many anthropogenic features, like old stone walls, banks and ditches, abandoned roads or clearings, and rows of old hedgerow trees embedded in younger forest, are obvious in a field or forest. Medieval plow marks have survived to the present in some places, forming distinctive “ridge and furrow” microtopography (fig. 3.1). This pattern indicates not only plowing in the past but also the lack of more modern plowing, since that would obliterate the old marks. Analysis of these features can generate a detailed reconstruction of past patterns of fields, woodlots, roads, and buildings.5 These patterns appear in aerial photographs and in other remotely sensed images at many scales, revealing past activities ranging from large-scale irri-

Field Studies  37

gation projects to individual farm sites and woodlots. Because some patterns are easier to discern at a distance, remote sensing often bridges the gap between written documents and actual field reconnaissance. It may be difficult to locate a specific site on the ground, but if it has left a distinctive mark visible in aerial photography or other remote imaging, the location can be found. The patterns on the landscape can then be analyzed using geographic information systems (GIS) to discern both ancient landscape systems and change over time.6 An additional remote imaging technique is lidar, which uses a laser beam to reveal images of surface features. While aerial photography detects surface features best in nonforested landscapes, lidar can reveal features hidden beneath a forest canopy. It also provides imagery in areas that are difficult to access because they are remote or have very thick vegetation. Finding features such as old field boundaries or buildings hidden beneath presumed ancient forests using lidar has opened new vistas for interpretation of forest patterns in both tropical and temperate forests where there is evidence of ancient abandoned cultural sites.7 Soils, too, reflect past activities. In some soils, cutting trees initiates severe erosion, which may leave steep-sided gullies that persist long after agriculture or logging has ceased. Sheet erosion removes surface soil, sometimes to the extent of washing off all the soil and leaving exposed bedrock. Plowing homogenizes the surface horizons of soil, destroying upper-level zonation. In floodplains, buried soils and sedimentary sequences show patterns of deposition and erosion related to human activities. In the Crau, a dry steppe in Mediterranean France, plowing centuries ago brought the underlying conglomerate bedrock to the surface, which continues to affect soil processes (fig. 3.2). There is no other evidence for plowing at this site.8 Stable isotope analysis of organic remains in soils and fossils allows reconstruction of many aspects of communities and climate. In Patagonia, analysis of fur seal remains in archeological sites indicates that decimation of hunting starting in the nineteenth century did not change the proportions of primary producers, kelp, phytoplankton, and bacteria in their diets.9 Stable isotopes of oxygen can indicate climate, while those of carbon relate to photosynthetic pathways. Finding stable isotopes of carbon in soils in Guatemala indicated the local areas where maize had been cultivated.10 Forests that have regenerated on old maize fields could then be compared with those that had never been cultivated. Fires, which people may have ignited, often leave distinct layers of charcoal in soils of grasslands and tropical rain forests as well as temperate forests. While pollen is usually not preserved in terrestrial soils, charcoal and

38  Questions and Clues

Figure 3.2. Bedrock conglomerate brought to the surface by plowing in the nineteenth century. The soils in this plowed area are still affected by this prior activity. A combination of field study and analysis of historical documents and local archaeology has permitted this interpretation of the pattern. (Photo by Christine Römermann, IMBE.)

phytoliths (silica bodies formed within plant cells), which can be identified often to genus, frequently are (fig. 3.3). As charcoal is made of carbon, larger pieces can be dated by carbon 14, and so can tell a story of change over time. Charcoal studies are providing novel interpretations of past environments in a wide range of habitats, from Brazilian savannas and Mediterranean steppes to European forests.11 Agriculture often raises the pH of soil by bringing base-rich minerals from deep soil horizons as well as by the addition of lime, while a regenerating forest usually lowers pH. Conversely, deforestation in the uplands of England millennia ago probably contributed to acidification of the soil as heaths replaced the forests. These processes may then have indirect effects on the pH of local streams and lakes. The existence of any of these conditions indicates past land use and directly affects current ecosystems as well.12 Archeological studies using physical remains of specifically human activities indicate how people lived in the past and to some extent how they used natural resources. For example, archeologists have used assemblages of house sites in the Maya region of Guatemala to estimate changing populations there, interpreting a slow increase as long ago as 3000 BCE to a peak of perhaps two hundred persons per square kilometer around 800 CE. The population then declined precipitously in the ninth century and did not ex-

Field Studies  39

Figure 3.3. Charcoal fragments (dark entities), mainly from Quercus (oak) and Fagus sylvatica (beech) (×10), and other remains in a sample from shallow, dry grasslands soils in the Seine Valley, Upper Normandy, France. The charcoal fragments have yielded dates ranging from ca. 4000 BCE to recent times. (Michel Thinon, IMBE.)

ceed its former density until recent times. These population changes can be related to changing conditions of a local lake.13 Radiocarbon-dated organic remains coupled with population models provide even more detail and precision for population change, though very large sample sizes are required for accuracy.14 Historic as well as prehistoric archaeology permits reconstruction of past environments and uses. For example, in northern Australia, extensive stockyards are common in an area of few trees. The numbers and sizes of the logs in these structures as well as historical documents indicate that the region was much more heavily forested in the past. Once cut, the trees have been unable to recover, but their remains testify to their former existence.15 The plants growing on a site also hint at its history. An even-aged stand of trees generally originated after some major disturbance, whether natural or anthropogenic. Other indicators may suggest the nature of the disturbance, for example, natural disturbances usually do not have rectilinear outlines, whereas artificial ones like agricultural fields often do. Changes in age and composition across a straight line thus suggest a human element. Some plantations are obvious, but others that have utilized native species can be less so. Field data may indicate past harvesting when a species is lacking on sites where it is expected and the sites are near locations of past human activities.16

40  Questions and Clues

Fig. 3.4. Hazel coppice, Wallis Wood, Surrey, England. Note the regular spacing of the sprouting stools, which were cut in the previous year. (Photo by the author, 1985.)

Another clue to past use lies in the structure of individual plants, for example, in coppices. These usually consist of one species with more or less regularly spaced clumps of stems (fig. 3.4). Scattered among the coppice stems are often larger individuals grown as “standards” to produce larger wood. The basic structure of the stand may persist long after the coppice management has ceased.17 Even in North America, where coppicing was usually practiced in a rather casual way, there are stands with many multiple-stemmed trees: that is, they may have been cut in the past and have regenerated as sprouts rather than as seedlings. Such an origin for the trees in a forest imposes a distinct pattern of age and size classes and species composition, known as sprout hardwoods, characterized by extensive stands of prolifically sprouting taxa such as American chestnut (Castanea dentata) and oaks. Flora also reflects past land use. In eastern Pennsylvania, for example, fortyyear-old tulip poplar stands that regenerated on an abandoned agricultural field had an herbaceous layer rich in such non-native weeds as field garlic (Allium vineale) and Japanese honeysuckle (Lonicera japonica), whereas these species were not present in the herbaceous flora of tulip poplar patches of approximately similar age and size in sites that were in all probability never used as fields. In Bavaria, several herbaceous species were associated with known ancient or recent grasslands (as established from historical doc-

Field Studies  41

uments and from pollen and charcoal found in soil), suggesting that they could be used as indications of the history of grasslands where the history is not known from documents.18 In England, old field edges can often be located by the presence or absence of woodland species that spread only very slowly into nearby regenerated woods. Others, such as sanicle (Sanicula europea), indicate old fields.19 Such observations made on sites with a known history can then be used to help identify past land use on other sites and to study the characteristics of species’ responses to disturbances in the past.20 In a study of the relation of understory flora to origin and type of woods in New England, Whitney and Foster found that the origin of a stand as oldgrowth or old-field succession influenced the distribution of more species than did the relative numbers of conifers and broadleaved trees present in the stand (table 3.1). The interpretation of such differences depends on thorough knowledge of the local flora and site histories. These and many other indicators of land use that are fairly easy to discern in the field present the historical ecologist with a static picture, a snapshot of the present consequences of a past inferred activity. There are several ways to tease out a more dynamic, moving picture of change over time. These involve finding traces on the landscape of a succession of past events, studying time series in the field, and modeling the changes and their possible causes. As woody plants die, their remains are often buried by additional litter before they decompose completely, especially in cool moist climates and acid soils. Trees that are toppled by storms produce distinctive mound and pit patterns where the roots were pulled from the soil. Erosion and decay slowly smooth these mounds, but their traces may remain for centuries, especially if other trees fall on them. Farming also smooths these remains, so that an old agricultural field that has reverted to forest leaves a surface unscarred by these mounds, pits, and ridges. The consequences of this kind of microtopography for regeneration of plants and distribution of animals indicate ways a younger forest may be distinct from an older one even after many decades of growth. Even where stems have broken off rather than pulling up their roots, the tangle of superimposed stems can be dissected to trace the history of tree deaths. This very painstaking procedure yields a detailed picture of changing species composition and disturbance events, which may then be related to patterns of natural and human disturbance.21 Although studies such as this can be done only on small areas and their results are site-specific, they afford unique, detailed glimpses into temporal patterns of change and the importance of disturbances to the structure of the forest vegetation.

Table 3.1. Comparison of common understory species in different woodland types

Species/habitat

Primary woodlands

Secondary woodlands

Conifer

Conifer

Broadleaf

Species with generally >10% occurrence in primary woodlands, 10% occurrence in secondary woodlands, 5 >2 n/a >5

3 0 n/a n/a n/a n/a

Data sources: Callmander et al. 2011; Goodman and Bensted 2005; http://hbs bishopmuseum.org/hispp.htm; http://issg.org/database/.

At least twenty-five species of animals, including at least fifteen species of lemurs, all larger than their surviving near relatives, became extinct on Madagascar after the arrival of people between four thousand to eight thousand years ago. Bird species, especially the large, flightless ones, were decimated after people arrived on the Hawai’ian Islands and also on Madagascar.47 Elimination of the large birds and mammals suggests that hunting played a major role in their demise but it appears that, as with the extinction of the late Pleistocene megafauna, the causes of the extinctions were more complex. Recent research indicates that people were present on Madagascar at least two thousand years before the major extinctions.48 In both Hawai’i and Madagascar, people cleared vegetation to make farms, using fire in both regions. They also brought domesticated animals and cultivated plants, along with unintentional passengers like rats and weeds, and hunted many native animals. Theories of changed habitat in Madagascar have long blamed people for the spread of fire-prone vegetation in a land

194  Contributions of Historical Ecology

lacking fire before human colonization. Recent sedimentologic evidence and the existence of endemic genera of grasses characteristic of burned grasslands and savannas are more consistent with a much longer, prehuman history of fire and grasslands on the island. Many species of animals that are now extinct survived for at least two millennia after people arrived on the island. While human activities led to a major expansion of fire-adapted habitats, they did not initiate them. Drying climate also likely contributed to the spread of drought-tolerant vegetation at that time. These factors point to the inadequacy of a simplistic explanation for extinctions. Focusing on only one factor misses the synergistic effects of multiple causation.49 Some physiological characteristics of species reflect past conditions, not the present. For example, some species of plants in Madagascar have defense systems that would be effective only against birds that are now extinct.50 In other words, one cannot explain the current patterns of activity or characteristics without reference to fairly recent changes brought about by people; to try to explain them by current physiology and ecosystem dynamics may lead to erroneous interpretations. The extinction of large lemurs may have had as yet undetermined impacts on the dispersal of plant species as well. Losses may not always be evident; forests may appear to be fully functioning in terms of tree species diversity but may have suffered changes in their fauna or soil that limit their potential for continued survival. While it is clear that in Hawai’i Polynesians planted extensive fields, caused the extinction of large numbers of birds, and increased the incidence of fire, the process that accounts for the correlations of loss of forest, increased charcoal, and extinctions with human impact is not as clear. Agriculture spread first on the more fertile and well-watered areas on the western, windward parts of the islands, and only centuries later spread onto the dryer uplands and leeward sides. Political changes apparently were critical for this process and the ecological changes that accompanied it.51 While agricultural fields replaced forest, it appears that at least in some areas, the loss of forest was hastened by the arrival of Polynesian rats (Rattus exulans), which likely eliminated trees before the spread of agriculture or loss of large birds. These herbivorous rats can devour large numbers of seeds before birds can get them, and may also have contributed to the loss of the birds that depended on the seeds for food. The amount of fire on the landscape seems to have been increased by herbaceous growth after loss of the trees, rather than as a consequence of agricultural clearing.52 The eighteenth-century colonists introduced more domesticated animals and plants, more inadvertently carried species, and more intensive and extensive use of the land. They converted almost all of the lowlands to such

Diversity and Species Extinctions  195

agricultural uses as monocultures of pineapples and bananas (see fig. 1.2). Much larger fields replaced indigenous agriculture in the lowlands, and large areas of uplands were converted to pastures. On the island of Hawai’i, such introduced plants as Opuntia and non-native grasses overran much of the northern regions, which lie in the rain shadow of the major volcanic peaks. Goats roamed everywhere, destroying reproduction of such rare species as silversword (Argyroxiphium sandwicense). The dynamics of the systems were, however, complex. A successful effort to remove the goats from Volcanoes National Park reduced the grazing pressure but allowed grasses, a ready supply of fuel for large fires, to spread. Uncontrollable fires in these grasses have destroyed some persisting stands of Metrosideros (see fig. 6.4). The spread of the non-native, nitrogen-fixing tree Myrica faya is changing the nutrient relations of primary succession on lava flows.53 The consequences of the human occupation of Madagascar and Hawai’i have been dramatic; many endemic species have been lost and replaced by non-native species in the last two millennia. Additionally, patterns of land use over time on both islands have reflected not only the natural resources but culture, technology, and population change, so the concentration of land clearing has varied in location and intensity over time. The patterns of human disturbances in these relatively small, isolated landscapes may throw into relief the consequences of these disturbances in ways that are not obvious in larger landscapes with a longer history of human impact. Studying them may help us understand not only their distinctive systems but also, by extrapolation, potential impacts in larger areas.

Conclusion Ecosystems that exist today are a single node in the continuing evolution and migration of species. They have been shaped not only by the prehuman processes of evolution, migration, and climate, but also by such impacts as human-caused extinctions and modification of habitats. In most of the world, ecosystems contain novel associations of species brought together through human agency in habitats that have been altered by human activities. Because of this, the maintenance of native species diversity calls for a major input of effort (energy) to counteract human-dispersed species and modified habitats. Regardless, the historical record shows that communities have constantly changed over all time scales from evolutionary and geologic down to seasonal. Human-induced changes are superimposed on those forced by climate, plate tectonics, evolution, and geomorphology. Current ecosystems carry with

196  Contributions of Historical Ecology

them the baggage of human-induced modifications over the past millennia. Assuming otherwise—that is, assuming that they are in equilibrium with conditions that can be measured now—will almost surely lead to erroneous conclusions about the relation between current conditions and processes that drive the systems. Evolutionary studies of the patterns and processes that govern biodiversity usually include only native species’ lineages that are assumed not to have been altered by people.54 For studies of the importance of biodiversity for ecosystem function, almost all systems that can be studied have experienced some alteration by people, however. These alterations have included changes in fire frequency, season, and intensities; introduced species; selective harvest and manipulation of the landscape for agriculture; and changes in technology and cultural systems. For every site, there is an overlay of these impacts, which have usually changed over time. Some of the most diverse systems are maintained by cultural activities, and date from millennia ago. In others, diversity may be attributed to human activities, but may actually predate these. Relict interactions among species may be obscure if the forces that led to their evolution have changed. As the examples discussed here show, errors of interpretation can flow from misinterpretation of the history of a site.

11

Biospheric Sustainability in a Changing World

Can the earth continue to support life as we know it for the foreseeable future? Sustainability often implies, either directly or indirectly, some form of use that can be continued indefinitely into the future, usually for human benefit. Ecologically, however, sustainability implies the continuation of ecosystems independently of human requirements. In the twentieth century and before, the goal of sustainability was often taken to be a system that resembles what exists in the present or sometime in a prelapsarian past. Toward the end of the twentieth century and into the twenty-first, maintenance of processes has often taken precedence over specific structure or composition, in recognition of the ubiquity of shifts in driving forces such as climate.1 The global environment is currently changing at an unprecedented rate. Burning fossil fuels has increased CO2 concentrations above any seen in at least eight hundred thousand years, altering climate and acidifying the oceans; production of nitrogen fertilizers from N2 in the air has altered global nitrogen cycles; and species extinction rates are heading for levels not seen in millions of years. In addition, these global changes are superimposed on, and to some extent caused by, other, more ancient changes wrought by people, such as altering fire regimes, transporting species across oceans, massive use of natural resources, replacing forests and prairies with agricultural crops, and generally divorcing many human activities from natural cycles and processes.

197

198  Contributions of Historical Ecology

These past changes require that we include in studies to understand how to maintain an ecosystem sustainably not only processes that can be discerned in systems today, but also those that have been affecting systems over the last centuries or millennia.2 Conscious concern that humans disrupt natural environments is not new. In 200 CE, for example, Tertullian, a Carthaginian Christian writer, observed that “our numbers are burdensome to the world, which can hardly supply us from its natural elements.”3 People had explored everywhere and exploited everything. Wilderness areas had become estates and forests cultivated fields. Swamps had been drained, deserts planted, and domestic herds had replaced wild beasts. Nature was failing to provide its usual sustenance. Depending on the translation, Tertullian described the natural world that had existed as “famous wilderness areas of the past” or as “once dreary and dangerous wastes,” an ambiguity still with us today.4 The theme of overpopulation recurred sporadically through the centuries, but the dire predictions were fulfilled only locally. After decimation by major wars, famines, or diseases, population usually rebounded to even higher levels, spurred by new techniques of production. Immigration to lightly settled regions relieved local population pressure, so that the real limits for global population did not ever appear to be reached. Some people mourned the loss of “natural” beauty where human populations became dense, yet most applauded the taming of wilderness, putting it to human use. Early apprehension about sustainability focused on resources, especially food and forest products, as do most goals on the United Nations’ Sustainability Agenda; this is not a new concern, though global perspectives are novel.5 As populations grew, the demand on agriculture to produce adequate supplies of food and fiber and on forestry to produce adequate wood products grew apace. Even where agriculture produced surpluses, wind and water erosion threatened future productivity in many areas. People devised a wide range of techniques to perpetuate the productivity of agriculture, constructing terraces to hold soil and using fertilizers, such as human and livestock manure and other waste, to replenish nutrients removed by crops or eroded away.6 Shifting agriculture relied on soil processes to return nutrients to depleted soil. In some areas of the tropics, and possibly even northern in Europe, soil amendments created entirely new kinds of soils, terra preta, by the addition of a kind of charcoal, which retained both nutrients and moisture. In dry, windy areas, windbreaks helped to protect the earth from wind erosion.7 In the mid- to late nineteenth century, deriving energy from burning fossil fuels led to major trade in materials used for fertilizer, for example, bones,

Biospheric Sustainability  199

guano, and oilseed cake, fed to livestock mainly for its value in producing excellent manure. About the same time, the nascent chemical industry introduced the first major manufactured fertilizer, superphosphate. Bigger and stronger equipment made it possible to cultivate soils that had previously been too heavy to plow, such as those found in prairies.8 With chemical fertilizers and pesticides, it became possible to increase productivity quickly and greatly. These, combined with technological advances in pest and weed control, new crop varieties, and larger equipment, led to more intensive and extensive use of land, with major increases in productivity. Ever deeper wells mined fossil water for human needs. Historically, native habitats had survived in many landscapes, mostly in marginal areas that were not intensively cultivated, or in hay fields or seasonal pastures. Intense cultivation, heavy use of pesticides and fertilizers, and new pasture grasses contributed to loss of even these habitats. Heavy demands for wood products led to the planting and managing of trees as well.9 Both coppices and plantations of fast-growing, often non-native species, especially Eucalyptus and conifers, furnished a constant supply of wood, although they, too, replaced naturally reproducing forests. Space exploration inspired quite a different concept of sustainability. People generally had accepted the concept, based on theory and exploration, that the earth was finite, but this concept crystallized when the Apollo 8 astronauts sent back pictures of the earth from the moon.10 We became conscious that we really are like Antoine de Saint-Exupéry’s Little Prince, tending our planet, which will no longer support us if we do not keep the destructive elements under control. Even more, we began to realize that we have the ability to alter the entire planet, not just individual parts of it, and that altering one bit will very likely have indirect impacts on many others. The earth as we know it is the result of billions of years of change. Not only the species that constitute the biosphere but also the physical life-­support systems of earth, water, and atmosphere have developed through eons of evolution. The rapid changes that we see today appear to many to be retrograde, and the possibility of natural adaptation, migration, or evolution to mitigate them appears remote. The message of sustainability of the biosphere is that it is up to humans to redirect their energies to learn to live within the limits imposed by current tolerances of the living and nonliving earth and actively to manage ecosystems where necessary to keep them from deleterious alterations. In the terminology of economics, use of the environment should not deplete its capital.11 That such self-determination might not be possible in the face of an increasing population was suggested by Thomas Malthus in the late eighteenth century when he observed that whereas population increases

200  Contributions of Historical Ecology

at a geometric rate, productivity increases only at an arithmetic rate, so will inevitably be overwhelmed—but the limits have been repeatedly stretched. Sustainability thus has meaning mainly in terms of its implications and expectations for the future. If sustainability means maintaining the requisite conditions for life to continue more or less as we know it, it implies a dynamic system that changes within certain boundaries, within which fluctuations are damped to avoid total destruction. Sustainability cannot mean maintaining some status quo ante, for example, some hypothetical system that may have existed in the past in a world without people. It also cannot mean maintaining a static system, because climate and culture will continue to change, as they have changed in the past. Rates of both kinds of changes are accelerating, however. Projecting current processes and patterns to the future is instructive but can represent only a small portion of the potential consequences of altered conditions, because the biosphere, and certainly the biosphere plus people, is an open system, one that will change unpredictably.12 The concept of sustainability thus suffers from a loss of specificity when it is removed from the realm of sustaining specific resources or ecosystems. Sustaining ecosystem services, for example, cycling carbon dioxide and nutrients and maintaining hydrologic regimes or biodiversity, are laudable goals, but until we know more about how these function, including their regional and temporal variability and resilience, we can only insist, not unreasonably, that any change is for the worse. We know little about how to maintain processes or species, either individually or in assemblages. We do know that climate change will require that species alter their ranges and most likely their associations, as has happened repeatedly since the end of the last glaciation, if they are to survive.13 We do not know which species will be able to do this and which will not, and what will be the consequences of losing those that cannot. The usefulness of historical study in evaluating potential change and formulating policy may be illustrated with a specific example: problems of sustaining the oak-dominated forests of the northeastern United States. In analyzing the problem of oak regeneration, Lorimer observed that “recent evidence . . . has accumulated to the point where there is no longer any question that oak will be displaced on many sites” in its range in North America. In part because this is regarded as a major silvicultural problem (“one of the most serious silvicultural problems in the eastern United States”), foresters and ecologists have devoted considerable research to trying to understand its causes. The use of historical study, both to assess the vegetation of the area before major deforestation in the last few centuries and to interpret changes in factors responsible for oak regeneration, constitutes a model for applying

Biospheric Sustainability  201

historical analyses to current ecological questions. This research raises many questions about sustainability as it refers to preserving seminatural ecosystems, processes, and genetic diversity.14 Among these are: 1. 2. 3.

4.

What community do we want to maintain? When a community is changing, how do we know which changes are temporary and which signal future loss of the community? How have past human impacts and climate change affected the potential to maintain the community, whether by nonintervention or by active management? How can we combine theory, experimental evidence, and observations to arrive at deeper understanding of ecosystems?

Critical Features of Oak-Dominated Forests Oaks are synonymous with the forest vegetation of large areas of eastern North America (fig. 11.1). Pollen analysis shows oak reaching its current extent and dominance by about eight thousand years ago. Variations over the last millennia in relative amounts of pine, hickory, chestnut, and other secondary taxa as well as charcoal have been correlated with other, independent indicators of climate fluctuations over these eight thousand years, but have not altered the basic importance of oak regionally. The earliest evidence for the species of oaks, rather than just the genus, are eighteenth-century land surveys, since one cannot separate oak species by pollen. These surveys confirm pollen evidence that oak was dominant, along with hickory and chestnut, and show that white oak was the predominant species in many areas in that period. Because this forest type was widespread when European settlers arrived in North America and for many millennia before then and supports a wide diversity of both flora and fauna, many ecologists believe that it embodies the natural biodiversity and typical ecosystem processes of the region and should be sustained. Foresters are especially interested in its maintenance as a source of valuable timber.15 At finer spatial scales, these forests were highly variable, depending on myriad factors such as slope, aspect, and soils. Human disturbance history over the last several hundred years has greatly modified these forests, but they have maintained the basic latitudinal pattern in composition, with more abundant oaks in the southern parts of the region and maples toward the north. Climate appears to be the overriding factor determining the patterns on a regional scale, even in the face of major novel disturbance regimes (fig. 4.3).16 The goal of foresters, then, has been a forest where white oak is the domi-

202  Contributions of Historical Ecology

Figure 11.1. Hardwood volume in the eastern United States. (Data from Oswalt et al. 2014, table 24.)

nant tree, with associated hickory, other species of oaks, and a variety of other species of trees. Chestnut has succumbed to a blight, so is unlikely ever again to be a major part of future forests. This forest would have a very diverse native herbaceous and shrub flora, and a diverse fauna. The majority of the regeneration would be oak.

Current Changes While there are stands today that fit these criteria, most do not. Advance regeneration (seedlings and saplings) in most stands consists of the more shade-tolerant and mesic species—eastern hemlock, beech, and sugar maple—or the fast-growing species characteristic of more open sites—red maple, tulip poplar (Liriodendron tulipifera), black gum (Nyssa sylvatica), ash, and black birch. Shade-tolerant sugar maple dominates even the early successional stages of some of these forests, indicating a permanent change in forest composition. Patterns of secondary succession are highly variable, and after even a century or more on former plowed fields do not lead to reestablishment of the diverse herbaceous flora associated with forests that have never been cleared.17 Two categories of forest stands contain clues about changes in these forests: old-growth stands, which contain many very old trees and a diverse flora; and older successional stands that have grown after the abandonment of farming or after logging. Neither of these is an exclusive category; overlaps are common. For example, many stands that include very old trees have been grazed and selectively logged, as have most secondary forests. Finally,

Biospheric Sustainability  203

the composition of any of these stands can be compared with that inferred for precolonial forests to indicate changes in the interim. Almost all extensive forest stands have regenerated after repeated logging or abandonment of agriculture as well as being subject to such natural disturbances as hurricanes. In such stands in Connecticut in 1907, more than 90 percent of the trees had originated in about 1870–75 as sprouts that had overtopped the numerous but slower-growing seedling trees (table 11.1).18 Seedlings in these stands also suffered from burial by heavy litter, fires, and browsing by squirrels, rabbits, and mice. Although they could sprout back from such damage several times, they gradually lost vigor and died, lasting maybe as long as thirty years. In the mid- to late twentieth century, studies in many oak forests in the northeastern United States encountered very few oak saplings in such sprout-dominated forests. In some years, oak seedlings might be numerous, but most disappeared in a very few years and other species were more common as small trees.19 Not only is reproduction apparently insufficient to replace older oak trees but drought, frost, outbreaks of fungi, and insect pests, often exacerbating each other’s effects, have killed mature trees prematurely throughout their range in the past few centuries. Hard hit, for example, have been stands in the Appalachian Mountains, where mortality has been associated with defoliation by non-native insects. The historical factors of uniform age in previously harvested stands, low diversity, and sprout origin are other factors contributing to the problem, and possibly also atmospheric deposition of pollutants such as nitrogen compounds, acidified rain, and ozone. The complexity of interacting historical factors makes it difficult to single out any one as most critical, but it is clear that much can be related to land-use change and industrialization.20 On the other hand, old-growth stands have never been clear-cut, so they constitute a very small remnant carrying the tenuous thread of continuity from the aboriginal forest into the present. Human activities have altered them all, however, to a greater or lesser extent; the continued existence of “pristine old growth” is indeed a myth.21 The few unlogged oak woodlands are small and isolated and have been subject to heavy use for firewood and grazing as well as to changed disturbance regimes; ongoing changes are apparent in most. The dominant white oaks in an old, twenty-four-hectare unlogged stand in New Jersey, for example, probably originated as seedlings, but in the late twentieth century there was little evidence of replacement from seedlings or saplings (table 11.2). Beech and red maple as well as white and black oaks dominated younger stands in the region, and in these there were more beech and red and sugar maple than oak saplings and seedlings.

Table 11.1. Average number of trees per acre in three different forest associations in the Housatonic Valley of Connecticut

Chestnut slope Sprouts From seed Mixed slope Sprouts From seed Oak ridge Sprouts From seed

Chestnut

Chestnut oak

White oak

Red oak

Black oak

Hickory

Sugar maple

Red maple

Other

Total

%

411 3

222 0

24 3

46 4

4 0

67 6

61 2

97 2

138 39

732 20

97 3

157 3

124 8

53 4

135 12

13 1

110 6

146 4

153 3

232 64

891 41

96 4

29 0

272 14

99 5

148 12

45 0

251 8

21 2

72 1

133 16

937 42

96 4

Average age (years) 28

32

38

Data source: Schwarz 1907.

Biospheric Sustainability  205 Table 11.2. Relative abundance of potential canopy tree species in Hutcheson Memorial Forest, New Jersey Species

Trees

Saplings

Seedlings

Acer negundo Acer platanoides Acer rubrum Acer saccharum Ailanthus altissima Carya spp. Fagus grandifolia Fraxinus americana Nyssa sylvatica Prunus avium Prunus serotina Quercus alba Quercus rubra Quercus velutina Sassafras albidum Ulmus spp.

0.7 3.4 8.9 0.7 0.3 10.5 0.9 15 0 9.2 0.3 30.5 9.5 8.8 1.2 0

2.3 5.7 25.2 6.5 0.5 2.6 4.4 44.7 0 3.9 0 2.1 0.8 0.3 1 0

0.3 5.3 4.6 10.2 0 2.1 0.8 21.9 2.1 4.3 2.4 1.1 0 42.7 1.8 0.3

Data source: Forman and Elfstrom 1975.

In other areas, such as east-central Indiana, other species are also apparently replacing oaks in old-growth oak-dominated woodlots.22 The differences between the composition of old-growth oak forests and precolonial forest, as reconstructed by survey data, suggest the possible effects of human impact in the past several centuries. In one such stand in Pennsylvania in 1992, 87 percent of the stems were chestnut oaks, red maples, and black and yellow birches; some of the chestnut oaks were at least three hundred years old. In the eighteenth century, however, these four species had accounted for only 9 percent of the 513 witness trees tallied for the area, while 80 percent of the trees were black and white oaks, chestnut, and hickory. Most of the saplings and seedlings were red maple and yellow birch in 1992, with some chestnut oak, implying yet a further change in composition. In Illinois, one-hundred- to two-hundred-year-old forest stands on south-facing and ridgetop sites were fairly similar to forests reconstructed from precolonial surveys on similar sites, but north-facing stands had more sugar maple and beech trees than the reconstructed presettlement forests (fig. 11.2).23 It is likely that precolonial oak-dominated forests were very patchy because of the sporadic nature of disturbances such as ice storms or wind

206  Contributions of Historical Ecology

Figure 11.2. Comparison of forests on six different site types in Illinois. On each site type (ridge south, south-facing slope, ridgetop, high and low north-facing slopes, and terrace), paired comparisons are made between prehistoric composition derived from survey data (PS), old-growth stands (OG), and successional stands (SG). (Data from Fralish et al. 1991.)

damage, and that the disturbances occurred at variable frequencies over centuries. Thus, extrapolating processes in current forests over only a few decades is unlikely to explain the processes that had been acting over centuries in the distant past.24 Many formerly common species have been eliminated from this region over the past several hundred years, including wolves, mountain lions, passenger pigeons, and chestnut trees. Some currently regionally rare species, for example, small whorled pogonia (Isotria medeoloides), American ginseng, and bald eagles, were formerly much more common in these forests. Others, like deer and beavers, were almost eliminated by the late nineteenth century or even earlier but have rebounded because of changed habitat and reintroductions by people. The ranges of many others, such as birds that live in forest interiors, have been greatly limited by the reduction in forest cover and the fragmentation of remaining forests in the nineteenth century.25 Chestnut, a major species in many areas in 1900, was eliminated as a forest tree in the early twentieth century by an introduced blight. How important was this loss to the remaining species composition and functioning of these forests? At the time, chestnut formed 50 percent or more of the basal area of many forest stands in the Appalachians, although it had been only locally common in precolonial forests. In most areas, it appears that species that were formerly important in the forests are replacing the chestnut, perhaps

Biospheric Sustainability  207

shifting the composition somewhat toward such earlier successional species as black birch, red maple, and tulip poplar, but in general having minor impact. The main replacements have been oaks. On the other hand, a forest modeling experiment indicates that the loss of chestnut will lead to long-term changes in forest function, as indicated by biomass and leaf area index, as well as in composition. Forests in which the chestnut trees died decades ago and that now have a closed canopy may still be responding to this loss that occurred well in the past.26 In many areas, the reintroduced white-tailed deer eliminate most vegetation less than a meter tall except species that they do not prefer, such as the prolific non-native Japanese barberry and native white snakeroot (Eupatorium rugosum) (see fig. 6.3). Especially where hunting is not allowed, changed habitat and extermination of most predators (except cars!) have contributed to the sizes of deer herds. In the small, old-growth stand in New Jersey, by the early twenty-first century, intense browsing by deer had removed almost all of the understory, which had been replaced by a carpet of non-native stilt grass (Microstegium vimineum), which deer do not eat. Where deer browsing is not very intense, such non-native vines as Japanese honeysuckle and oriental bittersweet inhibit regeneration of trees, while the non-native Norway maple often flourishes. On the other hand, in deep forest, far from edges, deer are less common, so the impact is less. The complex interactions of changing past and present land use, hunting, and species interactions contribute to the changing composition of these forests.27 To summarize, although there are still extensive oak-dominated forests in this region, the structure and diversity of these has been greatly changed over the past several centuries. These forests have value for such uses as timber production, recreation, and watershed protection. Foresters seem to agree that the value is highest if they are maintained as oak forests, yet it is possible that they might serve the other functions even if their composition were to change. Loss or severe further reduction in a major forest type, however, would entail a decrease in the diversity of the regional vegetation, possibly reducing its ability to respond to further perturbations. The question then becomes, do we know enough to sustain such a forest? Can it sustain itself, or have people modified conditions so much that continued active management would be necessary to maintain it? What has been the role of climate change in the last few centuries? And finally, just what kind of forest do we want? These questions may seem trivial in the face of the major losses of species and old-growth forests in tropical regions, but answers for this well-studied exemplar may provide some directions for research and management in complex tropical systems.

208  Contributions of Historical Ecology

The Contributions of History to Elucidating Causes of Changes Oak is a widely distributed genus that is found in North and South America, Eurasia, and North Africa. There are 450 or more species, all of which formed part of a variety of assemblages during the Tertiary period. As the climate alternated between glacial and interglacial during the Quaternary period, the differences among species in their tolerances of changed climate and their migration rates led to a shifting mosaic of both plant and animal communities. Because more time was spent in glacials than in interglacials, it is unlikely that the relations that characterize the interglacials, such as the one that we are currently experiencing, are highly evolved. In addition, just as the last glaciers began to retreat and the climate warmed and stabilized to some extent, humans arrived in North America.28 It is against this backdrop that we must evaluate the stability and dynamics of this forest vegetation. According to pollen records, oak migrated north rapidly in North America as the glaciers retreated, reaching its northern limit by about eight thousand years ago. We can only conjecture what the interactions of people were with these plant and animal communities. As large herbivores like mastodons and giant beavers became extinct, along with their predators, it is likely that smaller herbivores increased in density to substitute for them, changing the kinds of plants eaten. There is evidence there was an increase in fire and that a plant community that has no modern analogue became at least locally common in some areas.29 The people who immigrated to North America were adept at using fire; otherwise they could never have survived their long, arduous trek from Asia and the harsh cold of their early years on the continent. Their arrival in an area must have brought an increase not only in use of resources but also in fire ignitions. Over time, populations waxed and waned and cultures and climates changed, causing varied impacts on the biota. With the advent of agriculture in the Northeast about one thousand years ago, shifting agricultural clearing, sometimes extensive and at least semipermanent, introduced yet other potential impacts on the vegetation and fauna. Although the overall pattern of oak dominance in the region was fairly stable over the last eight thousand years, locally the proportion of oak fluctuated. Small disturbances or short-term changes in climate may have caused these variations, but such human activities as setting fires and local clearing may have been important, too. When these practices are superimposed on the overall pollen record for the region, however, the pattern that emerges is still one of overall oak dominance that lasted for many millennia, even as

Biospheric Sustainability  209

charcoal evidence suggests that fire regimes and climate fluctuated.30 Unfortunately, the record does not allow us to assess the contributions of different species of oak, most herbaceous plants except grasses, and insect-pollinated taxa. Nor does it allow fine-scaled studies of successional dynamics or landscape patterns. It does, however, indicate a wide range of conditions, from cool areas with more hemlock and beech to warm ones with hickory, from wet areas with alder to dry ones with pine. Seventeenth- and eighteenth-century colonists in eastern North America faced a daunting task: vast forests stood in the way of planting the crops they needed to sustain their lives. Some were able to exploit fields previously cleared by Amerindians, but most had to remove large trees to make fields. Deforestation was a goal to be achieved as quickly as possible, with the harvested trees often forming the first source of income from the new land. After farmers removed or killed the trees, they plowed the soil, planted crop plants, pastures, orchards, and, inadvertently, weeds, and sent their livestock into remaining forests for forage. Where soils were poor or hillsides too steep for crops, they cut trees for fuel. Where it was marshy, they dug drainage ditches. This process of deforestation was rapid and drastic. For example, in one agricultural region of New York State, a few small fields had been cleared by 1777, totaling perhaps 7 percent of the area. Dense forest surrounded these fields. About a hundred years later, in the mid-nineteenth century, more than 90 percent of the land was cultivated (fig. 8.4). In north-central North Carolina, farmland covered 80–90 percent of the land by the mid- to late nineteenth century. The general picture is clear: major deforestation for agriculture took place within a century or so of the arrival of European settlers. The effect on the regional vegetation pattern was equivalent to that of a major climate change.31 In agricultural areas, native trees remained in scattered woodlots, as small trees regenerating along field edges, and as plantings along roads and around buildings. Agricultural fields isolated most of these from other woodlots, although some hedgerows and stream-bank trees formed more or less continuous, though narrow, corridors (see figs. 9.3, 9.6). Farm use of the woodlots and hedges for fuel, building materials, litter for fertilizer, and wood ashes was intense. Farmers also collected hickory nuts and chestnuts and hunted small game and deer, reenacting the activities of a hunting and gathering culture. They eliminated some animals through overhunting, habitat reduction, and deliberate attempts to reduce the threat to domestic livestock and other resources. Passenger pigeons, for example, were killed in their roosts, both because they were edible and because their large concentrations damaged the forest trees. In their major roosting sites during migration they had provided a significant source of fertilizer to the forest floor.32

210  Contributions of Historical Ecology

Many forested areas, however, were not suitable for agriculture. Massive harvesting of wood from these never eliminated the forest completely but did change its structure. By 1900 CE in New Jersey, for example, forests covered only 10–20 percent of the land in fertile valleys, but extended over 80–100 percent of upland areas with rocky soils on steep slopes. The average age of the upland forests was only about thirty years, while woodlot trees were often much older. Such a pattern of land use favored species that sprout easily and those that thrive in open areas, for example, birch, replacing ­seedling-grown or root-sprouted oaks and chestnuts.33 Starting in the late nineteenth century and continuing into the twentieth, farmland in this region was abandoned, and heavy, destructive cutting of forests for charcoal ceased. Abandoned farms became young forest stands, usually with no active management (fig. 8.4). Most of the sprout forests also grew back without active management. Many of these would be a hundred years old by 2000 CE, with much of their disturbance history, caused by human activities, hidden in the past, although some grazing and selective logging persisted. In the late nineteenth century, surveys that catalogued the forest resources of the individual states generally concluded that the resources had been severely abused in the past and recommended future management to maintain the stock of timber and to protect watersheds. For example, in 1900 the New Jersey state geologist commented that the “superior quality of water from . . . wooded districts, over that gathered in a cleared farming country, makes it desirable that the forests in the Highlands should be kept, and not be cleared and put in farms.” The state geologist also noted the value of forests for education because “there trees, shrubs and herbaceous plants are found in their native habitat, and their relations to one another are there studied to the best advantage.” In addition, “it would be a public misfortune to lose any of our characteristic species or their natural grouping, as now existing, or to have our rich botanical heritage marred by general deforestation of the State.” Such forests should be managed wisely for timber, especially chestnut, probably as long-term coppice.34 The most serious threats to forest vegetation were thought to be clearing for farmland, fires, grazing, and overcutting. Removing the thick understory and dead, dying, and diseased trees and preventing forest fires and grazing would allow natural forces to produce healthy forests. Nature needed a little help from people but would basically do the job alone.35 The assumption was that increased disturbance by humans was upsetting the natural balance of the forests. The role of forest managers was to prevent this disturbance in order to let natural processes return to the forests.

Biospheric Sustainability  211

A shift in attitude about management of forests, from nonintervention to the reintroduction of disturbance, characterized the second half of the twentieth and the early twenty-first centuries. It has been argued that the only important factors likely to have altered forest composition over such a large region are changed fire regimes and lack of current canopy-opening disturbances.36 For example, eliminating historical disturbance, especially fire and possibly buffaloes, has been blamed for changes in old-growth forests in Illinois: “Our present old-growth forest is an artifact of near total protection while presettlement forest developed under a fire regime.” To maintain this forest type with its typical herbaceous understory, the management of oak forests may require the reintroduction of “disturbance at the level found in presettlement forest.” In Pennsylvania, too, fire was assumed to have been the major disturbance responsible for presettlement dominance of oak. Since then, however, in addition to the removal of disturbance, selective logging in the 1930s and 1940s is assumed to have accelerated succession to more mesic trees by releasing those that were in the understory (successful there because of the absence of fire). In the absence of major disturbance, these stands were assumed to be heading for dominance by mesic hardwoods.37 A more nuanced analysis of the history of this forest region, however, paints a more complex picture and offers other interpretations for the current apparent decline in oaks. Oak as represented in pollen cores did not shift with charcoal frequencies over the last ten thousand years in a detailed study in New York State. While at a coarse scale, there was continued oak dominance over the last millennium under a fairly stable climate, some notable changes are evident at a finer, centennial scale. Abundant evidence points to a period of relative warmth, often called the Medieval Warm Period (MWP), from about 800 to 1300 CE, followed by a cooler period, the Little Ice Age (LIA), from about 1400 to 1900 CE.38 In southern New York State, a detailed study indicates abundant pine and charcoal during the MWP. At 1300 CE there is an abrupt decrease in the amount of charcoal, that is, fires, with charcoal almost disappearing by 1400, when the amount of oak increased from 20 percent to 30 percent of the pollen. At a scale of specific sites, oak did not depend on fire.39 A detailed study of oak forest dynamics from about 1750 to 2000 integrates multiple drivers of change to explain the shift from oak to maple. Analysis of multiple paleoclimate records for the eastern United States shows that multiyear droughts characterized the early part of this period, to about 1900. This kind of climate regime would favor oak over maple. The last seventy-five years have been wetter than any period since 1500 CE, with fewer droughts. The recent moister climate would favor mesic species such

212  Contributions of Historical Ecology

as maple and beech over oak. It is also likely that there would be fewer fires. Moisture as well as temperature is an important factor in controlling forest species composition, even over fairly short time periods. Major deforestation, destruction of the passenger pigeon, white-tailed deer, and turkeys characterized the nineteenth century. The turn of the twentieth century also saw massive fires caused by poor logging practices and increased ignitions from steam engines, preventing regeneration of maples. This was the time of regeneration of much of the current oak-dominated forest and a time of maximum forest fragmentation. In the last century, populations of white-tailed deer have recovered to historic highs, as the climate has become less droughty and fires have been suppressed, leading to conditions that favor regeneration of maples.40 Other changed conditions caused by human impact have played a role in changing these forests. Seas of cultivated farmland that isolated woodlots inhibited the transport of animal-dispersed seeds from one woodlot to another while favoring the dispersal of species such as birch, tulip poplar, and maple with windblown seeds. Hedgerows favored species spread by birds. The absence of passenger pigeons may have inhibited long-distance dispersal of large seeds, even though blue jays may have partly filled this niche. In ­England, defoliation by caterpillars that drop from the canopy and then feed on young seedlings is blamed in part for the lack of oak regeneration under oak canopy. Perhaps the increase in non-native herbivorous insects such as gypsy moths is important in North America.41 The processes and composition of these forests are not only the consequences of factors that can be measured and analyzed currently. They also reflect logging history, changed disturbance regimes and landscape patterns, and altered soil processes over the life of the trees (and perhaps even longer). Generalizing over the landscape misses critical site-specific factors, such as local fire history or local contingencies arising from landscape patterns.42 Non-native earthworms are also altering nutrient cycling in the soils of many forests.43 Assuming that the landscape will return to some pristine, revirginized condition if left to its own devices flies in the face of the evidence by ignoring critical intervening factors. The sustainability of this landscape must be defined in light of these factors.

Prediction and Management for Sustainability Historical ecology has been used in restoration efforts mainly as a means to determine a condition sometime in the past to use as a goal, that is, using historical systems as ideal analogues. As has been seen, however, defin-

Biospheric Sustainability  213

ing these past conditions presents numerous issues, including both temporal and spatial scales, and often ignores the importance of changing ecological drivers forming the prologue to the present.44 These changing drivers often constrain the likelihood of restoring something from the past. In the case of the eastern oak forests, these would include such factors as altered age structures, climate, and introduced species. Historical ecology assists in setting goals by discovering possible alternatives and by analyzing the processes, both past and present, that may make one of these alternatives a better fit for broader restoration or sustainability goals, such as biodiversity, productivity, or improved quality of life. Historical studies reveal the importance of considering patterns at different spatial and temporal scales, and with different levels of precision. In a broad sense, climate and species constrain a large universe of possible states. Species cannot grow outside their climate limits, though long-lived species may persist well after the climate no longer supports regeneration. While earlier studies established that the climate stabilized to a broad extent after about eight thousand years ago, more recent research, using a wide variety of proxy data to infer past climate at sites around the world, indicates that there have been fluctuations in climate that have had significant effects on land cover and human populations during these last eight thousand or so years (table 11.3). Droughts can often be correlated with declines in human populations as well as with changed vegetation, regardless of the complexity of the civilizations and agricultural systems, apparently over time scales of centuries.45 The temporal scale of much historical ecological data does not permit detailed analysis of short-term changes in species composition, in part because of the coarseness of most dating techniques, except for written documents, and the lack of detail in the record. Long-term studies, some reaching back close to a century, suggest the variability of change over time; simple succession to climax along a predetermined path seems to be at most an exception, and certainly not the rule. Over the entire region of the oak-dominated forests, one can say with confidence that species of oaks have been the major genus for six thousand to eight thousand years, but the dynamics of regeneration are too fine to discern. Pollen studies do not sample most taxa that are insect-pollinated, for example, and the spatial scale of the area that is sampled is also coarse in most studies. Charcoal and other proxy data in sediments suffer from similar problems. Did droughts, fires, wind- and ice storms, and outbreaks of insects and diseases create a mosaic of stands of different ages, so that oak remained dominant over the landscape but varied locally?46 It is self-evident that one could not have predicted the current structure

214  Contributions of Historical Ecology Table 11.3. A selection of studies of climatic fluctuations over the Holocene epoch

Location Africa—Kalahari N. America—Nevada N. America S. America—southern Andes Europe—Levant Europe—Iberia Global

Kind of climate record

Reference

Multiproxy Pack rat middens Tree rings Multiproxy

Cordova 2017 Nowak, Nowak and Tausch 2017 deMenocal 2001 Morales et al. 2017

Pollen Multiproxy Multiproxy

Langgut, Finkelstein and Litt 2013 Aranbarri et al. 2014 Mayewski et al. 2004

of the forests of the Northeast from the dynamics and structure of those of the eighteenth century. Even if there had been ecologists in the eighteenth century who were able to decipher all of the processes that were active at the time, prediction would have been impossible. They would have had to foresee the effects of land clearing, grazing, a changed fire regime, timber cutting, changing climate, and the introductions of non-native species. They may have anticipated a dearth of seed sources, edge effects that would penetrate the borders of the forest, and the impacts of domestic livestock. They could have observed the decline in deer browsing as deer populations plummeted, and perhaps changes in other small-game species that were hunted, such as squirrels and rabbits. They could not have predicted the invention of the locomotive with concomitant increases in fire ignitions, the explosive increase in deer herds in the twentieth century, or which exotic species would invade the forests. The chestnut blight and the destructive use of hemlock for bark used to tan imported hides would also have been beyond the realm of reasonable expectations. These hypothetical eighteenth-century ecologists would have no doubt predicted expanded agriculture in the region and seen no reason for the pressure on the local forest products to decrease. Presumably such surprises lie in our future as well.

Conclusion What are the consequences of such insights for the questions of sustainability? First, we do not know the composition of vegetation or fauna at any particular historical period at any specific site. Even in North America, where conservation often aims to reconstruct or maintain ecosystems as they

Biospheric Sustainability  215

were before the arrival of European colonists only a few centuries ago, the details are hazy, and of course, such an observation is even more pertinent to the study of historical vegetation in regions settled by dense populations even farther back in time. Second, climate is only one of a complex of factors that influence vegetation. Disturbances, both “natural” and human-­ mediated, have played and continue to play major roles in determining the structure and composition of plant and animal communities. The magnitudes and kinds of such disturbances have changed in the past and can be expected to change unpredictably in the future. Most current ecosystems are likely not in a steady state, so if the goal of sustainability is to maintain systems as they exist today, complex disturbance regimes will probably be necessary. It has been observed with relation to oak forests in England that “if the objective of management is to produce something approaching mature forest then the management prescription should be unequivocally one of non-intervention. This will necessarily produce an ecosystem of zero net-productivity, probably a woodland of low aesthetic value and one which possibly would provide some surprises for ecologists in the final composition of its tree canopy!”47 We may not agree with this observation, but it is becoming ever more obvious that most ecosystems that we value have a long history of human intervention, both intentional and nonintentional, and to maintain them will require continued intervention. Proposals of how to maintain oak forests in the northeastern United States today require very focused, intensive manipulations, the results of which may very well turn out not to achieve the desired results. Decisions affecting sustainability include mitigating the effects of undesirable changes that have already occurred and eliminating causes of future changes perceived as deleterious. Dealing with the first level implies local solutions, both private and public, about managing land and ecosystems. Dealing with the second implies political solutions, convincing people and governments that some sacrifice now will pay off in the future. Both of these decisions imply the acquisition of potentially critical habitat to accommodate ecosystems that will be eliminated by climate change where they now exist. But do we know how to manage even those lands that have been acquired to preserve their current biota? Ecological modeling is one way to predict the consequences of changing conditions. Models are still unable, however, to reliably predict current conditions at the scale at which management decisions must be made. For example, two models that have been developed to predict species distributions are based on sampling known populations. When tested, they both have errors when used to predict distributions in a nearby area. Some of the reasons for

216  Contributions of Historical Ecology

failure are issues dealing with the scale of habitat diversity and historical landscape change.48 A major problem in designing these models is how to deal with species absence in the sampling data. For historical data, this issue is even more serious than for modern data. In addition, only very complex models can deal with changing drivers over time, which have been critical to developing current ecosystems and landscapes.49 The models are in essence ahistorical, even where they use past conditions for testing, because they cannot incorporate unanticipated contingencies or conditions that are novel. History also records the uniqueness of events caused by the juxtaposition of culture, climate, and other factors at a specific time. Models, however, are tested by multiple runs, producing a central tendency, the most likely outcome, based on the common elements that run through history. In the real world, however, the system will run only once, and even though the central tendency is the most likely result, no extreme outcome can be ruled out. This in no way invalidates the model, just as an erroneous weather report does not invalidate the model on which it was made. History documents such contingencies, indicating some of the range of possibilities for change and how difficult they are to predict. Understanding this range affords additional information for decision making, especially for indicating to the public what the uncertainties of the models are and why the models need not be ignored just because they are uncertain. We approve budgets for future income and expenditures, even though actual income and expenditures rarely conform to the expected because of unanticipated contingencies. A budget fraught with uncertainties is better than no budget at all, even though it approximates the future only roughly. By comprehending the range of past conditions we can make more realistic models for the future. Forests in many parts of eastern North America are less fragmented and heavily used today than they were a century or even fifty years ago. These young, recovering stands are growing on soils that have a history of disturbance, lack of old mound and pit topography, increased air pollution, high levels of carbon dioxide and available nitrogen, non-native diseases and pests, high levels of herbivory, absence of such formerly abundant species as passenger pigeons and large chestnut trees, changed climate, and other changed conditions known and unknown. That they are as similar as they are to precolonial forests suggests the critical basic role of species availability, macroclimate, topography, and some essential soil characteristics, all of which are amenable to conventional forest modeling. Most important tree species in this region are, however, already under the onslaught of disease or pests: gypsy moth, chestnut blight, beech bark disease, Dutch elm disease, hemlock woolly adelgid, and emerald ash borer, among others. In the face of

Biospheric Sustainability  217

these, the assumption that an unmanaged forest will develop in some “natural” way to maintain the diversity that existed before human intervention is a will-o’-the-wisp. Without knowing the history of an ecosystem, how it got to its present state, we have no way of guessing what kind of management is necessary either to deflect it from its current trajectory or to maintain the status quo. History teaches us that management for sustainability will always have to be flexible in the face of changing human impact and the inherent variability in ecological systems.

This page intentionally left blank

Conclusion Toward the Future: Research and Applications

Historical ecology delves into the perplexing realm between mechanistic, experimental science and descriptive “natural history.” To most people, the ideal of the natural world conjures up images that are eternal and unchanging. Add humans to this world, however, and it becomes dynamic and even precarious.1 Although we realize that the world independent of humans changes over time, the pace of such change has usually been slow, except for catastrophic events like volcanic eruptions. Mechanistic science based on the premise that constant and knowable laws govern both static and dynamic relations in this world has yielded much practical information, confirming the basics of the premise. Subject poinsettias to a certain light regime and they will flower; replace the terminal buds of flowers with auxins (a plant growth hormone) and lateral buds will remain inhibited. We may not understand all the details of these processes, but as far as we know at least some are consistent and can be predicted, and we expect that they will work the same a century from now as they do now. On the other hand, we cannot predict the potential activities of people in the future, except in a gross sense. Although all people have similar requirements for food, shelter, and reproduction, cultural developments have led to an enormous variety of systems for fulfilling these requirements and many others. These adaptations cannot have been predicted from basic, constant

219

220  Conclusion

laws, even those that include dynamic equilibria, just as future adjustments cannot be predicted. In addition, this changeability in adaptations means that observations of current human activities cannot allow one to reconstruct the past or to predict the future. Ecosystems fall somewhere between these extremes. Ecologists act under the assumption that there are basic laws that control the essential patterns of ecosystems, for example, that productivity of lower trophic levels limits that of higher ones. Superimposed on these are, however, the outcomes of evolutionary and climatic history and past disturbances, both human and nonhuman. The human-mediated changes are those least amenable to understanding according to general laws. Evolution, for example, has led to similar adaptations under similar climates, such as the reduced leaves and succulent stems in unrelated African euphorbias and American cacti. Ecologically, the pattern of recovery of a forest after a hurricane can be predicted within limits. But the invention of agriculture in its many ramifications did not follow logically from hunting and gathering. It is by distinguishing the unique contributions of humans to current ecosystems that we can incorporate them into our understanding of ecological processes. Consistent laws may govern the sequence of species that recolonize bare soil, but the kind of human disturbance that created the bare land may interfere with this regular pattern, or newly introduced species may disrupt it. Some laws may be robust in the face of the variety of human activities, for example, productivity or biomass accumulation may proceed the same regardless of prior conditions, but others may not hold constant, for example, prior grazing or plowing may alter nutrient cycling in ways that persist for centuries. Prior conditions may confound experiments carried out on sites that differ not only in current conditions but also in history. Human history is not repeatable (neither is any historical circumstance exactly repeatable) and is generally not reversible, yet by learning the contributions of history to present systems we can better understand structure and process to improve prediction and management. In addition, past human activities act as experiments from which we can deduce general laws, laws that would be much more difficult to understand if we had to rely on the usually slow workings of changing natural conditions.2 Studies using multiple lines of evidence, precise dating, and modeling help establish causation, including cultural drivers of human activities. For example, a model using a variety of scenarios for past environmental driving forces, coupled with field and sedimentological study, strongly suggests that the lack of trees in some valleys in subarctic Scandinavia results from winter use of wood by people a thousand years ago. In another study, models of various fire regimes, either natural or

Toward the Future  221

anthropogenic, contributed to better understanding of factors important in managing current forests in the New Jersey Pine Barrens.3 The contributions of historical study to the present are perhaps best understood in terms of considering history as analogue and history as prologue. Using history as analogue, we search for patterns that are repeated and use these to interpret the present and predict a range of options for the future, as well as often to set goals for conservation.4 Using history as prologue, we emphasize the mutability of systems and their drivers and their unique contributions to the present. Analogues often furnish examples of states that some judge to be preferable to the present. They also may provide cautionary tales or evidence of past processes and structure. History as prologue seeks to partition the impacts of the past into categories that can shed light on their potential contributions to current conditions and to predictions of the future. Both approaches have value, but together they reveal insights that are useful both for understanding ecological processes and for conservation. In the subarctic study, the historical analogue was a forested valley, indicating that the climate a millennium ago did support trees. The prologue to the present was the felling of these trees, followed by various processes related to lack of tree cover that changed soil characteristics so that today trees do not grow even if the climate is conducive to them. Both the analogue as a goal and the subsequent processes have created current conditions. A wild and rugged landscape with no ancient analogue is the open heathlands of Great Britain. Although the presence of sheep and stone walls in these areas attests to current human impacts, the very origin of these treeless heaths millennia ago is associated with past human activities, the prologue to the present. The ancient analogue would be a forest. A complex interaction of land clearance, starting in the Neolithic, deteriorating climate, and a variety of changes in agriculture, as in County Mayo in Ireland discussed in chapter 8, has resulted in the current heathland landscape. While forest did regenerate in some areas where farming was abandoned, because of increasingly wet and cold conditions, most of these areas are currently covered with blanket peat (except where it has been removed by mining).5 A combination of paleoecological and archaeological evidence suggests that peat might have encroached on forest even without agricultural clearing.6 Heaths here, and most likely in other parts of western Europe, result from a complex interaction of cultural and climatic factors acting over millennia. In addition, land management continues to affect them in complex ways. Because of economic incentives in some places there is less grazing, allowing nonheath shrubs to invade, while in others the incentives have encouraged farmers to plow the heath and plant grass for more intensive grazing. Much of the heath is being

222  Conclusion

mined for peat. The very existence of the heaths is tightly linked to changing human intervention as well as to climate in ways that have not yet been adequately elucidated.7

Incorporating Historical Research into Basic Ecological Studies Historical ecological studies have demonstrated that in a wide variety of landscapes, human activities in the past, often the very distant and dim past, have left legacies on current conditions.8 These studies can often provide data that are precise and accurate enough in time and space to help interpret current structure and function in a variety of ecosystems. For example, a combination of sedimentary studies, field work, and models in part of Tuscany, Italy, indicates that silver fir (Abies alba) was lost from the local forest because of moderate browsing by domestic livestock, and that without browsing or frequent fires it would thrive in current or warmer conditions. In the Czech Republic, landscape models using palynological as well as archival data indicate that conifers were a major component of Holocene forests until about two centuries ago. This brings into question the current emphasis on deciduous forest as the preclearance land cover of the region. Spatially precise historical data reveal the highly dynamic nature of seminatural grassland in Sweden.9 At the species scale, a consideration of the niche characteristics of species has been used to predict patterns of biodiversity, but the importance of nondeterministic factors makes this difficult. A species cannot be found outside its niche requirements in terms of climate and soil moisture, for example, but where within those conditions it will be found is much more difficult to predict. These factors are often described as stochastic, but many of them can be described more specifically in terms of site and regional history. Past cultural activities, both those that have ceased and others that have continued into the present, are examples of factors that can be defined and used to predict community diversity when combined with the deterministic characteristics of species.10 Historical research has established that the maintenance of the ideal habitats of some species is specifically related to cultural practices. Much of the high biodiversity of, for example, the Mediterranean steppe flora is maintained by the continued grazing by sheep. Similarly, regular harvest has led to the biodiversity of the coppice woodlands of western Europe and the British Isles. The niches of the species in these habitats depend on historical, cultural factors established by a long history of sustained use.11

Toward the Future  223

A study of forest recovery after massive hurricane damage illustrates how historical studies coupled with current experimental research can provide insight into community-level processes and highlights the importance of historical analysis. In 1938, a major hurricane devastated the mainly white pine forests of central Massachusetts. In the years after the hurricane, there was extensive erosion and the establishment of a forest consisting primarily of early successional birch and bird cherry (Prunus avium). Was it appropriate to extrapolate from this that a consequence of major disturbance of a forest was to set it back to an earlier successional stage? An ingenious one-hectare experimental plot was established at the Harvard Forest Long-Term Ecological Research site to test the impact on regrowth and soil conditions of knocking down all the trees. This study found that while the destruction of the trees had significant effects on the forest structure over succeeding years, species composition was little changed, as sprouts and saplings of the former tree canopy regenerated. Soil properties were also little affected. This led to the conclusion that many of the effects on the forest of the 1938 hurricane were most likely caused by widespread and intensive salvage logging. Previously, ignoring the importance of historical human impacts on forest recovery after major damages had led to misinterpretation of the process.12 There are two major messages from this example. First, using history to interpret ecological processes requires detailed analysis of the historical context of a past event. Second, changing drivers of change over time may affect processes. Lack of heavy machinery or of a market for timber would have led to a different outcome of this major disturbance. Cultural drivers of change affected the patterns of secondary succession. At a regional scale, multiproxy studies have made it increasingly clear that human impacts throughout the Holocene have not occurred against a backdrop of a stable climate. Many features of the landscape, such as patterns of grass-dominated ecosystems or frequency of fires, have changed in response to changing climate. People have most likely at times enhanced these changes, for example, by increasing fire ignitions at the same time that climate was becoming more conducive to fire. This is the case in many forested and grassland ecosystems, where nonanthropogenic fire has long been a component of the landscape.13 Historical ecology also provides insight into the process of extinction. The rates and patterns of this time-transgressive process can be determined only by long-term datasets.14 Especially long-lived species may persist long after conditions have changed, so that reproduction will be limited or eliminated, eventually causing extinction. Immigration is often delayed as well, as populations build slowly until they reach an exponential rate of increase. Species

224  Conclusion

ranges have also expanded rapidly when conditions have changed rapidly. Historical datasets with multiple indicators of both forcing factors, such as climate change or community fragmentation, and species are needed to anticipate the lag times for the future. For example, extinction of the Pleistocene megafauna in the Americas and Australia and of thylacines and devils in Australia are correlated with the arrival of Homo sapiens for the megafauna and the dingo for the thylacines and devils. The temptation is to infer causation from these correlations. However, more precise dating, modeling, and independent climate evidence are suggesting that causation in both cases is more complex, involving climate change in the first case and human land clearance in the second. Historical research reveals complexities of interactions in extinctions that are more difficult to establish in the short term. At a regional scale, the overall similarity of temperate forests in eastern Asia, eastern North America, and western Europe suggests that there are overarching physiological and evolutionary factors responsible for determining these patterns. These regions have very different histories of disturbance but similar vegetation because of climate, soils, and adaptations of deciduous trees. The diversity of the communities is very different, being much higher in Asia than in North America and higher in North America than in Europe, largely for reasons of evolution and glacial history. These factors, however, go only partway in explaining the differences in the systems. The effects of the variable lengths and intensities of human habitation on the three continents have had major consequences for the landscape, from the human-­dominated landscapes of Asia and Europe to the less obvious impacts in much of precolonial eastern North America. These differences have made themselves felt in the development of ecological studies in the different regions: in North America and northern Europe, ecologists until very recently focused on trying to understand ecological processes in the absence of human impact. In central and southern Europe, on the other hand, human impact was assumed to be universal. Thus not only the kinds of systems that we study but even the factors that we consider to be important are influenced by the past. Historical studies of past vegetation have indicated that most systems are in flux; as climate has changed, species distributions have also changed, with different species tracking climate change with different lag times and interactions.15 Periods of relative stability have been interspersed with periods of rapid change, sometimes referred to as times of disassembly and reassembly of communities, when the climate is changing more rapidly than species have adjusted. Although on a scale of several centuries and thousands of square kilometers the species may track climate closely, at the finer scales of time and space in which most ecologists work and species interactions take place, this

Toward the Future  225

equilibrium may not be apparent.16 For example, the overall patterns of forest dominants have remained stable in northeastern North America over the last several millennia even after a variety of human disturbances, suggesting that these patterns are robust and not just chance assemblages of species that are in flux. Similarly, a spatially detailed study of the prairie forest border in north-central North America before clearing by European settlers showed that the most important factor in the location of the border was climate, specifically moisture. In both regions, however, patterns at a finer scale were more complex, related to such features as topography and anthropogenic disturbance.17 It is routine to include physical attributes and geological history of a study site in analyses, and if such studies are to yield results of use in detecting underlying consistent patterns, it should become routine to include site history and human impact as well. That there is a legacy from past human activities, and that this legacy affects species distributions and landscape patterns in specific ways is becoming ever more apparent.18

Incorporating Historical Research into Conservation Why do people commit money and effort to preserving natural areas, interfering with conventional, economically remunerative uses of these lands? The answers to this question are deeply rooted in human perceptions of their role in the natural world. In many Western traditions, from which much of the modern industrial world developed, the nonhuman world is distinctly other, to be used, perfected, or otherwise manipulated by people. At the same time that it is crude, uncivilized, and dangerous, it is also pure and an example of a creation untainted by human sin and corruption. This ambivalence is reflected in the contrast between destructive manipulation of nature for human use and the attempt to rearrange it to emphasize its beauty. Running through these attitudes, however, is the continuous thread of thought that nature is there for the use of humans and that they have the right and responsibility to use and perfect it.19 Those who modify the natural world are usually called “developers,” not “destroyers.” Conservation can be seen as a form of development with the goal of protecting or recapturing the good qualities of nature. Some land has been preserved for saving rare species, some for protecting water supplies, some for propagating wildlife for hunting, some just for people’s enjoyment of open, nonbuilt spaces. The goal has dictated the type of management, but both goals and ways to achieve them change through time. In the mid-twentieth century, most “managers” of natural areas in North America would probably

226  Conclusion

have said that their main job was to prevent human intervention in the systems that they managed, allowing nature to take its course. They anticipated that this would lead to a natural system, in essence a climax community, a “wilderness,” ignoring research indicating that ecosystems would always be in flux with changing conditions, that there was no “climax.”20 Some assumed that the reason the systems were not developing in the desired direction was that prehistoric human disturbance had been removed; replacing it, especially in terms of fire, would put the site back on the path toward the goal, which was prehistoric systems, not systems with no human influence. Additional “active management,” a form of human disturbance, was necessary to correct changes wrought by more recent current changes, such as the spread of non-native species. By imposing such a new, directed disturbance regime on the system, one could reestablish and maintain prehistoric ecosystems, though now not strictly “natural” because of the need for human interaction. Although there was the hope that eventually one could erase the problems of the past and of the present surroundings and thus no longer have to manage in such an active way, the aim as a rule was not to remove management but to establish the ideal system. Changing conditions over the past centuries continued to exert influences that seemed not to be easily altered by ignoring them. This approach appears to use the historical record for analogues while ignoring history as prologue to the present, except as an impediment to reaching a goal. In parts of Europe, on the other hand, the most diverse and interesting communities often result from centuries, if not millennia, of human actions. In the last century, economic benefits from cultural systems such as coppicing and seasonal grazing declined, and for this reason and others the centuries-old disturbances ceased in many areas and were intensified in others. The systems that developed in the absence of human impact were less diverse than those that had been there and did not have the appeal for which they were being preserved, so active management had to be reinstituted.21 The historical intervention had to be continued to maintain the features of interest. We strive to preserve rare species, which may exist only in a very few sites, while we have not found evidence that the extinction of such major species as passenger pigeons, moas, or chestnut trees in the forest canopy has had major ecological consequences. (Which does not mean that these extinctions have not had such consequences, just that we have not yet clearly quantified them.) Can historical research indicate the consequences of these losses? Or is our interest in preserving rare species (apart from their possible individual value in supplying people with unique products) more a matter of values,

Toward the Future  227

ethics, and aesthetics? In understanding the world around us, do they contribute to a more satisfying story? 22 One all-encompassing concept of the value of ecological systems is that of ecosystem services. Some of these are commodity-based, for example, providing food, timber, and waterpower, while others relate to cultural and intellectual values. The rationale for preserving more or less intact ecosystems is that they provide services that can be specified and given value. There are several widely accepted ecosystem services that relate specifically to conservation, such as prevention of soil erosion, sequestration of carbon, protection of a diversity of species, and provision of areas for enjoyment and study. The major problems with using these for defending conservation values are quantification for comparison with other land use and providing spatially and temporally explicit values.23 Calling these “ecosystem services” is fairly new, but recognizing their value is not. For example, the report by the New Jersey state geologist in 1900 recognized the value of the forests for protecting water supplies, moderating climate, providing outdoor laboratories and classrooms for studying plants in their natural habitats, and preserving species, as well their utility as a harvestable resource. The state geologist did, however, acknowledge that protecting these resources would depend on the state, that is, regulations, because economic incentives would lead to other, destructive, uses.24 Values placed on ecosystems change over time for technological, cultural, political, and other reasons. A case study in Switzerland showed that, since 1900, some ecosystem services, such as the value of litter collected for fodder, have gone from being noted as very important to having no significance, while others, such as carbon sequestration and maintaining biodiversity, are only recently recognized.25 While there may be conflict between valuation of services for subsistence rather than for conservation, a retrospective study of several projects shows that this conflict can be resolved if local political and cultural conditions are taken into account as well as technology.26 To be effective for conservation, ecosystem values must be integrated into changing local values and cultural conditions. Often abandoning an agricultural practice, such as grazing, may be an economically favorable decision but may significantly decrease biodiversity and values that accompany it, such as tourism or education. On the other hand, intensifying agriculture may bring short-term economic gains while losing the potential for other values in the future. There is a long history of preserving remnants of natural systems, from ancient sacred groves to contemporary nature preserves and wilderness areas. Both the sacred groves and nature preserves are islands in a sea of cultivated

228  Conclusion

or built landscapes. In small islands such as Chinese naturalistic gardens, careful design creates landscapes in which every point of view gives a naturalistic impression. In large preserves, such as extensive wilderness areas, the aim has been to avoid any evidence of human intervention.27 The goal of an urban woodlot may be simply to provide urban residents with a glimpse of a system with minimal human disturbance, while that of a nature preserve may be to maintain the conditions needed for the survival of one or a suite of rare species. Sacred groves, gardens, and urban woodlots are often surrounded by intensively used land and often were set aside for protection while the land around them was being built. The very fact that they were preserved presumes that conflicting uses are occurring around them such that current conditions will not continue unchanged. In addition, the past always includes some human impact that has had some effects. It is not necessary to regard all human impacts as negative.28 Management decisions that ignore past land use both of the preserves and of surrounding areas are apt to be derailed by unexpected residual impacts of the past as well as by impacts caused by changes in the future.29 Historical study incorporated into conservation planning may provide surprising results. In the San Francisco Bay Area in California, extensive work in historical archives has allowed reconstruction on paper of the Bay Area before the major development of the last century. Careful analysis of multiple lines of evidence has revealed a past landscape of extensive tule (Scirpus lacustris) swamps and tidal marsh as well as riparian forests, most of which were unknown before this research. The diversity, locations, and composition of these plant communities are informing conservation efforts in ways that would not have been possible without the historical study. Finding that conifers were an important vegetation type in central Europe or that silver fir would most likely be able to flourish in Tuscany also leads to changed conservation agendas.30 Conservation decisions taken without regard to the history of a region are very often mistaken. Based on the outdated concept of “climax vegetation,” they often ignore sustainable change in the past and attribute current conditions that do not fit the theory to untested historical explanations. These historical explanations for current patterns and processes may not be the best templates for conservation or conservation activities.31

Global Change The concept of global change in the twenty-first century focuses on diverse human-mediated changes that affect the entire globe, among them

Toward the Future  229

increases in carbon dioxide, pesticides, and other anthropogenic chemicals in the atmosphere, decreases in stratospheric ozone, and a decrease in the number of species present on the earth.32 While it occurs against a background of nonhuman-mediated shifts in climate and evolution, recently accelerated change seems to overwhelm these and has led to a proposed new geologic epoch, the Anthropocene.33 At least three scientific journals have appeared in the twenty-first century focused on the Anthropocene, and the term is frequently used in papers published elsewhere.34 There is, however, no general agreement on the specific beginning of this epoch or even on the value of the concept.35 Arguments are based on paleoecological and historical evidence and offer differing interpretations. Support for the designation notes the urgency of concerns about current issues, while criticism focuses on the importance of considering the long history of human impacts, such as deforestation for agriculture, changing fire regimes, and extinction of the Pleistocene megafauna. Use of the term, however, establishes an artificial divide in understanding human interactions with the nonhuman environment, likely obscuring a better understanding of the temporal and spatial complexity of these relationships.36 Most global deforestation in temperate regions of the globe occurred millennia ago. Most regions of South and Central America that were deforested millennia ago have reverted to highly diverse tropical forest. In much of Asia, Africa, and Europe, however, agricultural practices continued for millennia have led to diverse systems that have replaced previous forests, savannas, or grasslands. Historical anthropogenic impacts are deeply embedded in all of these landscapes, even when these antecedents are not immediately apparent, and are very widespread. While concentrating on recent changes is important, it is also critical that they be put into their historical context to avoid misinterpreting the causes for the changes seen in the last century. Drivers of environmental change include culturally conditioned attitudes.37 In 1967, Lynn White argued in his provocative article “The Historical Roots of Our Ecologic Crisis” that we must seriously weigh the fundamental causes of our ecologic problems if we are to avoid making serious errors in trying to solve them. According to White, the roots of the problems lie in modern Western science and technology, which have developed from the Christian dogma that people have the right to assert their mastery over nature. His focus on the historical roots of perceived contemporary problems is cogent, although the extent of ecological problems in Eastern landscapes belies his emphasis on the crucial role of Christianity. In the past, much human-mediated change was regarded as good. Humans tamed the wilderness, irrigated barren deserts, reclaimed malarial marshes—

230  Conclusion

accomplishments that marked the advance of human civilization. In Australia in the mid-nineteenth century, lands that European settlers were not using were referred to as “wastelands” that should be “rescued from a state of nature.” Now, the protection of remaining wilderness, deserts, and marshes is a keystone of the wise use of land. This emerging paradigm, which values the nonbuilt environment, is in conflict with the old one, which sees no problems with using nature mainly to satisfy concrete human needs. Some see an inherent conflict between the pursuit of increased productivity and the preservation of nature. Historical analyses show, however, that the effects of the pursuit of productivity can destroy the basis of that productivity when the pursuit is heedless of historical evidence of earlier damage. Drained marshes may increase the amount of land available for hay or building and decrease the habitat of some disease-causing organisms, but eventually they lead to decreased productivity of adjacent waters and to increasing floods, which threaten the landscapes built on the drained land. History holds evidence of good and bad reclamation, and these lessons can direct the future. History can also show where current problems are exacerbated by past “improvements.” Destroying mangrove forests along coasts to make the area more economically valuable has led to devastating flooding in many parts of the world. The lessons learned from the clearing of forest for farms on poor soils in the Great Lakes area of North America could be applied to similar efforts in the tropics.38 General concern about global change must be linked to specific issues, which are in turn determined not only by the actual changes but also by perceptions, such as whether the change is perceived as natural or human-caused. Our perception of the major environmental problems has shifted over the past few decades, from pesticide residues to acid precipitation to global warming, invasive species, and losses of biodiversity. Other critical preoccupations will undoubtedly appear in the future. Solutions to one problem may not apply to others, and the old problems usually do not disappear as attention shifts elsewhere. Historical research may reveal overall patterns that merit study, for example, the effect of human modifications of ecosystems by whatever means. People simplify landscapes by replacing diverse ecosystems with cultivated fields, but in terms of landscape patchiness, development of new successional pathways, and increase in species numbers, human activities have made ecosystems more complex. Perhaps human activities render ecosystems less coherent—that is, the more human actions change them, the farther removed they are from a position in which they can remain somewhat stable in the face of stresses. The specific human activity may not be as important as the extent to which it diverts an ecosystem from some preexisting

Toward the Future  231

state.39 The amount of energy necessary to maintain “natural” systems in nature preserves alone implies that the systems have been so diverted. Even if we do not pass judgment on whether a speeded-up rate of change is good or bad, it does appear to be unprecedented in the evolutionary history of life on earth. By considering the long sweep of history we are able to place what we are doing in this context and to evaluate some potential consequences. It already appears that modifying the physical substratum and introducing species have been especially destructive to the ability of systems to recover after disturbance, at least on the scale of human history. Will this also be true of modifying climate and the atmosphere? Historical ecology has contributed to predicting the consequences of global change mainly by providing evidence of the responses of systems to change in the past. In addition, features of past adaptations can provide information critical to establishing effective conservation strategies. Focusing only on current features of landscapes, species, and ecosystem interactions ignores the lessons that can be learned from past adaptations, such as rates of migration. Historical studies provide information that will avoid pitfalls of solutions based on short-term research. In linking ecology and history, historical ecologists are gaining insights into the importance of a temporal perspective in ecology and the integral importance of human interactions in ecological systems over time. There are several broad conclusions that can be drawn. First, all landscapes are to some extent historical, cultural landscapes. The features related to cultural changes over time have been massive, though the rates may have been slower than today. Second, residual features that are not apparent in contemporary systems may continue to influence them cryptically. These may be the causes of patterns today that cannot be interpreted solely on the basis of currently measurable driving forces. This is generally attributable to some past alteration of the substrate or species composition that can often be discerned with precision and accuracy using current technology. Third, historical processes provide clues to causes and consequences of changes occurring today, and thus can be constructively incorporated into management decisions. Historical ecology is an observational science.40 Establishing causation with assurance is difficult, but hypotheses are effectively tested by finding information that may support or refute them. This process is a continuing one. New information and modeling lead to revising some previously held views on the causes of events in the past, such as the extinction of the Pleistocene megafauna or the role of human-set fire in maintaining some ecosystems. With continued research, historical ecologists will continue to strengthen their understanding of causation of past events to contribute to understanding processes occurring today.

This page intentionally left blank

Notes

Preface to the Second Edition   1. Crumley, Lennartsson and Westin 2018.  2. Williams 1989, 393–424; Bignal and McCracken 1996; Fairhead and Leach 2014.

Chapter 1. History Hidden in the Landscape   1. Richards 1973, 24; Willis, Gillson and Brncic 2004; Leal et al. 2016.   2. Three recent books include “historical ecology” in their titles. Egan and ­Howell’s Handbook of Historical Ecology (2005) focuses specifically on using historical studies for restoration ecology. Crumley’s Historical Ecology: Cultural Knowledge and Changing Landscapes (1994), Crumley, Lennartsson and Westin’s Issues and Concepts in Historical Ecology (2018), and Balée’s Advances in Historical Ecology (2002) treat the field from a complementary anthropological perspective, emphasizing the human element rather than ecosystem processes.   3. Bellemare, Motzkin and Foster 2002; Krzywinski, O’Connell and Küster 2009; Josefsson et al. 2010; Freschet et al. 2014; Reitalu, Kuneš and Giesecke 2014.   4. Utterström 1988; Williams 1994; Jackson et al. 2001; Lotze and McClenachan 2013.   5. Dambrine et al. 2007; Thompson et al. 2015.   6. Redmon 1999; Swetnam, Allen and Betancourt 1999; Foster et al. 2003; Rhem­ tulla and Mladenoff 2007.

233

234  Notes to Pages 5–20   7. Somerville 1853; quotation attributed by Marsh 1864 to H. Bushness.   8. Lowdermilk 1975; McEvoy 1986, 63.   9. Vankat 1979; Schlesinger et al. 1990; Cuddihy and Stone 1990. 10. Trimble 1992. 11. Bousquet and Fleming 2017. 12. Marks 1983; Terborgh 1989; Vickery and Herkert 1999. 13. Bradshaw et al. 2015. 14. Szabó et al. 2016. 15. Bland et al. 2018. 16. Anonymous 1992. 17. National Research Council 1992. 18. Szabó 2010. 19. Iverson 1988; Benjamin 2007. 20. Bürgi et al. 2015. 21. Hough 1878; Little 1950; Ellsworth 1975; Davis and Webb 1975; Bernabo and Webb 1977; Goodman and Lancaster 1990. 22. Anonymous 1926; Ellsworth 1975; Williams 1994. 23. Walsh 1967, 31; Cronon 1983; Bailes 1985b; Opie 1985; Worster 1988; Crosby 1995; Isenberg 2014; Szabó 2015. 24. Sauer 1941; Eyre and Jones 1966, 24; Turner et al. 1990; Knowles 2014. 25. Crumley 1994; Balée 2002. 26. Clements 1936; Küchler 1964; http://usnvc.org. 27. Council Directive 92/43/EEC of 21 May 1992, Article 1b. 28. Tansley 1923, 23; Rackham 1980. 29. Hutchinson 1957; Eriksson 2013; Feeley 2015. 30. Likens and Davis 1975. 31. Thomas 1956; Hamburg and Sanford 1986; Christensen 1989; Turner et al. 1990. 32. Nicholas 1988; Christensen 1989. 33. Davis et al. 1986. 34. Deevey 1969. 35. Agnoletti and Anderson 2000.

Chapter 2. Historical Records and Collections   1. Van Houtan, McClenachan and Kittinger 2013; Rogers 2016.   2. Bürgi, Hersperger and Schneeberger 2004.   3. Sanitary and Topographical Map of Hudson County, N.J. 1881; Geological Survey of New Jersey, 1900b.   4. Olson 1971; Robbins 1976.   5. Sinclair 2012.   6. Bloch 1953; Benjamin 2007.   7. van der Donck 1650, 276–277, 1854, 145–146.   8. Loeb 1982; Newton et al. 2009; Vellend et al. 2013.   9. Bloch 1967, 48. 10. A point made by Richard White 1989.

Notes to Pages 21–37  235 11. Vermeule 1900, 13–108; Black 1950, chap. 4. 12. Shapiro and Swain 1983. 13. Dobbs 1755. 14. Marsh 1882. 15. Nelson 1894, 11:280–281. 16. Forman and Russell 1983. 17. Darby 1976, 52–53. 18. Kreisler 1984, 37–39. 19. Kemble 1780–1785. 20. Guthorn 1966, 1972; Wilkinson 1777. 21. Russell 1988a. 22. Klett et al. 1984; Humphrey 1987; Dunwiddie 1989. 23. Chen et al. 2011. 24. Avery 1977. 25. Cousins 2001. 26. Lunt and Spooner 2005. 27. Clawson and Stewart 1965; Edmonds 2005; Whitney and DeCant 2005. 28. Rogers 1823. 29. Howell 1828–1831. 30. Ingvild 1988. 31. Kercheval 1833. 32. Southgate, unpublished research. 33. Ludwin 2005. 34. Leopold 1933; Wacker 1975. 35. Daube 1973. 36. Crawford, Lips and Bermingham 2010; Drábková 2014. 37. Thuiller et al. 2005; Bertin 2013. 38. Bond, Hobson and Branfireun 2015; Turner et al. 2015. 39. Rogers 2016. 40. Tulowiecki 2018. 41. For example, Rackham (1986) used primarily historical documents to show changes in the vegetation of the British Isles through the Middle Ages; Whitney (1994) compiled a compendium of forest changes for parts of North America.

Chapter 3. Field Studies   1. Rhemtulla, Mladenoff and Clayton 2007; Bonnell and Fortin 2014; Gimmi et al. 2016.   2. Emerson and Flint 1862; Schmidt 1946; Curtis 1956; Burgess and Sharpe 1981b; Ranney, Bruner and Levenson 1981.   3. Bard 1952.   4. Russell and Schuyler 1988.   5. Rackham 1980; Russell 1993a.   6. Knowles and Hillier 2008; Lillesand, Kiefer and Chipman 2015.   7. Georges-Leroy et al. 2009; Chase et al. 2011.

236  Notes to Pages 37–52   8. Martin and Bass 1940; Limbrey 1975; Lowdermilk 1975; Buol, Hole and McCracken 1980, 31–36; Olson 1981, 105–108, 113–114; Kain and Hooke 1982; Sanford et al. 1985; Clark, Merkt and Muller 1989; Trimble 2012; Thierry Dutoit, personal communication.   9. Nye et al. 2018. 10. Webb et al. 2007. 11. Scheel-Ybert et al. 2003; Nelle 2003; Dutoit et al. 2009; Robin et al. 2018. 12. Bridges 1970, 51; Birks et al. 1976; Krug and Frink 1983; Davis 1987; Goudie 1990, 130–131. 13. Binford et al. 1987. 14. Bamforth and Grund 2012; Contreras and Meadows 2014. 15. Wigston 1993. 16. Clark 1995. 17. Rackham 1986. 18. Poschlod et al. 2008. 19. Rackham 1980; C. D. Piggott, personal communication 1985; Russell 1988b. 20. Verheyen et al. 2003. 21. Swan and Swan 1974; Oliver and Stephens 1977. 22. McCaughey 1982; Lorimer 1985; Fritts 1991; Kipfmueller and Swetnam 2005; Cook et al. 2010; Cook et al. 2015. 23. Carloni 2005. 24. Oosting 1942; Bard 1952; Pickett 1989. 25. Hall 1905; Moss et al. 2004. 26. Frye 1978; Pickett 1983; Fraterrigo, Balser and Turner 2006; Meiners 2007. 27. Harmer et al. 2001. 28. https://lternet.edu/; https://www.ilter.network/. 29. Taylor 1989. 30. Ramey-Gassert and Runkle 1992. 31. Verheyen et al. 2018. 32. Vellend et al. 2007. 33. Allstadt et al. 2015. 34. Wares and Cunningham 2001.

Chapter 4. The Sedimentary Record   1. Davis 1994.   2. Lowe 2011.   3. Smart and Frances 1991; Kirchner 2011; Reimer et al. 2013.   4. Van Zant et al. 1979; A. M. Davis 1984.   5. Engstrom and Wright 1984; Brush and Davis 1984; Kemp et al. 2005.   6. Deevey 1984; Davis 1987.   7. Smol and Stoermer 2010.   8. Watts 1967; Birks 1973; Elias 1994; Willerslev et al. 2003; Giguet-Covex et al. 2014.   9. Anderson, Goudie and Parker 2013; Green and Speller 2017; Girona, Navarro and Morin 2018.

Notes to Pages 52–67  237 10. Polyak et al. 2001; Rhode 2005; Jouy-Avantin et al. 2003; https://www.neoto madb.org. 11. Details of sediment sampling, processing, and analysis can be found in such textbooks as Kummel and Raup 1965; Berglund 1986; Faegri, Kaland and Krzywinski 1989; Moore, Webb and Collinson 1991; Piperno 2006. 12. Davis and Deevey 1964; Smol et al. 1986; ter Braak and van Dame 1989; Giesecke and Fontana 2008; Bennett and Buck 2016. 13. Bradshaw 2016. 14. Pielou 1974; Everitt et al. 2011. 15. Webb 1973, 1974; Jacobson and Bradshaw 1981; Heide and Bradshaw 1982; Birks and Gordon 1985; Russell and Davis 2001. 16. Overpeck, Webb and Prentice 1985; Nakagawa et al. 2002; Gavin et al. 2003; Jackson and Williams 2004; Fyfe 2006. 17. Huntley and Birks 1983; Davis 1983; Williams et al. 2004; Ren 2007; Brewer et al. 2017. 18. www.ncdc.noaa.gov/data-access/paleoclimatology-data/datasets. 19. Sugita 2007a, 2007b. 20. Bunting and Middleton 2009; Bunting et al. 2018. 21. Henne et al. 2013. 22. Subcommission on Quaternary Stratigraphy 2018. 23. Faegri, Kaland and Krzywinski 1989, 177–199; Peteet 1995; González-Guarda et al. 2017; Finsinger et al. 2017. 24. Voosen 2018. 25. Firbas 1937; Iversen 1941; Birks 1986; Leal et al. 2016; Daura et al. 2016. 26. Russell et al. 1993. 27. Campbell and McAndrews 1993. 28. Nielsen and Odgaard 2010. 29. Bunting et al. 2018. 30. Bayon et al. 2012; Bostoen et al. 2015. 31. McGovern et al. 2007. 32. Williams et al. 2004.

Chapter 5. Fire   1. Bond, Woodward and Midgley 2005; Whitlock et al. 2010; Marlon et al. 2013.   2. Pyne 2001; fireresearchinstitute.org.   3. Noss et al. 2015 (data from the National Fire Detection Network, owned by Vaisala).   4. Taylor 1974; Goudie 1990; http://www.nifc.gov/nicc/predictive/intelligence/2013 _Statssumm/wildfire_charts_tables13.pdf.   5. Mutch 1970.   6. Keeley 1986; Gill, Hoare and Cheney 1990.   7. Schuyler and Stasz 1985.  8. Pyne 2015, 103; https://www.fs.fed.us/database/feis/plants/tree/pintae/all.html# FIRE ECOLOGY.

238  Notes to Pages 68–79   9. Retallack, Dugas and Bestland 1990; Bond, Midgley and Woodward 2003; Keeley and Rundel 2005; Bowman et al. 2011; Crisp et al. 2011. 10. Heinselman 1981; Christensen 1981. 11. Karkanas 2007. 12. Hough 1926; Stewart 1956. 13. Russell et al. 1993. 14. Loope 1991; Clark and Royall 1995; Whitlock et al. 2010; Leal et al. 2016. 15. Carloni 2005. 16. Patterson and Sassaman 1988. 17. Maezumi et al. 2017. 18. Innes and Blackford 2003. 19. Woeikof 1901; Day 1953; Sauer 1966. 20. Carver 1796, 187; Eldredge 1909; Blane 1918, 74–76; Lindeström 1925; Williams 1974, 5; Lewis 1982. 21. Pyne 1991, 61. 22. Acocks 1953. 23. Bond, Midgley and Woodward 2003; Vorontsova et al. 2016. 24. Chittendon 1905, 4:347; Gleason 1913; Blane 1918. 25. McInteer 1946; Wilkins et al. 1991; Baskin, Baskin and Chester 1994; Rhoades, Miller and Shea 2004. 26. Russell 1983; Foster et al. 2002a. 27. Brown and Davis 1973, 10; Stewart 1956, 128. 28. Huntley and Webb 1988; Marlon et al. 2013. 29. Russell et al. 1993. 30. Oswald et al. 2007. 31. http://inciweb.nwcg.gov/incident/4689/. 32. Birks 1986; Clark, Merkt and Muller 1989; Innes and Blackford 2003; Marlon et al. 2013. 33. Boyden 1987. 34. Overpeck, Rind and Goldberg 1990; Bond, Midgley and Woodward 2003; Bowman et al. 2009; Noss et al. 2015. 35. Power et al. 2008; Marlon et al. 2013. 36. Kerkkonen 1959; Saarnisto 1986; Niklasson and Granström 2000. 37. Whitehead 1881, 145–147; Hough 1882, 3:130–155. 38. Buell, Buell and Small 1954. 39. New Jersey State Geologist 1903, 51. 40. Brose et al. 2001. 41. van der Donck 1841, 20–21; Dwight 1969, 38–40; Russell 1983; Pyne 2015. 42. Shaler 1884, 29. 43. Maxwell 1910, 73. 44. Hough 1882, 206–207; Sarvis 1993. 45. Wright and Heinselman 1973. 46. Hough 1882, 60, 206–207. This report includes extensive detail about the fires that were reported as well as a summary of laws regulating fires in the various states. 47. Pinchot 1900, 108.

Notes to Pages 79–93  239 48. Clements 1916. 49. Brown 1949, 477–479; Forman and Boerner 1981. 50. Hartman 1949. 51. Hammatt 1949, 479. 52. Niering and Goodwin 1962; Ohmann and Buell 1968. 53. Brown 1960; Niering, Goodwin and Taylor 1970; Chandler et al. 1983; Bork, Hudson and Bailey 1997; Franklin, Robertson and Fralish 2003; Signell and Abrams 2008. 54. Parsons et al. 1986; Christensen et al. 1989. 55. Beatty and Taylor 2008. 56. Cuddihy and Stone 1990; Levine et al. 2003. 57. Parsons 1994; Power et al. 2018. 58. Williams 1989. 59. Zumbrunnen et al. 2009.

Chapter 6. Extending Species’ Ranges   1. Preston, Pearman and Hall 2004.   2. Martin, Canham and Marks 2009; Beauséjour et al. 2015.   3. Darwin 1966, 380–381.   4. Elton 1958; Groves and Burden 1986; Mooney and Drake 1986; Drake et al. 1989; Simberloff 2013.   5. http://www.merriam-webster.com/dictionary/invade.   6. Elton 1958, 33.   7. Davis 1983; Davis et al. 1986; Webb 1988; Scheller and Mladenoff 2005; Hu, Hampe and Petit 2009.   8. Orians 1986; Marshall 1988; Hoorn and Flantua 2015.   9. Snow 1980; Dorney 1981. 10. Turner 1970, Kornas 1983; Mack and Lonsdale 2001. 11. Bates 1956; Goudie 1990; Buisson, Dutoit and Wolff 2004. 12. Corbet 1974; Darby 1976, 98; Huntley and Birks 1983; Sheail 1984; Behre 1988. 13. Elton 1958, 118; Garrard and Streeter 1983, 248. 14. Crosby 1972. 15. Crosby 1972, 175; Lever 1985, 401–406; Wingate 1990; Cuddihy and Stone 1990, 40. 16. Schmidt 1973; Russell 1979; Van Zant et al. 1979. 17. Collinson 1738. 18. Bartram 1760. 19. Martin 1965. 20. Moulton et al. 2010. 21. von Wangenheim 1781; Landa 1988. 22. Gras 1930, 18; Power 1941, 47, 53. 23. Clark 1956; Densmore 1974; Van Zant et al. 1979; Crosby 1986, 155; Rackham 1986, chap.4. 24. Darlington 1969, 652; Wood 1987.

240  Notes to Pages 94–108 25. Schmid et al. 2015. 26. Athens et al. 2002; Hunt and Lipo 2007; Clout and Russell 2008. 27. Gurevitch and Padilla 2004; Ricciardi 2004; Simberloff 2013. 28. Orians 1986; Cuddihy and Stone 1990, 50. 29. Cuddihy and Stone 1990, 76–77, 84–85. 30. McClintock and Fitter 1964, 91; Polunin and Smythies 1973, 177; Garrard and Streeter 1983, 243. 31. Krzywinski, O’Connell and Küster 2009. 32. Elton 1958, 73; Dean and Voss 1965; Long 2003, 121–127. 33. Packard 1942; Clark 1956; Anonymous 1976; Lever 1985, 193–203; Caughley 1985, 15–21; Ratcliffe 1989; Mark and McSweeney 1990; Goudie 1990, 106; Wagner and Kay 1993; Fortin et al. 2005. 34. Waugh 1973; Townsend 2003; www.newzealand.com/us/fly-fishing/. 35. Moyle 1986; Simberloff 2013, 131–132. 36. Elton 1958; Moyle 1986; Kitchell and Carpenter 1993; Simberloff 2013, 36–39. 37. Elton 1958, 75–76; Moyle 1986; Hadfield and Miller 1992; Nichols et al. 2008; Simberloff 2013, 134. 38. Good 1965; Russell et al. 1993; Sinclair 2012. 39. Britton 1881; Gleason and Cronquist 1991; Simberloff 2013. 40. Orians 1986; Gleason and Cronquist 1991, 64. 41. Cuddihy and Stone 1990; Ainsworth and Kaufman 2010. 42. Tunison, D’Antonio and Loh 2001; Vitousek, D’Antonio and Asner 2011. 43. Mack 1981; West 1988; Billings 1990; Fritts 1991. 44. Sperry, Belnap and Evans 2006. 45. Bard 1952. 46. Kueffer 2017. 47. Haleakalâ 1990. 48. MacDougall and Turkington 2005; Thomas and Palmer 2015. 49. Wardle et al. 2016.

Chapter 7. Harvesting Natural Resources   1. Sauer 1969, 10–11; Anderson 1990; Krech 1998; Lotze and McClenachan 2013; Goudie 2013; Snir et al. 2015.   2. Devèze 1966; Anderson 1990.   3. Hempson, Archibald and Bond 2015.   4. Graham 2016.   5. Martin 1984.   6. Martin 1967; Martin 1984.   7. Seersholm et al. 2018.   8. Dragoo 1976.   9. Sutcliffe 1985, 206; Lorenzen et al. 2011. 10. Sutcliffe 1985, 181–185, 199; Haile et al. 2009; Barnosky et al. 2010; Prowse et al. 2014; Cooper et al. 2015; Halligan et al. 2016. 11. Robinson, Burney and Burney 2005; Ellison et al. 2005; Gill et al. 2009.

Notes to Pages 108–119  241 12. Erlandson 2001. 13. Erlandson et al. 2011; Rick et al. 2015. 14. Jackson et al. 2001. 15. Heizer 1955; Driver 1969; Densmore 1974; Simmons 1989, 59–60; Scarry 1993; Dunham 2009; Tushingham and Bettinger 2013. 16. Kalm 1937; Nabhan and Anderson 1941; Anderson 2005; Chazdon 2014, 16–21. 17. Yarnell 1964; Boyden 1987, 81n26; Cordain et al. 2000; Milton 2000; Johnson 2014. 18. Smith 1970; Chambers and Elliott 1989; Hörnberg et al. 2006; Ryan and Blackford 2010; Rautio, Josefsson and Östlund 2014. 19. Fairhead and Leach 2014. 20. Perlin 1989. 21. Rackham 1980. 22. Flemley 1979; McAndrews 1988; Sponsel 1992; Russell et al. 1993; Gaillard et al. 2009; Chazdon 2014, 21–32. 23. Betancourt, Dean and Hull 1986; Delcourt 1987. 24. Darby 1976, 52–53; Rackham 1986, 16. 25. Szabó 2013. 26. Old Testament, 2 Chronicles 2:8, 18; Lowdermilk 1975; Rollefson 1990. The sources differ in identification of the trees used. Some include cypress and algum (?) (also translated as almuggim), but the main point is that there were many trees, primarily gymnosperms. 27. Fairhead and Leach 2014. 28. Sponsel 1992; Horn 1992. 29. Lambert 1971, 49–50; Darby 1976, 24, 167; Hooke 1988, 301–310. 30. Hughes and Huntley 1988; Austad 1988; Gimmi et al. 2013; Watkins 2014, 120–128. 31. Schmidt 1973, 53, 66, 73; Russell 1979. 32. Peterken 1976; Rackham 1980; Kirby 1988; Rackham 1988. 33. Erlande-Brandenburg 1995; Williams 2000. 34. S. Webb 1986; Butler and Malanson 2005. 35. Leopold 1933, 420–423. 36. Marsh 1864, 240–244; Leopold 1933, 420–423; Devèze 1966; Duby 1968, 20; Bobiec 2002; http://whc.unesco.org/en/list/33. 37. Darby 1976, 55; Hooke 1988; Watkins 2014, 52–64, 140. 38. Devèze 1966; Watkins 2014, chap. 7. 39. Hough 1965; Anderson and Loucks 1979; McCabe and McCabe 1984; Flinn and Vellend 2005. 40. Fortin et al. 2005. 41. Bucher 1992; Ellsworth and McComb 2003; Hung et al. 2014. 42. Jackson et al. 2001; Erlandson and Rick 2010; Mánˇez et al. 2014; Poulson 2016. 43. Devèze 1966; Darby 1976, 65; Hooke 1988. 44. Devèze 1966; Darby 1976, 65, 221, 228–229; Métailié, Bonhote and Frauhauf 1988. 45. Piussi and Stiavelli 1988; Drescher-Schneider et al. 2007. 46. Braun 1950; Webb 1973; McAndrews 1988; Russell et al. 1993; Russell and Davis 2001.

242  Notes to Pages 119–133 47. Braun 1950, 338, 429; Whitney and Davis 1986. 48. Hanson 1992. 49. Lefébvre 1879; Hough 1882, 68–128; Darby 1976, 273. 50. Östlund et al. 2009; Rautio et al. 2015; Cramer 2015. 51. Bignal and McCracken 1996; Fairhead and Leach 2014. 52. Williams 1989, 393–424. 53. Sharpe et al. 1981; Pickett and White 1985; Forman and Godron 1986; Franklin and Forman 1987; Zipperer, Burgess and Nyland 1990; Spies, Ripple and Bradshaw 1994; Wallin, Swanson and Marks 1994. 54. Aubréville 1971. 55. Bourdo 1983. 56. McNeill 1988; Oliveira and Pillar 2004; Behling, Pillar and Bauermann 2005. 57. Hurst 1983; Ahlgren and Ahlgren 1983; Radeloff et al. 1999. 58. Gifford 1900, 292. 59. Devèze 1966. 60. Gifford 1900; Devèze 1966; Russell 1988a. 61. Evans 1992; Josefsson, Olsson and Östlund 2010. 62. Newton 1971, 530; Merchant 1985. 63. Vane-Wright 1993. 64. Walker and Chen 1987; Motzkin et al. 2004.

Chapter 8. Agriculture and Its Residual Effects   1. Allan 1965, quoted in Bigalke 1978, 13; Diamond 1987.   2. Hobbes 1909, 96.   3. Wolff, Tatin and Dutoit 2013; Normile 2016; Scott 2017.   4. Fuller 2010; Roberts 2013.   5. Anonymous 1961, 12; Harlan 1971; Cavalli-Sforza, Menozzi and Piazza 1993.   6. Sauer 1952, 21–22; Roosevelt 1980, 251–252; Fuller 2010.   7. Roberts 2013.   8. Garrod 1932; Isaac 1970, 61–62; Belfer-Cohen and Bar-Josef 2002; Snir et al. 2015.   9. Sauer 1952, 28, 32; Isaac 1970, chap. 5; Boyden 1987, 88–89; Larson and Fuller 2014. 10. Harlan 1971. 11. Olsen and Wendel 2013. 12. Smith 1989. In C3 plants, the first stable carbon compound formed by the fixation of carbon dioxide in photosynthesis includes three carbon atoms. The first carbon compound in C4 plants has four carbon atoms, hence the different descriptors. These processes use the stable carbon isotope C13 in slightly different ratios. 13. Firbas 1937; Iversen 1941; Moore and Chater 1969; McAndrews 1988; Clement and Horn 2001; Gaillard et al. 2009. 14. Iversen 1941; Iversen 1956. 15. Darlington 1969. 16. Curtis et al. 1998; Beach et al. 2002. 17. Chen et al. 2015.

Notes to Pages 134–151  243 18. Darlington 1969, 85–86; Isaac 1970, 45–46; Lowdermilk 1975; Simmons 1989, 93–97. 19. Renfrew 1979, 167–175. 20. Duby 1968, 109–110, 119; Boyden 1987, 91; Goudie 1990; Verhulst 1990, 17–28. 21. Fish 2000. 22. Curtis et al. 1998; Beach et al. 2002; Chase et al. 2011; Canuto et al. 2018. 23. Driver 1969, 64; Aikens 1983; McAndrews, 1988; Smith 1989; Grove and Rackham 2001; Hughes 2011. 24. Grigg 1982; Bucher 1988; Simonetti 1988. 25. McCallum and McCallum 1965; Cushman 2013. 26. Butzer 2005. 27. Marsh 1864, 117; Lowdermilk 1975. 28. Lowdermilk 1975; Hughes 2011. 29. Flemley 1979, 126; Denham, Haberle and Lentfer 2004. 30. Brugam 1978; Davis and Norton 1978; Brush and Davis 1984; Walter and Merritts 2008; Kirwan et al. 2011; Trimble 2012. 31. Williams 2000. 32. Richards and Tucker 1988; Williams 2000. 33. Russell and Schuyler 1988. 34. McDonnell and Stiles 1983; Southgate and Thompson 2014. 35. Evans 1975, 162. 36. Wilkinson 1777; Underwood et al. 1989; Austin 1992; Russell 1993a. 37. Binford et al. 1987; Hodell et al. 2001. 38. Evans 2018. 39. Caulfield 1981; Farrell and Doyle 2003; Caulfield et al. 2011; Guttmann-Bond et al. 2016. 40. Erlande-Brandenburg 1995. 41. Clout 1977, chap. 4. 42. Grigg 1982, 59. 43. Rhemtulla, Mladenoff and Clayton 2007. 44. Anderson 1960; Anonymous 1969; McDonnell and Stiles 1983. 45. Seymour 1970. 46. Marks 1983; Hughes and Huntley 1988; Naveh and Kutiel 1990. 47. Flemley 1979, 116–121; McAndrews 1988; Bush et al. 1992. 48. Birks 1986; Peglar 1993; Robinson 2000. 49. Russell 1993b; Bellemare, Motzkin and Foster 2002; Dupouey et al. 2002. 50. Naveh and Carmel 2004; Ruddiman, Kutzbach and Vavrus 2011; Singrayer et al. 2011; Mitchell et al. 2013. 51. Christensen 2014; Rull 2015; Veblen 2017.

Chapter 9. Patterns of Human Settlement and Industrialization   1. Cronon 1983.   2. Williams 1974, 67.   3. Lindborg and Eriksson 2004.

244  Notes to Pages 152–170   4. Wacker 1975, chap. 4.   5. Wacker 1975, 322–323.   6. Schmidt 1946, 72–73.   7. Wacker 1975, 245–246.   8. Russell 1979.   9. Dorney 1983. 10. Robbins 1976; Opie 1987; Meinig 1993, 404–405. 11. Robbins 1976, chap. 2; Opie 1987. 12. Irland 1986. 13. Hough 1882, 10; Johnson 1976, chap. 8; Ranney, Bruner and Levenson, 1981; McDonnell and Stiles 1983. 14. Forman and Godron 1981. 15. Devèze 1966. 16. Darby 1976, 25, 21–26. 17. Bahre and Bradbury 1978; Humphrey 1987. 18. Hersberger and Bürgi 2009. 19. Devèze 1966. 20. Naveh and Kutiel 1990. 21. Nguyen et al. 2016. 22. Daskins, Stalmers and Pringle 2016. 23. Taylor 1953; United States Federal Census 1850–1960; Trimble 1974. 24. Forman et al. 2003. 25. Rosania et al. 2008. 26. Ullman 1965; Reynolds 1998; Bunce, Pérez-Soba and Smith 2009. 27. Erlande-Brandenburg 1995. 28. Ullman 1965; Grigg 1982, 143. 29. Williams 1989, 97; Schabel 1990. 30. Grigg 1982, 14; Gates, Clarke and Harris 1983. 31. Fearnside 1990; Cronon 1991; Skole and Tucker 1993. 32. Hough 1878, 112–116, 120, 122, 206–207, Olson 1971. 33. Hough 1882, 419; Forman and Boerner 1981. 34. Mack 1981; Billings 1990; Roberts and Stuckey 1992. 35. Mumford 1956, 382–398; Glacken 1967, 117; Deevey et al. 1979; Kagan, Ozment and Turner 1979, 17; Chase et al. 2011; Munoz et al. 2014. 36. Hodell et al. 2001; deMenocal 2001; Kidder 2006. 37. Mumford 1956; Glacken 1967, 126; Simmons 1989, 271. 38. Rapaport 1993; Thompson et al. 2004; Arce-Nazario 2007; Grimm et al. 2008; Stewart et al. 2009; Bertoncini et al. 2012; Grove et al. 2015; Fischer et al. 2016. 39. McDonnell and Pickett 1990; McDonnell, Pickett and Pouyat 1993; McKinney 2008. 40. Mumford 1956, 387. 41. Solomon and Kroener 1971; Bowlus 1980. 42. Highlands Study Team 1992. 43. Bishop 1992. 44. Engstrom and Wright 1984; Binford et al. 1987.

Notes to Pages 170–186  245 45. Davis and Stokes 1986; Smol et al. 1986; Steinberg and Wright 1994; Cummings et al. 2012. 46. Stanley and Warne 1993. 47. Sanderson 2009.

Chapter 10. Diversity and Species Extinctions   1. Glacken 1967.   2. Culotta 1991; Weakley, Ludwig and Townsend 2012; Townsend 2016; Vellend 2017.   3. Marsh 1864, 76–77; Bucher 1992; Ellsworth and McComb 2003; Hirschfeld and Heyd 2005; Martelli 2015, 139–141.   4. Marsh 1864, 84.   5. Elton 1958; Glacken 1967, 84; Fukarek 1979, 77; Birks, Felde and Seddon 2016.   6. Tilman et al. 1994; Jackson and Sax 2010.   7. Grime 1997; Naeem et al. 1999; Kaiser 2000.   8. Eldridge 1991, 1–12, 134; Jablonski 1991; Erwin 1993, 16; Stanley and Xang 1994.   9. Glacken 1967, 678; Sagoff 1993. 10. Barnosky et al. 2004; Miller et al. 2005; Haile et al. 2009; Vignieri 2014. 11. Noss et al. 2015; Cooper et al. 2015; Halligan et al. 2016. 12. Peet 1975; Schluter and Ricklefs 1993, 1–10; Raven 1994. 13. Webb 1987. 14. Willis 1922; Hutchinson 1959; Luh and Pimm 1993; Hubbell 2006; Damschen, Harrison and Grace 2010; Connolly et al. 2014. 15. Brooks 1985; Schluter and Ricklefs 1993; Latham and Ricklefs 1993. 16. Good 1964, 157; Woodward 1992; Whittaker, Willis and Field 2001; Mittelbach et al. 2007. 17. Pausas and Austin 2001. 18. Hillebrand et al. 2007. 19. Huston 1993; Tilman and Pacala 1993. 20. Beadle 1966. 21. Peet, Glenn-Lewin and Wolf 1983; Whittaker, Willis and Field 2001. 22. Whittaker and Niering 1965. 23. Beadle 1966, 997. 24. Marsh 1864, 77; Soulé 1991; Snyder 1993; Finkelstein et al. 2012. 25. Pielou 1991, 263; Erwin 1991. 26. Severinghouse and Brown 1956. 27. Marsh 1864, 78–79; Russell et al. 1993; Hung et al. 2014; Murray et al. 2017. 28. Jacobs and Bartgis 1987; https://saveplants.org/national-collection/plant-search /plant-profile/?CPCNum=4331. 29. Davis et al. 1986; Carlton 2003. 30. “The Green League Table” 1993. 31. Prowse et al. 2014. 32. Jackson et al. 2001; Gurevitch and Padilla 2004; Didham et al. 2005; Dunn 2005; VanDyck et al. 2009; Lotze and McClenachan 2013; Hewson et al. 2014.

246  Notes to Pages 188–200 33. Burgess 1988; Pimm and Gittleman 1992; Nepstad et al. 1996. 34. Tilman and Downing 1994. 35. Stohlgren, Barnett and Kartesz 2003; Thomas and Palmer 2015. 36. http://www.unesco.org/new/en/natural-sciences/environment/ecological-sciences /biosphere-reserves/europe-north-america/. 37. Fairbrothers 1979. 38. Pinelands Commission 1980. 39. Applegate, Little and Marucci 1979. 40. Hough 1882; Wacker 1979; Forman and Boerner 1981. 41. Fairbrothers 1979. 42. Peterken 1996; Rackham 1980. 43. Hughes and Huntley 1988; Hauge 1988; Kull and Zobel 1991; Baur et al. 2006. 44. Cousins 2001; Poschlod and Wallis-DeVries 2002; Cremene 2005; Poschlod et al. 2008; Henry, Talon and Dutoit 2010; Wolff, Tatin and Dutoit 2013. 45. Ashwal and Tucker 1999; United States Geological Survey 2004; HawiiHistory. org; Vorontsova et al. 2016. 46. McArthur and Wilson 1963; Darwin 1966, 388–410; Carlquist 1974. 47. Olson and James 1982; Richard and Dewar 1991. 48. Dewar et al. 2013. 49. Burney 1987; Burney et al. 1987; Klein 2004; Burney et al. 2004; Bond et al. 2008. 50. Richard and Dewar 1991; Bond and Silander 2007. 51. Kirch 2007. 52. Athens et al. 2002. 53. Vitousek 1987. 54. Xing and Ree 2017.

Chapter 11. Biospheric Sustainability in a Changing World   1. Hiers et al. 2016; http://www.un.org/sustainabledevelopment/sustainable-develop ment-goals/.   2. Julius et al. 2013; Lewis and Maslin 2015; Gross et al. 2016.   3. Coxe 1885, 3:210.   4. First translation quoted in Herlihy 1980, 103, second in Glacken 1967, 296.   5. Goodland 1995; http://www.un.org/sustainabledevelopment/sustainable-develop ment-goals/.   6. Costanza et al. 2007.   7. Smith 1980.   8. Caird 1851; Marsh 1864; Hough 1878, 1882; Haber 1958, 61–62; Thompson 1968; Lowdermilk 1975; Smith 1989; Wiedner et al. 2015.   9. Rackham 1980. 10. Poole 2008. 11. Pearce 1989; Lubchenko 1991. 12. Shearman 1990; Oreskes, Shrader-Freshette and Belitz 1994. 13. Hunter, Jacobson and Webb 1988.

Notes to Pages 201–216  247 14. Lorimer 1989, 2, 3; Parsons 1994. 15. Braun 1950; Vankat 1979; Millers, Shriner and Rizzo 1989; Whitney 1994, 78– 79; Healy and McShea 2002; McWilliams et al. 2002; Delcourt and Delcourt 2004; Oswalt et al. 2014. 16. Overpeck, Webb and Prentice 1985; Russell and Davis 2001; Foster and Aber 2004, chap. 5. 17. Bellemare, Motzkin and Foster 2002; McEwan, Dyer and Pederson 2011; Southgate and Thompson 2014. 18. Schwartz 1907. 19. Lutz 1928; Buell et al. 1966; Russell and Schuyler 1988; Mikan, Orwig and Abrams 1994. 20. Millers, Shriner and Rizzo 1989. 21. Barnes 1989. 22. Bard 1952; Forman and Elfstrom 1975; Parker, Leopold and Eichenberger 1985. 23. Fralish et al. 1991; Mikan, Orwig and Abrams 1994. 24. White and White 1996; Schuster 2011. 25. Whitcomb et al. 1981. 26. Korstian and Stickel 1927; Good 1965; Shugart and West 1977; Southgate 2006; Wang and Hu 2015. 27. Baiser et al. 2008. 28. Braun 1950, 457; Good 1965, 85; Willis 1973, 972. 29. Robinson, Burney and Burney 2005; Gill et al. 2009. 30. Watts 1979; Gaudreau and Webb 1985; Parshall and Foster 2002; Power et al. 2008. 31. Davis 1965; Schmidt 1973; Russell 1979. 32. Wilson and Bonaparte 1828, 2:253–261; Gifford 1900; Curtis 1956. 33. Gifford 1900; Russell et al. 1993. 34. Rothrock 1894; Vermeule 1900, 7–8, 40–41; Shreve et al. 1910. 35. Smith 1900; Westveld 1949, 84. 36. Lorimer 1989; Dey 2014. 37. Abrams and Scott 1989; Fralish et al. 1991, 307. 38. Maenza-Gmelch 1997; Cronin et al. 2003; Mann et al. 2009. 39. Mayewski et al. 2004; Pederson et al. 2005. 40. Pederson et al. 2015. 41. Shaw 1974; Johnson and Atkisson 1985; Johnson and Webb 1989. 42. Pickett, Cadenasso and Bartha 2001; Hart and Buchanan 2006. 43. Beauséjour et al. 2015. 44. Agnoletti 2006. 45. deMenocal 2001; Mayewski et al. 2004. 46. Millers, Shriner and Rizzo 1989; Botkin and Nisbet 1992; White and White 1996. 47. Streeter 1974, 248; Foster and Motzkin 1998; Healy and McShea 2002; Dey 2014. 48. Soberón and Peterson 2005; Peterson, Papes and Eaton 2007; Phillips 2017.

248  Notes to Pages 216–229 49. Shugart and West 1980; Oreskes, Shrader-Freshette and Belitz 1994; Phillips et al. 2017.

Conclusion   1. Santmire 1973.   2. Seddon et al. 2014.   3. Scheller et al. 2008; Östlund et al. 2015.   4. Delcourt and Delcourt 1991.   5. Caulfield 1981; Farrell and Doyle 2003; Caulfield et al. 2011.   6. O’Connell and Molloy 2001.   7. Pennington 1970; Green 1981; Gimingham and De Smidt 1983; Behre 1988; Goudie 1990, 56–59.   8. Bürgi, Östlund and Mladenoff 2017.   9. Cousins 2001; Henne et al. 2013; Szabó et al. 2016. 10. Plue et al. 2008; Von Holle, Wei and Nickerson 2010; Josefsson et al. 2010; Chase and Myers 2011. 11. Rotherham 2007. 12. Foster et al. 2002b. 13. Parshall and Foster 2002; Bond and Parr 2009; Noss et al. 2015. 14. Jackson and Sax 2010. 15. M. B. Davis 1984; T. Webb 1986; Cornell and Lawton 1992; Seddon et al. 2014; Gill et al. 2015. 16. Chesson and Case 1984; Behling, Pillar and Bauermann 2005. 17. Russell and Davis 2001; Danz et al. 2011. 18. Rhemtulla and Mladenoff 2007; Freschet et al. 2014. 19. White 1967; Glacken 1967; Williams 1974; Piasecki 1985; Bratton 1989; Wynn 1992; Russell 1995. 20. Palmer 1992. 21. Hunziker 1993; Watkins and Kirby 1998. 22. Henberg 1994. 23. Nelson et al. 2009. 24. Geological Survey of New Jersey 1900a, 10; Russell 1988a. 25. Bürgi et al. 2015. 26. Tallis et al. 2008. 27. Tuan 1968. 28. Overpeck, Bartlein and Webb 1991; Delcourt and Delcourt 1991; Webb 1992. 29. Moreno-Mateos et al. 2012; Gill et al. 2015; McClenachan et al. 2017; Padgett et al. 2017. 30. Henne et al. 2013; Beller, Askevold and Grossinger 2014; Szabó et al. 2016. 31. Williams 1989, 393–424; Bignal and McCracken 1996; Swetnam, Allen and Betancourt 1999; Fairhead and Leach 2014. 32. McNeill 2000. 33. Crutzen and Stoermer 2000.

Notes to Pages 229–231  249 34. The journals I am aware of are Anthropocene, Anthropocene Review, and Elementa: Science of the Anthropocene. 35. Ruddiman et al. 2015; Waters et al. 2016. 36. Bauer and Ellis 2018. 37. Sauer 1941; Glacken 1967; Bürgi, Hersberger and Schneeberger 2004. 38. Santmire 1973; Williams 1974, 14–15; Nash 1982, chap. 1; Milbraith 1985. 39. Barnosky et al. 2017. 40. McNutt 2013.

This page intentionally left blank

References

Abrams, M. D. and M. L. Scott. 1989. Disturbance-mediated accelerated succession in two Michigan forest types. Forest Science 35:42–49. Acocks, J. P. H. 1953. Veld types of South Africa. Memoirs of the Botanical Survey of South Africa 28:1–192. Agnoletti, M. 2006. Conservation of Cultural Landscapes. Centre for Agriculture and Bioscience International (CABI), Wallingford, UK. Agnoletti, M. and Anderson, S., eds. 2000. Methods and Approaches in Forest History. Vol. 3. Centre for Agriculture and Bioscience International (CABI), Wallingford, UK. Ahlgren, C. E. and I. F. Ahlgren. 1983. The human impact on northern forest eco­systems. Pp. 33–51 in Flader 1983. Aikens, C. M. 1983. Environmental archaeology in the western United States. Pp. 239– 251 in Late-Quaternary Environments of the United States, vol. 2, The Holocene, H. E. Wright Jr., ed., University of Minnesota Press, Minneapolis. Ainsworth, A. and J. B. Kaufman. 2010. Interactions of fire and non-native species across an elevation/plant community gradient at Hawaii Volcanoes National Park. Biotropica 42:647–55. Allan, W. 1965. The African Husbandsman. Oliver and Boyd, Edinburgh. (Quoted in Bigalke, R. C. 1978. Present-day mammals of Africa. Pp. 1–16 in Evolution of African Mammals, J. J. Maglio and H. B. S. Cooke, eds., Harvard University Press, Cambridge.) Allstadt, A. J. et al. 2015. Temporal variation in the synchrony of weather and its consequences for spatiotemporal population dynamics. Ecology 96:2935–2946.

251

252  References Anderson, D. E., A. S. Goudie and A. G. Parker. 2013. Global Environments through the Quaternary: Exploring Environmental Change. 2nd ed. Oxford University Press, Oxford. Anderson, K. 2005. Tending the Wild: Native American Knowledge and the Management of California’s Natural Resources. University of California Press, Berkeley. Anderson, M. K. 1990. California Indian horticulture. Fremontia 18(2):7–14. Anderson, R. and O. L. Loucks. 1979. White-tail deer (Odocoileus virginianus) influence on structure and composition of Tsuga canadensis forests. Journal of Applied Ecology 16:855–861. Anderson, W. L. 1960. Making land produce useful wildlife. Farmers’ Bulletin #2035, U.S. Department of Agriculture. Anonymous. 1926. On the processes of tanning, leather-dressing, and dying, etc. from Eakin’s Dictionary of Chymistry. Journal of the Franklin Institute 1:117–120, 143–146. ———. 1961. Hammond’s Advanced Reference Atlas. C. S. Hammond, Maplewood, NJ. ———. 1969. Invite Birds to Your Home. U.S. Department of Agriculture, PA-940, Soil Conservation Service. ———. 1976. New Jersey’s White-Tailed Deer: A Report on New Jersey’s Deer Management Program for Fiscal Year 1975–1976. New Jersey Division of Fish, Game and Shellfisheries, Trenton. ———. 1992. People vs. the ecosystem. Science 255:155. Applegate, J. E., S. Little and P. E. Marucci. 1979. Plant and animal products of the Pine Barrens. Pp. 25–36 in Pine Barrens: Ecosystem and Landscape, R. T. T. Forman, ed., Academic Press, New York. Aranbarri, J. et al. 2014. Rapid climatic changes and resilient vegetation during the Lateglacial and Holocene in a continental region of south-western Europe. Global and Planetary Change 114:50–65. Arce-Nazario, J. A. 2007. Human landscapes have complex trajectories: Reconstructing Peruvian Amazon landscape history from 1948 to 2005. Landscape Ecology 22:89– 101. Ashwal, L. D. and R. D. Tucker. 1999. Geology of Madagascar: A brief outline. Gon­ dwana Research 2:335–339. Athens, J. S. et al. 2002. Avifaunal extinctions, vegetation change, and Polynesian impacts in prehistoric Hawai’i. Archaeology in Oceania 37:57–78. Aubréville, A. 1971. Regeneration patterns in the closed forests of Ivory Coast, S.  R.  Eyre, trans. Pp. 40–55 in World Vegetation Types, S. R. Eyre, ed., Palgrave Macmillan, London. Austad, I. 1988. Tree-pollarding in western Norway. Pp. 11–30 in Birks et al. 1988. Austin, K. A. 1992. Gray dogwood (Cornus racemosa Lam.) as a refuge from herbivory in old fields of Saratoga National Historical Park, New York. PhD diss., State University of New York, College of Environmental Science and Forestry, Syracuse. Avery, T. E. 1977. Interpretation of Aerial Photographs, 3rd ed. Burgess, Minneapolis. Bahre, C. J. and D. E. Bradbury. 1978. Vegetation change along the Arizona-Sonora boundary. Annals of the Association of American Geographers 68:145–165. Bailes, K. E., ed. 1985a. Critical Issues in Comparative Perspective. University Press of America, Lanham, MD.

References  253 ———. 1985b. Critical issues in environmental history. Pp. 1–21 in Bailes 1985a. Baiser, B. et al. 2008. A perfect storm: Two ecosystem engineers interact to degrade deciduous forests of New Jersey. Biological Invasions: 10:785–795. Balée, W. L. 2002. Advances in Historical Ecology. Columbia University Press, New York. Bamforth, D.B. and B. Grund. 2012. Radiocarbon calibration curves, summed probability distributions, and early Paleoindian population trends in North America. Journal of Archaeological Science 39:1768–1774. Bard, G. E. 1952. Secondary succession on the piedmont of New Jersey. Ecological Monographs 22:195–215. Barnes, B. V. 1989. Old-growth forests of the northern lake states: A landscape ecosystem perspective. Natural Areas Journal 9:45–57. Barnosky, A. D. et al. 2004. Assessing the causes of late Pleistocene extinctions on the continents. Science 306:70–75. ———. 2017. Merging paleobiology with conservation biology to guide the future of terrestrial ecosystems. Science 355:594, science.eaah4787. Bartram, J. 1760. Letter to Peter Collinson, 24 June 1760. Pp. 85–86 in John and William Bartram’s America, H. G. Cruickshank, ed., Devin-Adair, Greenwich, CT, 1990. Baskin, J. M., C. C. Baskin and E. W. Chester. 1994. The Big Barrens region of Kentucky and Tennessee: Further Observations and Considerations. Castanea 59:226–254. Bates, M. 1956. Man as an agent in the spread of organisms. Pp. 788–804 in Thomas 1956. Bauer, A. M. and E. C. Ellis. 2018. The Anthropocene divide. Current Anthropology 59:209–227. Baur, B. et al. 2006. Effects of abandonment of subalpine hay meadows on plant and invertebrate diversity in Transylvania, Romania. Biological Conservation 132:261–273. Bayon, G. et al. 2012. Intensifying weathering and land use in Iron Age Central Africa. Science 335:1219–1222. Beach, T. et al. 2002. Upland agriculture in the Maya lowlands: Ancient Maya soil conservation in northwestern Belize. Geographical Review 92:372–397. Beadle, N. C. W. 1966. Soil phosphate and its role in molding segments of the Australian flora and vegetation with special reference to xeromorphy and sclerophylly. Ecology 47:992–1007. Beatty, R. M. and A. H. Taylor. 2008. Fire history and the structure and dynamics of a mixed conifer forest landscape in the northern Sierra Nevada, Lake Tahoe Basin, California, USA. Forest Ecology and Management 255:707–719. Beauséjour, R. et al. 2015. Historical anthropogenic disturbances influence patterns of non-native earthworm and plant invasions in a temperate primary forest. Biological Invasions 17:1267–1281. Behling, H., V. D. Pillar and S. G. Bauermann. 2005. Late Quaternary grassland (Campos), gallery forest, fire and climate dynamics, studied by pollen, charcoal and multi­­ variate analysis of the São Francisco de Assis core in western Rio Grande do Sul (southern Brazil). Review of Palaeobotany and Palynology 133:235–248. Behre, K-E. 1988. The rôle of man in European vegetation history. Pp. 633–667 in Huntley and Webb 1988. Belfer-Cohen, A. and O. Bar-Josef. 2002. Early sedentism in the Near East: A bumpy

254  References ride to village life. Pp. 19–37 in Life in Neolithic Farming Communities: Social Organization, Identity, and Differentiation, I. Kuijt, ed., Kluwer Academic, New York. Bellemare, J., G. Motzkin and D. R. Foster. 2002. Legacies of the agricultural past in the forested present: An assessment of historical land-use effects on rich mesic forests. Journal of Biogeography 29:1401–1420. Beller, E., R. Askevold and R. Grossinger. 2014. Futures past: Exploring California landscapes with the San Francisco Estuary Institute. Boom: A Journal of California 4(3):4–27. Benjamin, J. R. 2007. A Student’s Guide to History, 10th ed. Bedford/St. Martin’s, Boston. Bennett, K. D. and C. E. Buck. 2016. Interpretation of lake sediment accumulation rates. The Holocene 26:1092–1102. Berglund, B. E., ed. 1986. Handbook of Holocene Palaeoecology and Palaeohydrology. John Wiley and Sons, Chichester, UK. Bernabo, C. J. and T. Webb III. 1977. Changing patterns in the Holocene pollen record of northeastern North America: A mapped summary. Quaternary Research 8:64–96. Bertin, R. I. 2013. Changes in the native flora of Worcester County, Massachusetts. Journal of the Torrey Botanical Society 140:414–452. Bertoncini, A. P. et al. 2012. Local gardening practices shape urban lawn floristic communities. Landscape and Urban Planning 105:53–61. Betancourt, J. L., J. S. Dean, and H. M. Hull. 1986. Prehistoric long-distance transport of construction beams, Chaco Canyon, New Mexico. American Antiquity 51:370–375. Bignal, W. M. and D. I. McCracken. 1996. Low-intensity farming systems in the conservation of the countryside. Journal of Applied Ecology 33:413–424. Billings, W. D. 1990. Bromus tectorum, a biotic cause of ecosystem impoverishment in the Great Basin. Pp. 301–322 in Woodwell 1990. Bilsky, L. J., ed. 1980. Historical Ecology: Essays on Environment and Social Change. Kennikat, Port Washington, NJ. Binford, M. W. et al. 1987. Ecosystems, paleoecology and human disturbance in subtropical and tropical America. Quaternary Science Reviews 6:115–128. Birks, H. H. 1973. Modern macrofossil assemblages in lake sediments in Minnesota. Pp. 173–189 in Quaternary Plant Ecology, H. J. B. Birks and R. G. West, eds., Blackwell Scientific Publications, Oxford. Birks, H. H. et al. 1976. Recent paleolimnology of three lakes in northwestern Minnesota. Quaternary Research 6:249–272. ——— eds. 1988. The Cultural Landscape—Past, Present and Future. Cambridge University Press, Cambridge. Birks, H. J. B. 1986. Late-Quaternary biotic changes in terrestrial and lacustrine environments, with particular reference to north-west Europe. Pp. 3–65 in Berglund 1986. Birks, H. J. B., V. A. Felde and A. W. R. Seddon. 2016. Biodiversity trends within the Holocene. Holocene 26:994–1001. Birks, H. J. B. and A. D. Gordon. 1985. Numerical Analysis in Quaternary Pollen Analysis. Academic Press, London. Bishop, G. 1992. Newark mayor suggests urban-Highlands link. New York Times, 19 January 1992, sec. 1:22.

References  255 Black, J. D. 1950. The Rural Economy of New England. Harvard University Press, Cambridge. Bland, L. M. et al. 2018. Developing a standardized definition of ecosystem collapse for risk assessment. Frontiers in Ecology and the Environment 16:29–36. Blane, W. 1918. A tour in southern Illinois in 1822. Pp. 74–76 in Pictures of Illinois One Hundred Years Ago, M. M. Quaife, ed., R. R. Donnelley, Chicago. Bloch, M. 1953. The Historian’s Craft. Vintage Books, New York. ———. 1967. Land and Work in Medieval Europe, J. E. Anderson, trans. University of California Press, Berkeley. Bobiec, A. 2002. Białowieza primeval forest: The largest area of natural deciduous lowland forest in Europe. International Journal of Wilderness 8:33–37. Bond, A. L. K., A. Hobson and B. A. Branfireun. 2015. Rapidly increasing methyl mercury in endangered ivory gull (Pagophila eburnea) feathers over a 130 year record. Proceedings of the Royal Society B 282:20150032. doi:10.1098/rspb.2015.0032. Bond, A. L. K. and C. Parr. 2009. Beyond the forest edge: Ecology, diversity and conservation of the grassy biomes. Biological Conservation 143:2395–2404. Bond, W. J., G. F. Midgley and F. I. Woodward. 2003. What controls South African ­vegetation—climate or fire? South African Journal of Botany 69:79–91. Bond, W. J. and J. Silander. 2007. Madagascar’s grasslands: Anthropogenic or ancient? South African Journal of Botany 73:281. Bond, W. J., F. I. Woodward and G. F. Midgley. 2005. The global distribution of ecosystems in a world without fire. New Phytologist 165:525–537. Bond, W. J. et al. 2008. The antiquity of Madagascar’s grasslands and the rise of C4 grassy biomes. Journal of Biogeography 35:1743–1758. Bonnell, J. and M. Fortin. 2014. Using HGIS. Pp. 181–195 in Historical GIS Research in Canada, J. Bonnell and M. Fortin, eds., University of Calgary Press, Calgary. Bork, E. W., A. J. Hudson and R. W. Bailey. 1997. Populus forest characterization in Elk Island National Park relative to herbivory, prescribed fire, and topography. Canadian Journal of Botany 75:1518–1526. Bostoen, K. et al. 2015. Middle to late Holocene paleoclimatic change and the early Bantu expansion in the rain forests of Western Central Africa. Current Anthropology 56:367–368. Botkin, D. and R. L. Nisbet. 1992. Projecting the effects of climate change on biological diversity in forests. Pp. 277–293 in Global Warming and Biological Diversity, R. L. Peters and T. E. Lovejoy, eds., Yale University Press, New Haven. Bourdo, E. A. 1983. The forest the settlers saw. P. 3016 in Flader 1983. Bousquet, W. S. and G. P. Fleming. 2017. Floristics of the Abrams Creek wetlands, a calcareous fen complex in Winchester city and Frederick County, Virginia. Castanea 82:132–55. Bowlus, C. R. 1980. Ecological crisis in fourteenth century Europe. Pp. 86–99 in Bilsky 1980. Bowman, D. M. J. S. et al. 2009. Fire in the earth system. Science 324:481–484. ———. 2011. The human dimension of fire regimes on earth. Journal of Bio­geography 38:2223–2236.

256  References Boyden, S. 1987. Western Civilization in Biological Perspective. Clarendon, Oxford. Bradshaw, R. 2016. Spatial scale in the pollen record. Pp. 49–60 in Modelling Ecological Change, D. R. Harris and K. D. Thomas, eds., Routledge, London. Bradshaw, R. et al. 2015. Forest continuity and conservation value in western Europe. The Holocene 25:194–202. Bratton, S. P. 1989. Oaks, wolves and love: Celtic monks and northern forests. Journal of Forest History 33:4–20. Braun, E. L. 1950. Deciduous Forests of Eastern North America. Blakiston, Philadelphia. Brewer, S. et al. 2017. Late-glacial and Holocene European pollen data. Journal of Maps 13:921–928. Bridges, E. M. 1970. World Soils. Cambridge University Press, Cambridge. Britton, N. L. 1881. A Preliminary Catalogue of the Flora of New Jersey. Geological Survey of New Jersey, New Brunswick, NJ. Brooks, D. R. 1985. Historical ecology: A new approach to studying the evolution of ecological associations. Annals Missouri Botanical Garden 72:660–680. Brose, P. et al. 2001. Bringing fire back: The changing regimes of the Appalachian mixed-oak forests. Journal of Forestry 99:30–35. Brown, A. A. 1949. Progress, but still a problem. Pp. 477–479 in Stefferud 1949. Brown, A. A. and K. P. Davis. 1973. Forest Fire Control and Its Use. 2nd. ed. McGrawHill, New York. Brown, J. R. 1960. The role of fire in altering the species composition of forests in Rhode Island. Ecology 41:310–316. Brugam, R. B. 1978. Pollen indicators of land-use change in southern Connecticut. Quaternary Research 9:349–362. Brush, G. S and F. W. Davis. 1984. Stratigraphic evidence of human disturbance in an estuary. Quaternary Research 22:91–108. Bucher, E. H. 1988. Impact of European colonization on the chaco savannas of Argentina. Bulletin of the Ecological Society of America 69:86. ———. 1992. The causes of extinction of the passenger pigeon. Current Ornithology 9:1–36. Buell, M. F., H. F. Buell, and J. A. Small. 1954. Fire in the history of Mettler’s Woods. Bulletin of the Torrey Botanical Club 81:253–255. Buell, M. F. et al. 1966. The upland forest continuum in northern New Jersey. Ecology 47:416–432. Buisson, E., T. Dutoit and A. Wolff. 2004. Bilan de trente années de recherches en écologie dans la steppe de Crau (Bouches-du-Rhône, sud-est de la France). Ecologia Mediterranea 30:7–24. Bunce, R. G. H., M. Pérez-Soba and M. Smith. 2009. Assessment of the extent of agroforestry systems in Europe and their role within transhumance systems. Pp. 321– 329 in Agroforestry in Europe: Current Status and Future Prospects, A. Rigueiro-­ Rodríguez et al., eds., Springer Science and Business Media, n.p. Bunting, M. J. and R. Middleton. 2009. Equifinality and uncertainty in the interpretation of pollen data: The Multiple Scenario Approach to reconstruction of past vegetation mosaics. The Holocene 19:799–803.

References  257 Bunting, M. J. et al. 2018. Maps from mud—Using the multiple scenario approach to reconstruct land cover dynamics from pollen records: A case study of two Neolithic landscapes. Frontiers in Ecology and Evolution 6:1–20. Buol, S., W. F. D. Hole and R. J. McCracken. 1980. Soil Genesis and Classification. 2nd ed. Iowa State University Press, Ames. Burgess, R. L. 1988. Community organization: Effects of landscape fragmentation. Canadian Journal of Botany 66:2687–2690. Burgess, R. L. and D. M. Sharpe, eds. 1981a. Forest Island Dynamics in Man-­Dominated Landscapes. Springer-Verlag, New York. ———. 1981b. Introduction. Pp. 1–5 in Burgess and Sharpe 1981a. Bürgi, M. et al. 2015. Linking ecosystem services with landscape history. Landscape Ecology 30:11–20. Bürgi, M., A. M. Hersperger and N. Schneeberger. 2004. Driving forces of landscape direction—Current and new directions. Landscape Ecology 19:857–868. Bürgi, M., L. Östlund and D. Mladenoff. 2017. Legacy effects of human land use: Ecosystems as time-lagged systems. Ecosystems 20:94–103. Burney, D. A. 1987. Late Quaternary stratigraphic charcoal records from Madagascar. Quaternary Research 28:274–286. Burney, D. A. et al. 1987. L’environnement au cours de L’Holocène et la disparition de la mégafaune à Madagascar: Quel rapport avec la conservation de la nature? Pp. 137– 143 in Priorités en Matière de Conservation des Espèces à Madagascar, R. A. Mit­termeier et al., eds., Antananarivo, Madagascar. ——— 2004. A chronology for late prehistoric Madagascar. Journal of Human Evolution 47:25–63. Bush, M. B. et al. 1992. A 14,300-yr paleoecological profile of a lowland tropical lake in Panama. Ecological Monographs 62:251–275. Bushness, H. 1864. Sermon on the power of an endless life. P. 1 in Man and Nature, G. P. Marsh, Belknap Press of Harvard University, Cambridge, 1965. Butler, G. R. and B. P. Malanson. 2005. The geomorphic influences of beaver dams and failures of beaver dams. Geomorphology 71:48–60. Butzer, K. W. 2005. Environmental history in the Mediterranean world: Cross-­ disciplinary investigation of cause-and-effect for degradation and soil erosion. Journal of Archaeological Science 32:1773–1800. Caird, J. 1851. English Agriculture in 1850–1. Longman, Brown, Green and Longmans, London. Callmander, M. W. et al. 2011. The endemic and non-endemic vascular flora of Madagascar updated. Plant Ecology and Evolution 144:121–225. Campbell, I. D. and J. H. McAndrews. 1993. Forest disequilibrium caused by rapid Little Ice Age cooling. Nature 366:336–338. Canuto, M. A. et al. 2018. Ancient lowland Maya complexity as revealed by airborne laser scanning of northern Guatemala. Science 361:1355. dx.doi.org/10.1126/science .aau0137. Carloni, K. R. 2005. The ecological legacy of Indian burning practices in southwestern Oregon. PhD diss., Oregon State University, Corvallis. Carlquist, S. J. 1974. Island Biology. Columbia University Press, New York.

258  References Carlton, J. T. 2003. Community assembly and historical biogeography in the North Atlantic Ocean: The potential role of human-mediated dispersal vectors. Hydro­biologia 503:1–8. Carver, J. 1796. Three Years Travels through the Interior Parts of North America. Key and Simpson, Philadelphia. Caughley, G. 1985. The Deer Wars: The Story of Deer in New Zealand. Heinemann, Auckland. Caulfield, S. 1981. Forest use and land use in Mayo around 3000 B.C. Irish Forestry 38:92–100. Caulfield, S. et al. 2011. Excavations on Céide Hill, Behy & Glenulra, North Co. Mayo, 1963–1994. UCD School of Archaeology and the Irish Strategic Archaeological Research Programme (INSTAR). Cavalli-Sforza, L. L., P. Menozzi and A. Piazza. 1993. Demic expansion and human evolution. Science 259:639–646. Chambers, F. M. and L. Elliott. 1989. Spread and expansion of Alnus Mill. in the British Isles: Timing, agencies and possible vectors. Journal of Biogeography 16:541–550. Chandler, C. et al. 1983. Fire in Forestry. Vol. 1, Forest Fire Behavior and Effects. John Wiley and Sons, New York. Chase, A. F. et al. 2011. Airborne LiDAR, archaeology, and the ancient Maya landscape at Caracol, Belize. Journal of Archaeological Science 38:387–398. Chase, J. M. and J. A. Myers. 2011. Disentangling the importance of ecological niches from stochastic processes across scales. Philosophical Transactions of the Royal Society of London B: Biological Science 366:2351–2363. Chazdon, R. L. 2014. Second Growth: The Promise of Tropical Forest Regeneration in an Age of Deforestation. University of Chicago Press, Chicago. Chen, F. H. et al. 2015. Agriculture facilitated permanent human occupation of the Tibetan Plateau after 3600 B.P. Science 347:248–250. Chen, H. et al. 2011. Detecting one-hundred-year environmental changes in western China using seven-year repeat photography. PLOS One 6(9):325008. Chesson, P. L. and T. J. Case. 1984. Overview: Nonequilibrium community theories: Chance, variability, history. Pp. 229–239 in Community Ecology, J. Diamond and T. J. Cas, eds., Harper and Row, New York. Chittendon, H. M. 1905. Letters and Travels of Father Pierre-Jean de Smet, S.J., 1801– 1873. 4 vols. Francis P. Harper, New York. Christensen, N. L. 1981. Fire regimes in southeastern ecosystems. Pp. 112–136 in Mooney et al. 1981. Christensen, N. L. 1989. Landscape history and ecological change. Forest & Conservation History 33:116–125. ———. 2014. An historical perspective on forest succession and its relevance to ecosystem restoration and conservation practice in North America. Forest Ecology and Management 330:312–322. Christensen, N. L. et al. 1989. Interpreting the Yellowstone fires of 1988. Bio­Science 39:678–685. Clark, A. C. 1956. The impact of exotic invasion on the remaining New World mid-­ latitude grasslands. Pp. 737–762 in Thomas 1956.

References  259 Clark, D. C. 1995. Edaphic and human effects on landscape-scale distributions of tropical rain-forest palms. Ecology 76:2581–2594. Clark, J. S., J. Merkt, and H. Muller. 1989. Post-glacial fire, vegetation, and human history on the northern alpine forelands, southwestern Germany. Journal of Ecology 77:897–925. Clark, J. S. and P. D. Royall. 1995. Transformation of a northern hardwood forest by aboriginal (Iroquois) fire: Charcoal evidence from Crawford Lake, Ontario, Canada. The Holocene 5:1–9. Clawson, M. and C. L. Stewart. 1965. Land Use Information: A Critical Survey of U. S. Statistics Including Possibilities for Greater Uniformity. Resources for the Future, Washington, DC. Clement, R. M. and S. P. Horn. 2001. Pre-Columbian land-use history in Costa Rica: A 3000-year record of forest clearance, agriculture and fires from Laguna Zoncho. The Holocene 11:419–426. Clements, F. E. 1916. Plant Succession: An Analysis of the Development of Vegetation. Carnegie Institute of Washington Publication #242. ———. 1936. Nature and structure of the climax. Ecology 24:252–284. Clout, H. D., ed. 1977. Themes in the Historical Geography of France. Academic Press, London. Clout, M. N. and J. C. Russell. 2008. The invasion ecology of mammals: A global perspective. Wildlife Research 35:180–184. Collinson, P. 1738. Letter to John Bartram, 6 April 1738. Pp. 73–74 in John and William Bartram’s America, H. G. Cruickshank, ed., Devin-Adair, Greenwich, CT, 1990. Connolly, S. R. et al. 2014. Commonness and rarity in the marine biosphere. Proceedings of the National Academy of Science 111:8524–8529. Contreras, D. A. and J. Meadows. 2014. Summed radiocarbon calibrations as a population proxy: A critical evaluation using a realistic simulation approach. Journal of Archaeological Science 52:591–608. Cook, E. R. et al. 2010. Asian monsoon failure and megadrought during the last millennium. Science 328:486–489. ———. 2015. Old World megadroughts and pluvials during the Common Era. Science Advances 1:e1500561. Cooper, A. et al. 2015. Abrupt warming events drove Late Pleistocene Holarctic megafaunal turnover. Science 349:602–606. Corbet, G. B. 1974. The importance of oak to mammals. Pp. 312–323 in Morris and Perring 1974. Cordain, L. et al. 2000. Plant-animal subsistence ratios and macronutrient energy estimations in worldwide hunter-gatherer diets. Journal of Clinical Nutrition 71:682– 692. Cordova, C. E. 2017. Late Pleistocene-Holocene vegetation and climate change in the Middle Kalahari, Lake Ngami, Botswana. Quaternary Science Reviews 171:199–215. Cornell, H. V. and J. H. Lawton. 1992. Species interactions, local and regional processes and limits to the richness of ecological communities: A theoretical perspective. Journal of Animal Ecology 61:1–12. Costanza, R. et al. 2007. Sustainability or collapse: What can we learn from integrating

260  References the history of humans and the rest of nature? AMBIO: A Journal of the Human Environment 36:522–527. Council Directive 92/43/EEC of 21 May 1992 on the conservation of natural habitats and of wild fauna and flora, EU Natura 2000. Cousins, S. A. O. 2001. Analysis of land-cover transitions based on 17th and 18th century cadastral maps and aerial photographs. Landscape Ecology 16:41–54. Coxe, A. C. 1885. Latin Christianity: Its Founder, Tertullian, Ante-Nicene Fathers. Hendrickson, Peabody, MA. Cramer, D. 2015. The Narrow Edge: A Tiny Bird, an Ancient Crab, and an Epic Journey. Yale University Press, New Haven. Crawford, A. J., K. R. Lips and E. Bermingham. 2010. Epidemic disease decimates amphibian abundance, species diversity, and evolutionary history in the highlands of central Panama. Proceedings of the National Academy of Science 107:13777–13782. Cremene, C. et al. 2005. Alterations of steppe-like grasslands in eastern Europe: A threat to regional biodiversity hotspots. Conservation Biology 19:1606–1618. Crisp, M. D. et al. 2011. Flammable biomes dominated by eucalypts originated at the Cretaceous-Palaeogene boundary. Nature Communications 2:1–8. Cronin, T. M. et al. 2003. Medieval Warm Period, Little Ice Age and 20th century temperature variability from Chesapeake Bay. Global and Planetary Change 36:17–29. Cronon, W. 1983. Changes in the Land: Indians, Colonists, and the Ecology of New England. Hill and Wang, New York. ———. 1991. Nature’s Metropolis. Chicago and the Great West. W. W. Norton, New York. Crosby, A. W. 1972. The Columbian Exchange: Biological and Cultural Consequences of 1492. Greenwood, Westport, CT. ———. 1986. Ecological Imperialism: The Biological Expansion of Europe, 900–1900. Cambridge University Press, Cambridge. ———. 1995. The past and present of environmental history. American Historical Review 100:1177–1189. Crumley, C. L., ed. 1994. Historical Ecology: Cultural Knowledge and Changing Landscapes. School of American Research Press, Santa Fe. Crumley, C. L., T. Lennartsson and A. Westin. 2018. Issues and Concepts in Historical Ecology: The Past and Future of Landscapes and Regions. Cambridge University Press, Cambridge. Crutzen, P. J. and E. F. Stoermer. 2000. The “Anthropocene.” Global Change Newsletter 41:17–18. Cuddihy, L. W. and C. P. Stone. 1990. Alteration of Native Hawaiian Vegetation. Cooperative National Park Resources Studies Unit, University of Hawaii, Honolulu. Culotta, E. 1991. Biological immigrants under fire. Science 234:1444–1447. Cummings, B. F. et al. 2012. Tracking Holocene Climatic Change with Aquatic Biota from Lake Sediments: Case Studies of Commonly used Numerical Techniques. Springer, Dordrecht. Curtis, J. H. et al. 1998. A multiproxy study of Holocene environmental change in the Maya Lowlands of Peten, Guatemala. Journal of Paleolimnology 19:139–159.

References  261 Curtis, J. T. 1956. The modification of mid-latitude grasslands and forests by man. Pp. 721–762 in Thomas 1956. Cushman, G. T. 2013. Guano and the Opening of the Pacific World: A Global Ecological History. Cambridge University Press, Cambridge. Dambrine, E. et al. 2007. Present forest biodiversity patterns in France related to former Roman agriculture. Ecology 88:1430–1439. Damschen, E. I., S. Harrison and J. B. Grace. 2010. Climate change effects on an endemic-rich edaphic flora: Resurveying Robert H. Whittaker’s Siskiyou sites (Oregon, USA). Ecology 91:3609–3619. Danz, N. P. et al. 2011. Vegetation controls vary across space and spatial scale in a historic grassland-forest biome boundary. Ecography 34:402–414. Darby, H. C. 1976. A New Historical Geography of England Before 1600. Cambridge University Press, Cambridge. Darlington, C. D. 1969. The Evolution of Man and Society. Simon and Schuster, New York. Darwin, C. 1966. On the Origin of Species: A Facsimile of the First Edition. Harvard University Press, Cambridge. Daskins, J. H., M. Stalmers and R. M. Pringle. 2016. Ecological legacies of civil war: 35-year increase in savanna tree cover following wholesale large-mammal declines. Journal of Ecology 104:79–89. Daube, D. 1973. The self-understood in legal history. Juridical Review, n.s., 18:126–134. Daura, J. et al. 2016. Palaeoenvironmental record of the Cal Maurici wetland sediment archive in Barcelona (NE Iberian Peninsula) between c. 6000 and 4000 cal. yr BP. The Holocene 26:1020–1039. Davis, A. M. 1984. Dating with pollen: Methodology, applications, limitations. Pp. 283–297 in W. C. Mahaney, ed., Quaternary Dating Methods, Elsevier, Amsterdam. Davis, M. B. 1965. Phytogeography and palynology of northeastern United States. Pp. 377–401 in The Quaternary of the United States, H. E. Wright and D. G. Frey, eds., Princeton University Press, Princeton, NJ. ———. 1983. Holocene vegetational history of the eastern United States. Pp. 166–181 in Wright 1983. ——— 1984. Climatic instability, time lags, and community disequilibrium. Pp. 269– 228 in Community Ecology, J. Diamond and T. J. Case, eds., Harper and Row, New York. Davis, M. B. and E. S. Deevey Jr. 1964. Pollen accumulation rates: Estimates from late-glacial sediment of Rogers Lake. Science 145:1293–1295. Davis, M. B. et al. 1986. Dispersal versus climate: Expansion of Fagus and Tsuga into the Upper Great Lakes region. Vegetatio 67:93–103. Davis, O. K. 1994. Aspects of Archaeological Palynology: Methodology and Applications. American Association of Stratigraphic Palynologists Contribution Series 29, AASP Foundation, Dallas. Davis, R. B. 1987. Paleolimnological diatom studies of acidification of lakes by acid rain: An application of Quaternary science. Quaternary Science Reviews 6:147–163. Davis, R. B. and S. A. Norton. 1978. Paleolimnologic studies of human impact on lakes

262  References in the United States, with emphasis on recent research in New England. Polskie Archiwum Hydrobiology 25:99–115. Davis, R. B. and P. M. Stokes. 1986. Overview of historical and paleoecological studies of acidic air pollution and its effects. Water, Air and Soil Pollution 30:311–318. Davis, R. B. and T. Webb III. 1975. The contemporary distribution of pollen in eastern North America: A comparison with the vegetation. Quaternary Research 5:395–434. Day, G. M. 1953. The Indian as an ecological factor in the northeastern forest. Ecology 34:329–346. Dean, P. B. and A. de Vos. 1965. The spread and present status of the European hare, Lepus europaeus hybridus (Demarest), in North America. Canadian Field Naturalist 79:38–48. Deevey, E. S. 1969. Coaxing history to conduct experiments. BioScience 19:40–43. ———. 1984. Stress, strain, and stability of lacustrine ecosystems. Pp. 203–229 in Haworth and Lund 1984. Deevey, E. S. et al. 1979. Mayan urbanism: Impact on a tropical karst development. Science 206:298–306. Delcourt, H. R. 1987. The impact of prehistoric agriculture and land occupation on natural vegetation. Trends in Ecology and Evolution 2:39–44. Delcourt, H. R. and P. A. Delcourt. 1991. Quaternary Ecology: A Paleoecological Perspective. Chapman and Hall, London. Delcourt, P. A. and H. R. Delcourt. 2004. Prehistoric Native Americans and Ecological Change: Human Ecosystems in Eastern North America since the Pleistocene. Cambridge University Press, Cambridge. deMenocal, P. B. 2001. Cultural responses to climate change during the late Holocene. Science 292:667–673. Denham, T., S. Haberle and C. Lentfer. 2004. New evidence and revised interpretations of early agriculture in Highland New Guinea. Antiquity 78:839–857. Densmore, F. 1974. How Indians Use Wild Plants for Food, Medicine and Crafts. Dover, New York. (Unabridged republication of “Uses of plants by the Chippewa Indians,” pp. 275–397 in 44th Annual Report of the Bureau of American Ethnology, Government Printing Office, Washington, D.C., 1928.) Devèze, M. 1966. Forêts françaises et forêts allemandes. Revue Historique 236:47–68. Dewar, R. E. et al. 2013. Stone tools and foraging in northern Madagascar challenge Holocene extinction models. Proceedings of the National Academy of Sciences 110:12583–12588. Dey, D. C. 2014. Sustaining oak forests in eastern North America: Regeneration and recruitment, the pillars of sustainability. Forest Science 60:926–942. Diamond, J. 1987. The worst mistake in the history of the human race. Discover, May, 64–66. Didham, R. K. et al. 2005. Are invasive species a major cause of extinctions? Trends in Ecology and Evolution 20:470–474. Dobbs, A. 1755. 24 August letter “to the Board.” P. 361 in Colonial Records of North Carolina, vol. 5, W. L. Saunders, ed., P. M. Hale, Raleigh, 1886. Dorney, J. R. 1981. The impact of native Americans on presettlement vegetation in southeastern Wisconsin. Wisconsin Academy of Science, Arts and Letters 69:26–36.

References  263 ———. 1983. Increase A. Lapham’s pioneer observations and maps of landforms and natural disturbances. Wisconsin Academy of Science, Arts and Letters 71:25–30. Drábková, L. Z. 2014. DNA Extraction from Herbarium Specimens. Springer, New York. Dragoo, D. W. 1976. Some aspects of eastern North American prehistory: A review, 1975. American Antiquity 41:3–27. Drake, J. A. et al., eds. 1989. Biological Invasions: A Global Perspective, SCOPE vol.37. John Wiley and Sons, New York. Drescher-Schneider, R. et al. 2007. Vegetation history, climate and human impact over the last 15,000 years at Lago dell’Accesa (Tuscany, Central Italy). Vegetation History and Archaeobotany 16:279–299. Driver, H. L. 1969. Indians of North America, 2nd ed. University of Chicago Press, Chicago. Duby, G. 1968. Rural Economy and Country Life in the Medieval West, Cynthia Postan, trans. University of South Carolina Press, Columbia. Dunham, S. 2009. Nuts about acorns: A pilot study on acorn use in Woodland period subsistence in the eastern Upper Peninsula of Michigan. Wisconsin Archeologist 9:113–130. Dunn, R. R. 2005. Modern insect extinctions, the neglected majority. Conservation Biology 19:1030–1036. Dunwiddie, P. W. 1989. Forest and heath: The shaping of the vegetation on Nantucket Island. Journal of Forest History 33:126–133. Dupouey, J. L. et al. 2002. Irreversible impact of past land use on forest soils and biodiversity. Ecology 83:2978–2984. Dutoit, T. et al. 2009. Sampling soil wood charcoals at a high spatial resolution: A new methodology to investigate the origin of grassland plant communities. Journal of Vegetation Science 20:349–358. Dwight, T. 1969. Travels in New England and New York. B. M. Solomon, ed. Harvard University Press, Cambridge. (Originally published in 1841, printed by the author, New Haven, CT.) Edmonds, M. 2005. The pleasures and pitfalls of written records. Pp. 73–99 in Egan and Howell 2005. Egan, D. and E. Howell. 2005. Handbook of Historical Ecology. Island Press, Washington, DC. Eldredge, Z. S. 1909. Note made 7 November 1769, in The March of Portola and the Discovery of the Bay of San Francisco. California Promotion Committee, San Francisco. Eldridge, N. 1991. The Miner’s Canary: Unraveling the Mysteries of Extinction. ­Prentice-Hall, New York. Elias, S. A. 1994. Quaternary Insects and Their Environments. Smithsonian Institution Press, Washington, D.C. Ellison, A. M. et al. 2005. Loss of foundation species: Consequences for the structure and dynamics of forested ecosystems. Frontiers in Ecology and the Environment 3:479–486. Ellsworth, J. W. and McComb, B. C. 2003. Potential effects of passenger pigeon flocks

264  References on the structure and composition of presettlement forests of eastern North America. Conservation Biology 17:1548–1558. Ellsworth, L. F. 1975. Craft to National Industry in the Nineteenth Century: A Case Study of the Transformation of the New York State Tanning Industry. Arno, New York. Elton, C. S. 1958. The Ecology of Invasions by Animals and Plants. Methuen, London. Emerson, G. B. and C. L. Flint. 1862. Manual of Agriculture for the School, the Farm and the Fireside. Brewer and Tileston, Boston. Engstrom, D. R. and H. E. Wright Jr. 1984. Chemical stratigraphy of lake sediments as a record of environmental change. Pp. 11–64 in Haworth and Lund 1984. Eriksson, O. 2013. Species pools in cultural landscapes—Niche construction, ecological opportunity and niche shifts. Ecography 36:403–413. Erlande-Brandenburg, A. 1995. Cathedrals and Castles: Building in the Middle Ages. R. Stonehewer, trans. Harry N. Abrams, New York. Erlandson, J. M. 2001. The archaeology of aquatic adaptations: Paradigms for a new millennium. Journal of Archaeological Research 9:287–350. Erlandson, J. M. and T. C. Rick. 2010. Archaeology meets marine ecology: The antiquity of maritime cultures and human impacts on marine fisheries and ecosystems. Annual Review of Marine Science 2:165–185. Erlandson, J. M. et al. 2011. 10,000 years of human predation and size changes in the owl limpet (Lottia gigantea) on San Miguel Island, California. Journal of Archaeological Research 38:1127–1134. Erwin, D. H. 1993. The Paleozoic Crisis: Life and Death in the Permian. Columbia University Press, New York. Erwin, T. L. 1991. An evolutionary basis for conservation strategies. Science 253:750– 752. Evans, D. 1992. A History of Nature Conservation in Britain, 2nd ed. Routledge, London. Evans, J. G. 1975. The Environment of Early Man in the British Isles. Paul Elek, London. Evans, P. E. et al. 2018. Quantification of drought during the collapse of the classic Maya civilization. Science 361:498–501. Everitt, B. S. et al. 2011. Cluster Analysis. 5th ed. Wiley and Sons, Chichester, UK. Eyre, S. R. and R. J. Jones. 1966. Introduction. Pp. 1–29 in Geography as Human Ecology: Methodology by Example, S. R. Eyre and R. J. Jones, eds., Edward Arnold, London. Faegri, K., P. E. Kaland and K. Krzywinski. 1989. Textbook of Pollen Analysis, 4th ed. John Wiley and Sons, Chichester, UK. Fairbrothers, D. E. 1979. Endangered, threatened, and rare vascular plants of the Pine Barrens and their biogeography. Pp. 395–405 in Pine Barrens: Ecosystem and Landscape, R. T. T. Forman, ed., Academic Press, New York. Fairhead, J. and M. Leach. 2014. False forest history, complicit social analysis: Rethinking some west African environmental narratives. Pp. 14–30 in The Social Lives of Forests, S. B. Hecht, K. D. Morrison and C. Padoch, eds., University of Chicago Press, Chicago.

References  265 Farrell, C. A. and G. J. Doyle. 2003. Rehabilitation of industrial cutaway Atlantic blanket bog in County Mayo, north-west Ireland. Wetlands Ecology and Management 11:21–35. Fearnside, P. M. 1990. Deforestation in Brazilian Amazonia. Pp. 211–238 in Woodwell 1990. Feeley, K. J. 2015. Moving forward with species distributions. American Journal of Botany 102:73–75. Fidino, M. and S. B. Magle. 2017. Trends in long-term urban bird research. Pp. 161–184 in Ecology and Conservation of Birds in Urban Environments, E. Murgui, M. H ­ edblom, eds., Springer International, Heidelberg. Finkelstein, M. E. et al. 2012. Lead poisoning and the deceptive recovery of critically endangered California condors. Proceedings of the National Academy of Science 109:11449–11454. Finsinger, W. et al. 2017. Emergence patterns of novelty in European vegetation assemblages over the past 15 000 years. Ecology Letters 20:336–346. Firbas, F. 1937. Der pollenanalytische Nachweis des Getreidebaus. Zeitschrift Für Botanik 31:447–478. Fischer, L. K. et al. 2016. Drivers of biodiversity patterns in parks of a growing South American megacity. Urban Ecosystems 19:1231–1249. Fish, S. K. 2000. Hohokam impacts on Sonoran desert environment. Pp. 251–280 in Imperfect Balance: Landscape Transformations in the Precolumbian Americas, D. L. Lentz, ed., Columbia University Press, New York. Flader, S. L., ed. 1983. The Great Lakes Forest: An Environmental and Social History. University of Minnesota Press and Forest History Society, Minneapolis. Flemley, J. R. 1979. The Equatorial Rain Forest: A Geological History. Butterworth, London. Flinn, K. M. and M. Vellend. 2005. Recovery of forest plant communities in post-­ agricultural landscapes. Frontiers in Ecology and the Environment 3:243–250. Forman, R. T. T. and R. E. Boerner. 1981. Fire frequency and the Pine Barrens of New Jersey. Bulletin of the Torrey Botanical Club 108:34–50. Forman, R. T. T. and B. A. Elfstrom 1975. Forest structure comparison of Hutcheson Memorial Forest and eight old woods on the New Jersey Piedmont. William L. Hutcheson Memorial Forest Bulletin 3:44–51. Forman, R. T. T. and M. Godron. 1981. Patches and structural components for a landscape ecology. BioScience 31:733–740. ———. 1986. Landscape Ecology. John Wiley and Sons, New York. Forman, R. T. T. and E. W. B. Russell. 1983. Evaluation of historical data in ecology. Bulletin of the Ecological Society of America 64:5–7. Forman, R. T. T. et al. 2003. Road Ecology. Science and Solutions. Island Press, Washington, DC. Fortin, D. et al. 2005. Wolves influence elk movements: Behavior shapes a trophic cascade in Yellowstone National Park. Ecology 86:1320–1330. Foster, D. R. and J. D. Aber. 2004. Forests in Time: The Environmental Consequences of 1,000 Years of Change in New England. Yale University Press, New Haven. Foster, D. R. and G. Motzkin. 1998. Grasslands, heathlands and shrublands in coastal

266  References New England: Historical interpretations and approaches to conservation. Journal of Biogeography 29:1569–1590. Foster, D. R. et al. 2002a. Oak, chestnut and fire: Climatic and cultural controls of longterm forest dynamics in New England, USA. Journal of Biogeography 29:1359–1379. ———. 2002b. Simulating a catastrophic hurricane. Pp. 235–258 in Foster and Aber 2004. ———. 2003. The importance of land-use legacies to ecology and conservation. BioScience 53:77–88. Fralish, J. L. et al. 1991. Comparison of presettlement, second-growth and old-growth forest on six site types in the Illinois Shawnee Hills. American Midland Naturalist 125:294–309. Franklin, J. F. and R. T. T. Forman. 1987. Creating landscape patterns by logging: Ecological consequences and patterns. Landscape Ecology 1:5–18. Franklin, S. B., P. A. Robertson and J. S. Fralish. 2003. Prescribed burning effects on upland Quercus forest structure and function. Forest Ecology and Management 184:315–335. Fraterrigo, J.M., T. C. Balser and M. G. Turner. 2006. Microbial community variation and its relationship with nitrogen mineralization in historically altered forests. Ecology 7:570–579. Freschet, G. T. et al. 2014. Aboveground and belowground legacies of native Sami land use on boreal forest in northern Sweden 100 years after abandonment. Ecology 95:963–977. Fritts, H. C. 1991. Reconstructing Large-Scale Climatic Patterns from Tree-Ring Data: A Diagnostic Analysis. University of Arizona Press, Tucson. Frye, R. J. II. 1978. Structural dynamics of early old-field succession on the New Jersey Piedmont: A comparative approach. PhD diss., Rutgers University. Fukarek, F. 1979. Pflanzenwelt der Erde. Urania-Verlag, Leipzig. Fuller, D. Q. 2010. An emerging paradigm shift in the origins of agriculture. General Anthropology 17:1, 8–12. Fyfe, R. 2006. GIS and the application of a model of pollen deposition and dispersal: A new approach to testing landscape hypotheses using the POLLANDCAL models. Journal of Archaeological Science 33:483–493. Gaillard, M. J. et al. 2009. European cultural landscapes: Insights into origins and development. Pp. 35–46 in Krzywinski, Connell, and Küster 2009. ­ ondon. Garrard, I. and D. Streeter. 1983. The Wildflowers of the British Isles. Macmillan, L Garrod, D. A. E. 1932. A new Mesolithic industry: The Natufian of Palestine. Journal of the Royal Anthropological Institute of Great Britain and Ireland 62:257–269. Gates, D. M., C. H. D. Clarke and J. T. Harris. 1983. Wildlife in a changing environment. Pp. 52–81 in Flader 1983. Gaudreau, D. C. and T. Webb III. 1985. Late-quaternary pollen stratigraphy and isochrone maps for the northeastern United States. Pp. 247–280 in Pollen Records of Late-Quaternary North American Sediments, V. M. Bryant and R. G. Holloway, eds., AASP Foundation, Dallas. Gavin, D. G. et al. 2003. A statistical approach to evaluating distance metrics and analog assignments for pollen records. Quaternary Research 60:356–367.

References  267 Geological Survey of New Jersey.1900a. Annual Report of the State Geologist for 1899: Report on Forests. Geological Survey of New Jersey, Trenton. ———. 1900b. Forests of Northern New Jersey. 6 sheets. Special Collections, Rutgers University, New Brunswick. Georges-Leroy, M. et al. 2009. Le massif forestier, objet pertinent pour la r­echerche archéologique: L’exemple du massif forestier de Haye (Meurthe-et-Moselle). Revue Géographique de l’Est 49(2/3):2–18. Giesecke, T. and S. L. Fontana. 2008. Revisiting pollen accumulation rates from Swedish lake sediments. The Holocene 18:293–305. Gifford, J. 1900. Forestal conditions and silvicultural prospects of the coastal plain. Pp. 235–292 in Annual Report of the State Geologist for 1899, Geological Survey of New Jersey, Trenton. Giguet-Covex, C. et al. 2014. Long livestock farming history and human landscape shaping revealed by lake sediment DNA. Nature Communications 5:3211. doi: 10. 1038/ncomms4211. Gill, A. M., J. R. L. Hoare and N. P. Cheney. 1990. Fires and their effects in the wet-dry tropics of Australia. Pp. 159–177 in Fire in the Tropical Biota: Ecosystem Processes and Global Challenges, J. G. Goldammer, ed., Springer-Verlag, Berlin. Gill, J. L. et al. 2009. Pleistocene megafaunal collapse, novel plant communities, and enhanced fire regimes in North America. Science 326:1100–1103. ———. 2015. A 2.5-million-year perspective on coarse-filter strategies for conserving nature’s stage. Conservation Biology 29:640–648. Gimingham, C. H. and J. D. De Smidt. 1983. Heaths as natural and semi-natural vegetation. Pp. 185–199 in Holzner, Werger, and Ikusima 1983. Gimmi, U. et al. 2013. Soil carbon pools in Swiss forests show legacy effects from historic forest litter raking. Landscape Ecology 28:835–846. ———. 2016. Assessing accuracy of forest cover information on historical maps. Prace Geograficzne 146:7–18. Girona, M. M., L. Navarro and H. Morin. 2018. A secret hidden in the sediments: Lepidopteran scales. Frontiers in Ecology and Evolution 26:2. doi.org/10.3389/fevo .2018.00002. Glacken, C. 1967. Traces on the Rhodian Shore. University of California Press, Berkeley. Gleason, H. A. 1913. The relation of forest distribution and prairie fires in the Middle West. Torreya 13:173–181. Gleason, H. A. and A. Cronquist. 1991. Manual of the Vascular Plants of Northeastern United States and Adjacent Canada. New York Botanical Garden, New York. González-Guarda, E. et al. 2017. Late Pleistocene ecological, environmental and climatic reconstruction based on megafauna stable isotopes from northwestern Chilean Patagonia. Quaternary Science Reviews 170:88–202. Good, N. 1965. A study of natural replacement of chestnut in New Jersey. PhD diss., Rutgers University. Good, R. 1964. The Geography of the Flowering Plants, 3rd ed. John Wiley and Sons, New York. Goodland, R. 1995. The concept of environmental sustainability. Annual Review of Ecology and Systematics 26:1–24.

268  References Goodman, R. M. and K. Lancaster. 1990. Tsuga canadensis (L.) Carr. Eastern hemlock. Pp. 604–612 in Silvics of North America, vol. 1, Agriculture, R. M. Burns and B. H. Honkala, eds., Handbook 654, USDA Forest Service, Washington, DC. Goodman, S. M. and J. P. Benstead. 2005. Updated estimates of biotic diversity and endemism for Madagascar. Oryx 39:73–77. Goudie, A. 1990. The Human Impact on the Natural Environment, 3rd ed. MIT Press, Cambridge. ———. 2013. The Human Impact on the Natural Environment, 7th ed. John Wiley and Sons, Oxford, UK. Graham, R. W. 2016. Timing and causes of mid-Holocene mammoth extinction on St. Paul Island, Alaska. Proceedings of the National Academy of Science 113:9310–9314. Gras, N. S. B. 1930. The Economic and Social History of an English Village (Crawley, Hampshire), A.D. 1909–1928. Harvard University Press, Cambridge. Green, B. 1981. Countryside Conservation. Resource Management Series 3. George Allen and Unwin, London. Green, C. S. and R. J. Speller. 2017. Novel substrates as sources of ancient DNA: Prospects and hurdles. Genes 8(7). doi:10.3390/genes8070180. “The Green League Table.” 1993. Science 262:1815. Table taken from a report by the New Economics Foundation in London. Grigg, D. 1982. The Dynamics of Agricultural Change: The Historical Experience. St. Martin’s, New York. Grime, J. P. 1997. Biodiversity and ecosystem function: The debate deepens. Science 277:1260–1261. Grimm, N. B. et al. 2008. Global change and the ecology of cities. Science 319:756–760. Gross, J. E. et al. 2016. Adapting to climate change: Guidance for protected area managers and planners. Best Practice Protected Area Guidelines Series 24. International Union for the Conservation of Nature. portals.iucn.org/library/node/46685. Grove, A. T. and O. Rackham. 2001. The Nature of Mediterranean Europe: An Ecological History. Yale University Press, New Haven. Grove, J. M. et al. 2015. The Baltimore School of Urban Ecology: Space, Scale, and Time for the Study of Cities. Yale University Press, New Haven. Groves, R. J. and J. J. Burden, eds. 1986. Ecology of Biological Invasions. Cambridge University Press, New York. Guo, Q. et al. 2015. A unified approach for quantifying invasibility and degree of invasion. Ecology 96:2613–2621. Gurevitch, J. and D. K. Padilla. 2004. Are invasive species a major cause of extinctions? Trends in Ecology and Evolution 19:470–474, 619–620. Guthorn, P. G. 1966. American Maps and Mapmakers of the Revolution. Philip Freneau Press, Monmouth Beach, NJ. ———. 1972. British Maps of the American Revolution. Philip Freneau Press, Monmouth Beach, NJ. Guttmann-Bond, E. B. et al. 2016. Early Neolithic agriculture in county Mayo, Republic of Ireland: Geoarchaeology of the Céide fields, Belderrig, and Rathlackan. Journal of the North Atlantic 30:1–32.

References  269 Haber, L. F. 1958. The Chemical Industry during the Nineteenth Century. Oxford University Press, Oxford. Hadfield, M. G. and S. E. Miller. 1992. Ecology and conservation of endemic Hawai’ian tree snails (subfamily Achatinellinae). Bulletin Ecological Society of America 73(2):196. Haile, J. et al. 2009. Ancient DNA reveals late survival of mammoth and horse in interior Alaska. Proceedings of the National Academy of Sciences 106:22352–22357. Haleakalâ. 1990. Hosmer Grove Trail Guide. Haleakalâ National Park, National Park Service, U.S. Department of the Interior, n.p. Hall, A. D. 1905. The Book of the Rothamsted Experiments. J. Murray, London. Halligan, J. J. et al. 2016. Pre-Clovis occupation 14,550 years ago at the Page-Ladson site, Florida, and the peopling of the Americas. Science Advances 2:e1600375. Hamburg, S. P. and R. L. Sanford Jr. 1986. Disturbance, Homo sapiens, and ecology. Bulletin of the Ecological Society of America 67:169–171. Hammatt, R. F. 1949. Bad business; your business. P. 479 in Stefferud 1949. Hanson, J. H. 1992. Extractive economies in a historical perspective: Gum arabic in West Africa. Pp. 107–114 in Non-Timber Products from Tropical Forests: Evaluation of a Conservation and Development Strategy, D. C. Nepstad and S. Schwartzman, eds., Advances in Economic Botany 9, New York Botanical Garden, New York. Harlan, J. P. 1971. Agricultural origins: Centers and noncenters. Science 174:468–474. Harmer, R. et al. 2001. Vegetation changes during 100 years of development of two secondary woodlands on abandoned arable land. Biological Conservation 101:291–304. Hart, J. L. and M. L. Buchanan. 2006. History of fire in eastern oak forests and implications for restoration. Proceedings of the 4th Fire in Eastern Forests Conference GTR-NRS-P-102:34–51. Hartman, A. W. 1949. Fire, friend and enemy: Fire as a tool in southern pine. Pp. 517– 527 in Stefferud 1949. Hauge, L. 1988. Galdane, Iserdal, western Norway—Management and restoration of the cultural landscape. Pp. 31–45 in Birks et al. 1988. Haworth, E. Y. and J. W. G. Lund, eds. 1984. Lake Sediments and Environmental History: Studies in Palaeolimnology and Palaeoecology in Honour of Winifred Tutin. University of Minnesota Press, Minneapolis. Healy, M. J. and E. L. Jones. 1962. Wheat yields in England, 1815–1849. Journal of the Royal Statistical Society, ser. A, 125:574–579. Healy, W. M. and W. J. McShea. 2002. Goals and guidelines for managing oak ecosystems for wildlife. Pp. 333–341 in Oak Forest Ecosystems: Ecology and Management for Wildlife, W. M. Healy and W. J. McShea, eds., Johns Hopkins University Press, Baltimore. Heide, K. and R. Bradshaw. 1982. The pollen-tree relationship within forests of Wisconsin and upper Michigan, USA. Review of Palaeobotany and Palynology 36:1–24. Heinselman, M. L. 1981. Fire intensity and frequency as factors in the distribution and structure of northern ecosystems. Pp. 7–57 in Mooney et al. 1981. Heizer, R. F. 1955. Primitive man as an ecological factor. Kroeber Anthropological Society Papers 13:1–31.

270  References Hempson, G. P., S. Archibald and W. J. Bond. 2015. A continent-wide assessment of the form and intensity of large mammal herbivory in Africa. Science 350:1056–1061. Henberg, M. 1994. Wilderness, myth and American character. Key Reporter 11(4):7–11. Henne, P. D. et al. 2013. Impacts of changing climate and land use on vegetation dynamics in a Mediterranean ecosystem: Insights from paleoecology and dynamic modeling. Landscape Ecology 28:819–833. Henry, F., B. Talon and T. Dutoit. 2010. The age and history of the French Mediterranean steppe revised by soil wood charcoal analysis. The Holocene 20:25–34. Herlihy, D. J. 1980. Attitudes toward the environment in medieval society. Pp. 100–116 in Bilsky 1980. Hersberger, A. M. and M. Bürgi. 2009. Going beyond landscape change description: Quantifying the importance of driving forces of landscape change in a central Europe case study. Land Use Policy 26:640–648. Hewson, I. et al. 2014. Densovirus associated with sea-star wasting disease and mass mortality. Proceedings of the National Academy of Science 111:17278–17283. Hiers, J. K. et al. 2016. The precision problem in conservation and restoration. Trends in Ecology and Evolution 31:820–830. Highlands Study Team. 1992. New York–New Jersey Highlands Regional Study. United States Forest Service, n.p. Hillebrand, H. et al. 2007. Consumer versus resource control of producer diversity depends on ecosystem type and producer community structure. Proceedings of the National Academy of Sciences 104:10904–10909. Hirschfeld, A. and A. Heyd. 2005. Mortality of migratory birds caused by hunting in Europe: Bag statistics and proposals for the conservation of birds and animal welfare. Berichte zum Vogelschutz 42:47–74. Hobbes, T. 1909. Leviathan. Oxford University Press, Oxford. (Originally published in 1651.) Hodell, D. A. et al. 2001. Solar forcing of drought frequency in the Maya lowlands. Science 292:1376–1380. Holzner, W., M. J. A. Werger and I. Ikusima, eds. 1983. Man’s Impact on Vegetation. Dr. W. Junk, The Hague. Hooke, D. 1988. Woodland utilization in England, AD 800–1100. Pp. 301–310 in Salbitano 1988. Hoorn, C. and S. Flantua. 2015. An early start for the Panama land bridge. Science 348:186–187. Horn, S. 1992. Microfossils and forest history in Costa Rica. Pp. 16–30 in Changing Tropical Forests: Historical Perspectives on Today’s Challenges in Central and South America, H. K. Steen and R. P. Tucker, eds., Forest History Society, n.p. Hörnberg, G. et al. 2006. Effects of Mesolithic hunter-gatherers on local vegetation in a non-uniform glacio-isostatic land uplift area, northern Sweden. Vegetation History and Archaeobotany 15:13–26. Hough, A. F. 1965. A twenty-year record of understory vegetational change in a virgin Pennsylvania forest. Ecology 46:370–373. Hough, F. B. 1878. Report upon Forestry, vol. 1. U.S. Department of Agriculture, Washington, DC.

References  271 ———. 1882. Report upon Forestry, vol. 3. U.S. Department of Agriculture, Washington, DC. Hough, W. 1926. Fire as an agent in human culture. United States National Museum Bulletin 139. Howell, E. 1828–1831. Account Books, Harrisonville, NJ. Special Collections, Alexander Library, Rutgers University, New Brunswick, NJ. Hu, F. S., A. Hampe and R. J. Petit. 2009. Paleoecology meets genetics: Deciphering past vegetational dynamics. Frontiers in Ecology and the Environment 7:371–379. Hubbell, S. T. 2006. Neutral theory and the evolution of ecological equivalence. Ecology 87:1387–1398. Hughes, J. and B. Huntley. 1988. Upland hay meadows in Britain—Their vegetation, management and future. Pp. 91–110 in Birks et al. 1988. Hughes, J. D. 2011. Ancient deforestation revisited. Journal of the History of Biology 44:43–57. Humphrey, R. R. 1987. Ninety Years and 535 Miles: Vegetation Changes along the Mexican Border. University of New Mexico Press, Albuquerque. Hung, C-M. et al. 2014. Drastic population fluctuations explain the rapid extinction of the passenger pigeon. Proceedings of the National Academy of Sciences 111:10636– 10641. Hunt, T. L. and C. P. Lipo. 2007. Chronology, deforestation, and “collapse”: Evidence vs. faith in Rapa Nui prehistory. Rapa Nui Journal 21:85–97. Hunter, M. L., Jr., T. L. Jacobson, Jr. and T. Webb III. 1988. Paleoecology and the coarse-filter approach to maintaining biological diversity. Conservation Biology 2:375–385. Huntley, B. and H. J. B. Birks. 1983. An Atlas of Past and Present Pollen Maps for Europe: 0–13 000 Years Ago. Cambridge University Press, Cambridge. Huntley, B. and T. Webb III, eds. 1988. Vegetation History. Kluwer Academic, Dordrecht, Netherlands. Hunziker, M. 1993. Wenn wiesen zu Wald werden . . . ein Verlust für das Landschaftserlebnis? Informationsblatt des Forschungsbeiches Landschaftsökologie 18:1–2. Hurst, J. W. 1983. The institutional environment of the logging era in Wisconsin. Pp. 137–155 in Flader 1983. Huston, M. 1993. Biological diversity, soils, and economics. Science 262:1676–1680. Hutchinson, G. E. 1957. Concluding remarks. Population Studies: Animal Ecology and Demography. Cold Spring Harbor Symposium on Quantitative Biology 22:415– 427. ———. 1959. Homage to Santa Rosalia; or, Why are there so many kinds of animals? American Naturalist 93:1145–1159. Ingvild, A. 1988. Tree-pollarding in western Norway. Pp. 11–29 in Birks et al. 1988. Innes, J. B. and J. J. Blackford. 2003. The ecology of late Mesolithic woodland disturbances: Model testing with fungal spore assemblage data. Journal of Archaeological Science 30:185–194. Irland, L. C. 1986. Rufus Putnam’s ghost: An essay on Maine’s public lands, 1783– 1820. Journal of Forest History 30:60–69. Isaac, E. 1970. Geography of Domestication. Prentice-Hall, Englewood Cliffs, NJ.

272  References Isenberg, Andrew C., ed. 2014. The Oxford Handbook of Environmental History. Oxford University Press, Oxford. Iversen, J. 1941. Land occupation in Denmark’s Stone Age. Danmarks Geologiske Undersøgelse, Raekke II 66:1–68. ———. 1956. Forest clearance in the stone age. Scientific American 194:36–41. Iverson, L. R. 1988. Land-use changes in Illinois, USA: The influence of landscape attributes on current and historic land use. Landscape Ecology 2:45–61. Jablonski, D. 1991. Extinction: A paleontological perspective. Science 253:754–757. Jackson, A. L. 2011. Comparing isotopic niche widths among and within communities: SIBER-Stable Isotope Bayesian Ellipses in R. Journal of Animal Ecology 80:595–602. Jackson, J. B. C. et al. 2001. Historical overfishing and the collapse of coastal ecosystems. Science 293:629–637. Jackson, S. T. and D. F. Sax. 2010. Balancing biodiversity in a changing environment: Extinction debt, immigration credit and species turnover. Trends in Ecology and Evolution 25:153–160. Jackson, S. T. and J. W. Williams. 2004. Modern analogs in Quaternary paleoecology: Here today, gone yesterday, gone tomorrow? Annual Review of Earth and Planetary Sciences 32:495–537. Jackson, W. and J. Piper. 1989. The necessary marriage between ecology and agriculture. Ecology 70:1591–1593. Jacobs, J. F. and R. L. Bartgis. 1987. The running buffalo clover. Pp. 438–43 in Audubon Wildlife Report, R. L. DiSilvestris, ed., Academic Press, Orlando. Jacobson, G. L., Jr. and R. H. W. Bradshaw. 1981. The selection of sites for paleo­ vegetational studies. Quaternary Research 16:80–96. Johnson, A. L. 2014. Exploring adaptive variation among hunter-gatherers with Binford’s frames of reference. Journal of Archaeological Research 22:1–42. Johnson, H. B. 1976. Order upon the Land: The U.S. Rectangular Land Survey and the Upper Mississippi Country. Oxford University Press, New York. Johnson, W. C. and C. S. Atkisson. 1985. Dispersal of beech (Fagus grandifolia) nuts by blue jays (Cyanocitta cristata) in fragmented landscapes. American Midland Naturalist 113:319–324. Johnson, W. C. and T. Webb III. 1989. The role of blue jays (Cyanocitta cristata L.) in the postglacial dispersal of fagaceous trees in eastern North America. Journal of Biogeography 16:561–571. Josefsson, T., J. Olsson and L. Östlund. 2010. Linking forest history and conservation efforts: Long-term impact of low-intensity timber harvest on forest structure and wood-inhabiting fungi in northern Sweden. Biological Conservation 143:1803–1811. Josefsson, T. et al. 2010. Historical human influence on forest composition and structure in boreal Fennoscandia. Canadian Journal of Forest Research 40:872–884. Jouy-Avantin, F. et al. 2003. A standardized method for the description and the study of coprolites. Journal of Archaeological Science 30:367–372. Julius, S. H. et al. 2013. Climate Change and the U.S. Natural Resources: Advancing the Nation’s Capability to Adapt. Issues in Ecology 18. Ecological Society of America, n.p. Kagan, D., S. Ozment and F. M. Turner. 1979. The Western Heritage. Macmillan, New York.

References  273 Kain, R. J. P. and J. M. Hooke. 1982. Historical Change in the Physical Environment: A Guide to Sources and Techniques. Butterworth Scientific, London. Kaiser, J. 2000. Biodiversity divides ecologists. Science 289:1282–1283. Kalm, P. 1937. Peter Kalm’s Travels in North America: The English Version of 1770, trans. and ed. A. B. Benson, vol. 2. Wilson-Erickson, New York. Karkanas, P. 2007. Evidence for habitual use of fire at the end of the Lower Paleolithic: Site-formation processes at Qesem Cave, Israel. Journal of Human Evolution 53:197–212. Keeley, J. E. 1986. Resilience of Mediterranean shrub communities to fires. Pp. 95–112 in Resilience in Mediterranean-type Ecosystems, B. Dell, A. J. M. Hopkins, and B. B. La­ mont, eds., Dr. W. Junk, Dordrecht, Netherlands. Keeley, J. E. and P. W. Rundel. 2005. Fire and the Miocene expansion of C4 grasslands. Ecology Letters 8:683–690. Kemble, P. 1780–1785. Mt. Kemble, NJ, Diary, 21 April 1780–25 December 1785. Special Collections, Alexander Library, Rutgers University, New Brunswick, NJ. Kemp, W. M. et al. 2005. Eutrophication of Chesapeake Bay: Historical trends and ecological interactions. Marine Ecology Progress Series 303:1–29. Kercheval, S. 1833. A History of the Valley of Virginia. S. H. Davis, Winchester, VA. Kerkkonen, M. 1959. Peter Kalm’s North American Journey: Its Ideological Background and Results. N.p., Helsinki. Kidder, T. R. 2006. Climate change and the Archaic to Woodland transition (3000–2500 Cal B.P.) in the Mississippi River Basin. American Antiquity 71:195–231. Kipfmueller, K. F. and T. W. Swetnam. 2005. Using dendrochronology to reconstruct the history of forest and woodland ecosystems. Pp. 199–228 in Egan and Howell 2005. Kirby, K. J. 1988. Conservation in British woodland—Adapting traditional management to modern needs. Pp. 79–89 in Birks et al. 1988. Kirch, P. V. 2007. Hawaii as a model system for human ecodynamics. American Anthropologist 109:8–26. Kirchner, G. 2011. 210 Pb as a tool for establishing sediment chronologies: Examples of potentials and limitations of conventional dating models. Journal of Environmental Radioactivity 102:490–494. Kirwan, M. L. et al. 2011. Rapid wetland expansion during European settlement and its implication for marsh survival under modern sediment delivery rates. Geology 39:507–510. Kitchell, J. F. and S. R. Carpenter. 1993. Variability in lake ecosystems: Complex responses by the apical predator. Pp. 125–140 in McDonnell and Pickett 1993. Klein, J. 2004. Deforestation in the Madagascar highlands—Established “truth” and scientific uncertainty. GeoJournal 56:191–199. Klett, M. et al. 1984. The Rephotographic Survey Project. University of New Mexico Press, Albuquerque. Knowles, A. K. 2014. Why we must make maps: Historical geography as a visual craft. Historical Geography 42:3–26. Knowles, A. K. and A. Hillier. 2008. Placing History: How Maps, Spatial Data, and GIS Are Changing Historical Scholarship. ESRI, Redlands, CA.

274  References Kornas, J. 1983. Man’s impact upon the flora and vegetation of Central Europe. Pp. 277–286 in Holzner, Werger and Ikusima 1983. Korstian, C. F. and P. W. Stickel. 1927. The natural replacement of blight-killed chestnut in hardwood forests of the northeast. Journal of Agricultural Research 34:631–648. Krech, S., III. 1998. Ecology, conservation, and the buffalo jump. Pp. 139–164 in Stars Above, Earth Below: American Indians and Nature, M. Bol, ed., Roberts Rinehart, Niwot, CO. Kreisler, F. F. 1984. Domesday Book. Pages 237–239 in Dictionary of the Middle Ages, J. R. Strayer, ed., Charles Scribner’s Sons, New York. Krug, E. C. and C. R. Frink. 1983. Acid rain on acid soil: A new perspective. Science 221:520–525. Krzywinski, K., M. O’Connell and K. Küster. 2009. Cultural Landscapes of Europe: Fields of Demeter, Haunts of Pan. Aschenbeck Media, Bremen. Küchler, A. W. 1964. Potential Natural Vegetation of the Conterminous United States. Special Publication 36. American Geographical Society, New York. Kueffer, C. 2017. Plant invasions in the Anthropocene. Science 358:724–725. Kull, K. and M. Zobel. 1991. High species richness in Estonian wooded meadow. Journal of Vegetation Science 2:715–718. Kummel, B. and D. M. Raup. 1965. Handbook of Paleontological Techniques. W. H. Freeman, San Francisco. Lambert, A. M. 1971. The Making of the Dutch Landscape: An Historical Geography of the Netherlands. Academic Press, London. Landa, M. 1988. Ecological changes in the history of forestry and game management in Czechoslovakia, especially southern Bohemia. Pp. 287–299 in Salbitano 1988. Langgut, D., I. Finkelstein and T. Litt. 2013. Climate and the Late Bronze collapse: New evidence from the Southern Levant. Tel Aviv 40:149–175. Larson, G. and D. Q. Fuller. 2014. The origin of animal domestication. Annual Review of Ecology, Evolution, and Systematics 45:115–136. Latham, R. E. and R. E. Ricklefs. 1993. Continental comparisons of temperate-zone tree species diversity. Pp. 294–314 in Species Diversity in Ecological Communities: Historical and Geographical Perspectives, R. E. Ricklefs and D. Schluter, eds., University of Chicago Press, Chicago. Leal, A. et al. 2016. Late-Holocene gallery forest retrogression in the Venezuelan Guayana: New data and implications for the conservation of a cultural landscape. The Holocene 26:1049–1063. Lefébvre, J. G. 1879. Les forêts de l’Europe et de l’Amérique: Etude sur le régime des forêts et leur reconstruction (Paris and Havre, 1879), quoted in Hough 1882. Leopold, A. 1933. Game Management. Charles Scribner’s Sons, New York. Lever, C. 1985. Naturalized Mammals of the World, Longman, London. Levine, J. M. et al. 2003. Mechanisms underlying the impacts of exotic plant in­vasions. Proceedings of the Royal Society of London B: Biological Sciences 270:775–781. Lewis, H. T. 1982. Fire technology and resource management in aboriginal North America and Australia. Pp. 45–68 in Resource Managers: North American and Australian Hunter-Gatherers, N. M. Williams and E. S. Hunn, eds., AAAS Selected Symposium 67, Westview, Boulder, CO.

References  275 Lewis, S. L. and M. A. Maslin 2015. Defining the Anthropocene. Nature 519:171–180. Likens, G. E., ed. 1989. Long-Term Studies in Ecology. Springer-Verlag, New York. Likens, G. E. and M. B. Davis. 1975. Postglacial history of Mirror Lake and its watershed in New Hampshire, USA: An initial report. Verhandlungen der Internationaler Vereinigung für Theoretischen und angewandte Limnologie 19:982–993. Lillesand, T., R. W. Kiefer and J. Chipman. 2015. Remote Sensing and Image Interpretation, 7th ed. John Wiley and Sons, New York. Limbrey, S. 1975. Soil Science and Archaeology. Academic Press, London. Lindborg, R. and O. Eriksson. 2004. Historical landscape connectivity affects present plant species diversity. Ecology 84:1840–1845. Lindeström, P. M. 1925. Geographia Americae, with an Account of the Delaware Indians, A. Johnson, trans. Swedish Colonial Society, Philadelphia. (Written in 1691 from notes made in 1654–1656.) Little, E. L., Jr. 1950. Important forest trees of the United States. Pp. 763–814 in Trees, Yearbook of Agriculture, USDA Government Printing Office, Washington, DC. Loeb, R. E. 1982. Reliability of the New York City Department of Parks and Recreation’s forest records. Bulletin of the Torrey Botanical Club 109:537–541. Long, J. L. 2003. Introduced Mammals of the World: Their History, Distribution and Influence. CSIRO, Cambridge, MA. Loope, W. L. 1991. Interrelationships of fire history, land use history, and landscape pattern within Pictured Rocks National Lakeshore, Michigan. Canadian Field Naturalist 105:18–28. Loran, C., F. Kienast and M. Bürgi. 2018. Change and persistence: Exploring the driving forces of long-term forest cover dynamics in the Swiss lowlands. European Journal of Forests Research 137:693–706. Lorenzen, E. D. et al. 2011. Species-specific responses of Late Quaternary megafauna to climate and humans. Nature 479:359–364. Lorimer, C. G. 1985. Methodological considerations in the analysis of forest disturbance history. Canadian Journal of Forest Research 15:200–213. ———. 1989. The oak regeneration problem: New evidence on causes and possible solutions. Forest Resources Analyses, #8. Department of Forestry, University of Wisconsin, Madison. Lotze, H. K. and L. McClenachan. 2013. Marine historical ecology: Informing the future by learning from the past. Pp. 165–200 in Marine Community Ecology and Conservation, M. Bertness et al., eds., Oxford University Press, Oxford. Lowdermilk, W. C. 1975. Conquest of land through 7000 Years: Dr. Lowdermilk’s trip after the dustbowl. USDA, Agricultural Information Bulletin #99. Lowe, D.J. 2011. Tephrochronology and its application: A review. Quaternary Geochronology 6:107–153. Lubchenko, J. 1991. The sustainable biosphere initiative: An ecological research agenda. Ecology 72:371–412. Ludwin, R. S. 2005. Dating the 1700 Cascadia earthquake: Great coastal earthquakes in native stories. Seismological Research Letters 76:140–148. Luh, H.-K. and S. L. Pimm 1993. The assembly of ecological communities: A minimalist approach. Journal of Animal Ecology 62:749–765.

276  References Lunt, I. D. and P. G. Spooner. 2005. Using historical ecology to understand patterns of biodiversity in fragmented agricultural landscapes. Journal of Biogeography 32:1859–1873. Lutz, H. J. 1928. Trends of silvicultural significance of upland forest succession in southern New England. Yale School of Forestry Bulletin 22:1–68. MacDougall, A. S. R. and R. Turkington. 2005. Are invasive species the drivers or passengers of change in degraded ecosystems? Ecology 86:42–55. Mack, R. N. 1981. Invasion of Bromus tectorum L. into western North America: An ecological chronicle. Agro-Ecosystems 7:145–165. Mack, R. N. and W. M. Lonsdale. 2001. Humans as global plant dispersers: Getting more than we bargained for: Current introductions of species for aesthetic purposes present the largest single challenge for predicting which plant immigrants will become future pests. BioScience 51:95–102. Maenza-Gmelch, T. E. 1997. Holocene vegetation, climate, and fire history of the Hudson Highlands, southeastern New York, USA. The Holocene 7:25–37. Maezumi, S. Y. et al. 2017. Reassessing climate and preColumbian drivers of paleo­fire activity in the Bolivian Amazon. Quaternary International. doi.org/10.1016/j.quaint .2017.11.053. Mánˇez, K. S. et al. 2014. The future of the oceans past: Towards a global marine historical research initiative. PLOS One 9(7):e101466. Mann, M. E. et al. 2009. Global signatures and dynamical origins of the Little Ice Age and medieval climate anomaly. Science 326:1256–1260. Mark, A. F. and G. D. McSweeney. 1990. Patterns of impoverishment in natural communities: Case history studies in forest ecosystems—New Zealand. Pp. 151–176 in Woodwell 1990. Marks, P. L. 1983. On the origin of the field plants of the northeastern United States. American Naturalist 122:210–228. Marlon, J. R. et al. 2013. Global biomass burning: A synthesis and review of Holocene paleofire records and their controls. Quaternary Science Reviews 65:5–25. Marsh, G. P. 1864. Man and Nature. (Reprinted by Belknap Press of Harvard University, Cambridge, 1965.) ———. 1882. Letter to C. S. Sargent, 20 July, Marsh Collection, University of Vermont. P. 261n223 in G. P. Marsh, Man and Nature, Belknap Press of Harvard University, Cambridge, 1965. Marshall, L. J. 1988. Land mammals and the great American interchange. American Scientist 76:380–388. Martelli, C. A. 2015. Ducks and deer, profit and pleasure: Hunters, game and the natural landscapes of medieval Italy. PhD diss., York University. Martin, I. L. and T. C. Bass. 1940. Erosion and Related Land Use Conditions on the Lake Michie Watershed, Near Durham, North Carolina. USDA, Washington, DC. Martin, P. H., C. D. Canham and P. L. Marks. 2009. Why forests appear resistant to exotic plant invasions: Intentional introductions, stand dynamics, and the role of shade tolerance. Frontiers in Ecology and the Environment 7(3):142–149. Martin, P. S. 1967. Prehistoric overkill: Pleistocene extinctions. Proceedings VIIth Congress of INQUA 6:75–129.

References  277 ———. 1984. Prehistoric overkill: The global model. Pp. 354–403 in Quaternary Extinctions: A Prehistoric Revolution, P. S. Martin and R. G. Klein, eds., University of Arizona Press, Tucson. Martin, W. K. 1965. The Concise British Flora in Colour. Ebury Press and Michael Joseph, Norwich. Maxwell, H. 1910. The use and abuse of forests by the Virginia Indians. William and Mary College Quarterly Historical Magazine 19:73–103. Mayewski, P. A. et al. 2004. Holocene climate variability. Quaternary Research 62:243– 255. McAndrews, J. H. 1988. Human disturbance of North American forests and grasslands: The fossil pollen record. Pp. 673–97 in Huntley and Webb 1988. McArthur, R. H. and E. O. Wilson 1963. An equilibrium theory of insular zoogeography. Evolution 17:373–387. McCabe, R. E. and T. R. McCabe. 1984. Of slings and arrows: An historical retrospective. Pp. 19–72 in White-Tailed Deer: Ecology and Management, Stackpole Books, Mechanicsburg, PA. McCallum, H. T. and E. D. McCallum. 1965. The Wire That Fenced the West. University of Oklahoma Press, Norman. McCaughey, W. W. 1982. Understory tree release following harvest cutting in sprucefir forests of the Intermountain West. U.S. Department of Agriculture Forest Service Research Paper INT-285. McClenachan, L. et al. 2017. Ghost reefs: Nautical charts document large spatial scale of coral reef loss over 240 years. Science Advances 3:e1603155. McClintock, D. and R. S. R. Fitter. 1964. Guide des Plantes à Fleurs de l’Europe Occidentale. Delachaux et Niestlé, Neuchatal, Switzerland. McDonnell, M. J. and S. T. A. Pickett 1990. Ecosystem structure and function along gradients of urbanization: An unexploited opportunity for ecology. Ecology 71:1232– 1237. ———, eds. 1993. Humans as Components of Ecosystems: The Ecology of Subtle Human Effects and Populated Areas. Springer-Verlag, New York. McDonnell, M. J., S. T. A. Pickett and R. V. Pouyat. 1993. The application of the ecological gradient paradigm to the study of urban effects. Pp. 175–89 in McDonnell and Pickett 1993. McDonnell, M. J. and E. W. Stiles. 1983. The structural complexity of old-field vegetation and the recruitment of bird-dispersed plant species. Oecologia 56:109–116. McEvoy, A. F. 1986. The Fisherman’s Problem: Ecology and Law in the California Fisheries, 1850–1980. Cambridge University Press, Cambridge. McEwan, R. W., J. M. Dyer and N. Pederson. 2011. Multiple interacting ecosystem drivers: Toward an encompassing hypothesis of oak forest dynamics across eastern North America. Ecography 34:244–256. McGovern, T. H. et al. 2007. Landscapes of settlement in northern Iceland: Historical ecology of human impact and climate fluctuation on the millennial scale. American Anthropologist 109:27–51. McInteer, B. B. 1946. A change from grassland to forest vegetation in the “Big Barrens” of Kentucky. American Midland Naturalist 35:276–282.

278  References McKinney, M. L. 2008. Effect of urbanization on species richness: A review of plants and animals. Urban Ecosystems 11:161–176. McNeill, J. R. 1988. Deforestation in the araucaria zone of southern Brazil, 1900–1983. Pp. 15–32 in World Deforestation in the Twentieth Century, J. F. Richards and R. P. Tucker, eds., Duke University Press, Durham, NC. ———. 2000. Something New under the Sun: An Environmental History of the ­Twentieth-Century World. W. W. Norton, New York. McNutt, M. 2013. Risk. Science 341:109. McWilliams, W. J. et al. 2002. Distribution and abundance of oaks in North America. Pp. 13–34 in Oak Forest Ecosystems: Ecology and Management for Wildlife, W. M. Healy and W. J. McShea, eds., Johns Hopkins University Press, Baltimore. Meiners, S. J. 2007. Native and exotic plant species exhibit similar population dynamics during succession. Ecology 88:1098–1104. Meinig, D. W. 1993. The Shaping of America: A Geographic Perspective on 500 Years of History, vol. 2, Continental America, 1800–1867. Yale University Press, New Haven. Merchant, C. 1985. The women of the progressive conservation crusade. Pp. 153–175 in Bailes 1985a. Métailié, J. P., J. Bonhote and C. Frauhauf. 1988. A thousand years of forest history in the French Pyrenees Mountains: The Ariège example. Pp. 159–167 in Salbitano 1988. Mikan, C. J., D. O. Orwig and M. D. Abrams. 1994. Age structure and successional dynamics of a presettlement-origin chestnut oak forest in the Pennsylvania piedmont. Bulletin of the Torrey Botanical Club 111:13–23. Milbraith, L. W. 1985. Culture and environment in the United States. Environmental Management 9:161–172. Miller, G. H. et al. 2005. Ecosystem collapse in Pleistocene Australia and a human role in megafaunal extinction. Science 309:287–290. Millers, I., D. S. Shriner and D. Rizzo. 1989. History of Hardwood Decline in the Eastern United States. USDA Forest Service, General Technical Report NE-126. Milton, K. 2000. Hunter-gatherer diets—A different perspective. Journal of Archaeological Research 71:665–667. Mitchell, L. et al. 2013. Constraints on the late Holocene anthropogenic contribution to the atmospheric methane budget. Science 342:964–966. Mittelbach, G. G. et al. 2007. Evolution and the latitudinal diversity gradient: Speciation, extinction and biogeography. Ecology Letters 10:315–331. Mooney, H. A. and J. A. Drake, eds. 1986. Ecology of Biological Invasions of North America and Hawaii. Springer-Verlag, New York. Mooney, H. A., S. P. Hamburg and H. A. Drake. 1986. The invasion of plants and animals into California. Pp. 250–272 in Mooney and Drake 1986. Mooney, H. A. et al., eds. 1981. Fire Regimes and Ecosystem Properties. USDA Forest Service, General Technical Report WO-26. Moore, P. D. and E. H. Chater. 1969. The changing vegetation of west-central Wales in the light of human history. Journal of Ecology 57:361–379. Moore, P. D., T. A. Webb, and M. E. Collinson. 1991. Pollen Analysis. Blackwell Scientific, Oxford. Morales, M. R. et al. 2017. Exploring habitat diversity of mid-Holocene hunter-gath-

References  279 erers in the South-Central Andes: Multi-proxy analysis of Cruces Core 1 (TC1), Dry Puna of Jujuy, Argentina. Journal of Archaeological Science 18:708–721. Moreno-Mateos, E. et al. 2012. Structural and functional loss in restored wetland ecosystems. PLoS Biology 10(1):e1001247 Morris, M. G. and F. H. Perring, eds. 1974. The British Oak: Its History and Natural History. Classey, Farringdon, UK. Moss, S. R. et al. 2004. Symposium: The Broadbalk long-term Experiment at Rothamsted: What has it told us about weeds? Weed Science 52:864–873. Motzkin, G. et al. 2004. Forest landscape patterns, structure, and composition. Pp. 172–188 in Foster and Aber 2004. Moulton, M. P. and S. L. Pimm. 1986. Species introductions to Hawaii. Pp. 231–249 in Mooney and Drake 1986. Moulton, M. P. et al. 2010. The earliest house sparrow introductions to North America. Biological Invasions 12:2955–2958. Moyle, P. B. 1986. Fish introductions into North America: Patterns and ecological impact. Pp. 27–43 in Mooney and Drake 1986. Mumford, L. 1956. The natural history of urbanization. Pp. 382–398 in Thomas 1956. Munoz, S. E. et al. 2014. A record of sustained prehistoric and historic land use from the Cahokia region, Illinois, USA. Geology 42:499–502. Murray, G. B. R. et al. 2017. Natural selection shaped the rise and fall of passenger pigeon genomic diversity. Science 358:951–954. Mutch, R. W. 1970. Wildland fires and ecosystems—A hypothesis. Ecology 51:46–51. Nabhan, G. P. and M. K. Anderson. 1941. Gardeners in Eden. Wilderness 55(194):30. Naeem, S. et al. 1999. Biodiversity and Ecosystem Functioning: Maintaining Natural Life Support Processes. Issues in Ecology 4. Ecological Society of America, n.p. Nakagawa, T. et al. 2002. Quantitative pollen-based climate reconstruction in central Japan: Application to surface and Late Quaternary spectra. Quaternary Science Reviews 21:2099–2113. Nash, R. 1982. Wilderness and the American Mind. Rev. ed. Yale University Press, New Haven. National Research Council. 1992. Global Environmental Change: Understanding the Human Dimensions. National Academy Press. Report quoted in Anonymous 1992. Naveh, Z. and Y. Carmel. 2004. The evolution of the cultural Mediterranean landscape in Israel as affected by fire, grazing, and human activities. Pp. 337–409 in Evolutionary Theory and Processes: Modern Horizons, S. P. Wasser, ed., Springer, Dordrecht. Naveh, Z. and P. Kutiel. 1990. Changes in Mediterranean vegetation of Israel in response to human habitation and land use. Pp. 259–299 in Woodwell 1990. Nelle, O. 2003. Woodland history of the last 500 years revealed by anthracological studies of charcoal kiln sites in the Bavarian forest, Germany. Phytocoenologia 33:667– 682. Nelson, E. et al. 2009. Modeling multiple ecosystem services, biodiversity conservation, commodity production, and tradeoffs at landscape scales. Frontiers in Ecology and the Environment 7:4–11. Nelson, W., ed. 1894. Documents Relating to the Colonial History of the State of New Jersey. Series 1, vol. 11. Press Printing and Publishing, Patterson, NJ.

280  References Nepstad, D. C. et al. 1996. A comparative study of tree establishment in abandoned pasture and mature forest of eastern Amazonia. Oikos 76:25–39. New Jersey State Geologist. 1903. Report of the State Geologist for the Year 1902. Geological Survey of New Jersey, Trenton. Newton, A. C. et al. 2009. Impacts of grazing on lowland heathland in north-west Europe. Biological Conservation 142:935–947. Newton, N. T. 1971. Design of the Land: The Development of Landscape Architecture. Harvard University Press, Cambridge. Nguyen, H. H. et al. 2016. A review of the drivers of 200 years of wetland degradation in the Mekong Delta of Vietnam. Regional Environmental Change, 16:2303–2315. Nicholas, G. P. 1988. Introduction: Human behavior and Holocene ecology. Pp. 1–7 in Holocene Human Ecology in Northeastern North America, G. P. Nicholas, ed., Plenum, New York. Nichols, E. et al. 2008. Ecological functions and ecosystem services provided by Scarabaeinae dung beetles. Biological Conservation 141:1461–1474. Nielsen, A.B. and B. V. Odgaard. 2010. Quantitative landscape dynamics in Denmark through the last three millennia based on the Landscape Reconstruction Algorithm approach. Vegetation History and Archaeobotany 19:375–387. Niering, W. A. and R. H. Goodwin. 1962. Ecological studies of the Connecticut Arboretum natural area, I: An introduction and a survey of vegetation types. Ecology 43:41–54. Niering, W. A., R. H. Goodwin and S. Taylor. 1970. Prescribed burning in southern New England: Introduction to long-range studies Proceedings of the Annual Tall Timbers Fire Ecology Conference 8:267–286. Niklasson, M. and A. Granström. 2000. Numbers and sizes of fires: Long-term spatially explicit fire history in a Swedish boreal landscape. Ecology 81:1484–1499. Normile, D. 2016. Nature from nurture. Science 351:908–910. Noss, R. F. et al. 2015. How global biodiversity hotspots may go unrecognized: Lessons from the North American Coastal Plain. Diversity and Distributions 21:236–244. Nowak, R. S., C. L. Nowak and R. J. Tausch. 2017. Vegetation dynamics during last 35,000 years at a cold desert locale: Preferential loss of forbs with increased aridity. Ecosphere 8:2–23. Nye, J. W. et al. 2018. Cumulative human impacts on Pinnipeds over the last 7,500 years in southern South America. SAA Archaeological Record 18:47–52. O’Connell, M. K. and Molloy, K. 2001. Farming and woodland dynamics in Ireland during the Neolithic. Biology and Environment: Proceedings of the Royal Irish Academy 101B:99–128. Ohmann, L. H. and M. F. Buell. 1968. Forest vegetation of the New Jersey highlands. Bulletin of the Torrey Botanical Club 95:287–298. Oliveira, J. M. and V. D. Pillar. 2004. Vegetation dynamics on mosaics of Campos and Araucaria forests between 1974 and 1999 in southern Brazil. Community Ecology 5:197–202. Oliver, C. D. and E. P. Stephens. 1977. Reconstruction of a mixed-species forest in central New England. Ecology 58:562–572. Olsen, K. M. and J. F. Wendel. 2013. A bountiful harvest: Genomic insights into crop domestication phenotypes. Annual Review of Plant Biology 64:47–70.

References  281 Olson, G. W. 1981. Soils and the Environment. Chapman and Hall, New York. Olson, S. 1971. The Depletion Myth: A History of Railroad Use of Timber. Harvard University Press, Cambridge. Olson, S. L. and H. F. James. 1982. Fossil birds from the Hawaiian Islands: Evidence for wholesale extinction by man before Western contact. Science 117:633–635. Oosting, H. J. 1942. An ecological analysis of the plant communities of piedmont, North Carolina. American Midland Naturalist 28:1–126. Opie, J. 1985. Environmental history: Pitfalls and opportunities. Pp. 22–35 in Bailes 1985a. ———. 1987. The Law of the Land: Two Hundred Years of American Farmland Policy. University of Nebraska Press, Lincoln. Oreskes, N., K. Shrader-Freshette and K. Belitz. 1994. Verification, validation, and confirmation of numerical models in the earth sciences. Science 263:641–648. Orians, G. H. 1986. Site characteristics favoring invasions. Pp. 133–148 in Mooney and Drake 1986. Östlund, L. et al. 2009. Bark-peeling, food stress and tree spirits—The use of pine inner bark for food in Scandinavia and North America. Journal of Ethnobiology 29:94–112. ———. 2015. Intensive land use in the Swedish mountains between AD 800 and 1200 led to deforestation and ecosystem transformation with long-lasting effects. Ambio 44:508–520. Oswald, W. W. et al. 2007. Post-glacial changes in spatial patterns of vegetation across southern New England. Journal of Biogeography 34:900–913. Oswalt, S. N. et al. 2014. Forest Resources of the United States, 2012: A Technical Document Supporting the Forest Service; 2010 Update of the RPA Assessment. U.S. Forest Service General Technical Report WO-91. Overpeck, J. T., P. J. Bartlein and T. Webb III. 1991. Potential magnitude of future vegetation change in eastern North America: Comparisons with the past. Science 254:692–695. Overpeck, J. T., D. Rind, and R. Goldberg. 1990. Climate-induced changes in forest disturbance and vegetation. Nature 343:51–53. Overpeck, J. T., T. Webb III and I. C. Prentice. 1985. Quantitative interpretation of fossil pollen spectra: Dissimilarity coefficients and the method of modern analogs. Quaternary Research 23:87–108. Packard, F. M. 1942. Wildlife and aspen in Rocky Mountain National Park, Colorado. Ecology 23:478–482. Padgett, W. et al. 2017. Development of historical ecology concepts and their application to resource management and conservation. Pp. 19–28 in Historical Environmental Variation in Conservation and Natural Resource Management, J. A. Wiens et al., eds., John Wiley and Sons, New York. Palmer, T. 1992. The case for human beings. Atlantic Monthly 269(1):83–88. Parker, D. R., G. J. Leopold and J. K. Eichenberger. 1985. Tree dynamics in an oldgrowth, deciduous forest. Forest Ecology and Management 11:31–57. Parshall, T. and D. R. Foster. 2002. Fire on the New England landscape: Regional and temporal variation, cultural and environmental controls. Journal of Biogeography 29:1307–1313.

282  References Parsons, D. J. 1994. Objects or ecosystems? Giant sequoia management in national parks. USDA Forest Service General Technical Report PSW-151. Parsons, D. J., T. W. Swetnam and N. L. Christensen. 1999. Uses and limitations of historical variability concepts in managing ecosystems. Ecological Applications 9:1177– 1178. Parsons, D. J. et al. 1986. Natural fire management in National Parks. Environmental Management 10:21–24. Patterson, W. P., III and K. E. Sassaman. 1988. Indian fires in the prehistory of New England. Pp. 107–135 in Holocene Human Ecology in Northeastern North America, G. P. Nicholas, ed., Plenum, New York. Pausas, J. G. and M. P. Austin. 2001. Patterns of plant species richness in relation to different environments: An appraisal. Journal of Vegetation Science 12:153–166. Pearce, D. 1989. Sustainable futures: Some economic issues. Pp. 311–323 in Changing the Global Environment: Perspectives on Human Involvement, D. B. Botkin et al., eds., Academic Press, New York. Pederson, D. C. et al. 2005. Medieval warming, Little Ice Age, and European impact on the environment during the last millennium in the lower Hudson Valley, New York, USA. Quaternary Research 63:238–249. ———. 2015. Climate remains an important driver of post-European vegetation change in the eastern United States. Global Change Biology 21:2105–2110. Peet, R. K. 1975. Relative diversity indices. Ecology 56:496–498. Peet, R. K., D. C. Glenn-Lewin and J. W. Wolf. 1983. Prediction of man’s impact on plant species diversity: A challenge for vegetation science. Pp. 41–54 in Holzner, Werger and Ikusima 1983. Peglar, S. M. 1993. The mid-Holocene Ulmus decline at Diss Mere, Norfolk, UK: A yearby-year pollen stratigraphy from annual laminations. The Holocene 3:1–13. Pennington, W. 1970. Vegetation history in the north-west of England: A regional synthesis. Pp. 41–79 in Walker and West 1970. Perlin, J. 1989. A Forest Journey: The Role of Wood in the Development of Civilization. W. W. Norton, New York. Peteet, D. 1995. Global Younger Dryas. Quaternary International 28:93–104. Peterken, G. F. 1976. Long-term changes in the woodlands of Rockingham Forest and other areas. Journal of Ecology 64:123–146. ———. 1996. Natural Woodland: Ecology and Conservation in Northern Temperate Regions. Cambridge University Press, Cambridge. Peterson, A. T., M. Papes and M. Eaton. 2007. Transferability and model evaluation in ecological niche modeling: A comparison of GARP and Maxent. Ecography 30:550– 560. Phillips, S. J. et al. 2017. Opening the black box: An open source release of Maxent. Ecography 40:887–893. Piasecki, P. 1985. Environmental ambivalence: An analysis of implicit dangers. Pp. 83–98 in Bailes 1985a. Pickett, S. T. A. 1983. The absence of an Andropogon stage in old-field succession at the Hutcheson Memorial Forest. Bulletin of the Torrey Botanical Club 110:533–535. ———. 1989. Space-for-time substitution as an alternative to long-term studies. Pp. 110–135 in Likens 1989.

References  283 Pickett, S. T. A., M. L. Cadenasso and S. Bartha. 2001. Implications from the Buell-Small succession study for vegetation restoration. Applied Vegetation Science 4:41–52. Pickett, S. T. A. and T. S. White. 1985. Patch dynamics: A synthesis. Pp. 371–84 in The Ecology of Natural Disturbance and Patch Dynamics, S. T. A. Pickett and P. S. White, eds., Academic Press, Orlando. Pielou, E. C. 1974. The Interpretation of Ecological Data. Wiley and Sons, New York. ———. 1991. After the Ice Age. University of Chicago Press, Chicago. Pimm, S. L. and J. L. Gittleman. 1992. Biological diversity: Where is it? Science 255:940. Pinchot, G. 1900. The effects of fire. P. 108 in Annual Report of the State Geologist for 1899: Report on Forests, Geological Survey of New Jersey, Trenton. Pinelands Commission. 1980. Comprehensive Management Plan for the Pinelands National Reserve and Pinelands Area. Pinelands Commission, New Lisbon, NJ. Piperno, D. R. 2006. Phytoliths: A Comprehensive Guide for Archaeologists and Paleoecologists. AltaMira, Lanham, MD. Piussi, C. and S. Stiavelli. 1988. Forest history of the Cerbaie Hills (Toscona, Italy). Pp. 109–120 in Salbitano 1988. Plue, J. et al. 2008. Persistent changes in forest vegetation and seed bank 1,600 years after human occupation. Landscape Ecology 23:673–688. Polunin, O. and B. E. Smythies. 1973. Flowers of South-west Europe: A Field Guide. Oxford University Press, New York. Polyak, V. J. et al. 2001. Wetter and cooler late Holocene climate in the southwestern United States from mites preserved in stalagmites. Geology 29:643–646. Poole, R. 2008. Earthrise: How Man First Saw the Earth. Yale University Press, New Haven. Poschlod, P. and M. F. Wallis-DeVries. 2002. The historical and socioeconomic perspective of calcareous grasslands—Lessons from the distant and recent past. Biological Conservation 104:361–376. Poschlod, P. et al. 2008. The history of dry calcareous grasslands near Kallmünz (Bavaria) reconstructed by the application of palaeoecological, historical and recent-­ ecological methods. Pp. 130–143 in Human Nature: Studies in Historical Ecology and Environmental History, P. Szabó and R. Hédl, eds., Institute of Botany of the Czech Academy of Sciences, Brno, CZ. Poulson, B. 2016. Human archives: historians’ methodologies and past marine resource use. Pp. 71–87 in Perspectives on Oceans Past: A Handbook of Marine Environmental History, K. S. Mánˇez and B. Poulson, eds., Springer, Dordrecht. Power, E. 1941. The Wool Trade in English Medieval History. Oxford University Press, London. Power, M. J. et al. 2008. Changes in fire regimes since the Last Glacial Maximum: An assessment based on a global synthesis and analysis of charcoal data. Climate Dynamics 30:887–907. ———. 2018. Human fire legacies on ecological landscapes. Frontiers in Earth Science 6:151. Preston, C. D., D. A. Pearman and A. R. Hall. 2004. Archaeophytes in Britain. Botanical Journal of the Linnean Society 145:257–294. Prowse, T. A. A. et al. 2014. An ecological regime shift resulting from disrupted predator-prey interactions in Holocene Australia. Ecology 95:693–702. Pyne, S. J. 1991. Burning Bush: A Fire History of Australia. Holt, New York.

284  References ———. 2001. Fire: A Brief History. University of Washington Press, Seattle. ———. 2015. Between Two Fires: A Fire History of Contemporary America. University of Arizona, Tucson. Rackham, O. 1980. Ancient Woodland: Its History, Vegetation and Uses in E ­ ngland. Edward Arnold, London. ———. 1986. The History of the Countryside. J. M. Dent and Sons, London. ———. 1988. Trees and woodland in a crowded landscape—The cultural landscape of the British Isles. Pp. 53–77 in Birks et al. 1988. Radeloff, V. C. et al. 1999. Forest landscape change in the northwestern Wisconsin pine barrens from pre-European settlement to the present. Canadian Journal of Forest Research 29:1649–1659. Ramey-Gassert, L. T. and J. R. Runkle. 1992. Effect of land use practices on composition of woodlot vegetation in Greene County, Ohio. Ohio Journal of Science 91:25–32. Ranney, J. W., M. C. Bruner and J. B. Levenson. 1981. The importance of edge in the structure and dynamics of forest islands. Pp. 67–96 in Burgess and Sharpe 1981a. Rapaport, E. H. 1993. The process of plant colonization in small settlements and large cities. Pp. 190–207 in McDonnell and Pickett 1993. Ratcliffe, P. R. 1989. The control of red and Sika deer populations in commercial forests. Pp. 98–115 in Mammals as Pests, R. J. Putnam, ed., Chapman and Hall, London. Rautio, A. M., T. Josefsson and L. Östlund. 2014. Sami resource utilization and site selection: Historical harvesting of inner bark in northern Sweden. Human Ecology 42:137–146. Rautio, A. M. et al. 2015. People and pines 1555–1910: Integrating ecology, history and archaeology to assess long-term resource use in northern Fennoscandia. Landscape Ecology 31:337–349. Raven, P. H. 1994. Defining biodiversity. Nature Conservancy 44(1):11–15. Redmon, C. L. 1999. Human Impact on Ancient Environments. University of Ari­zona, Tucson. Reimer, P. H. et al. 2013. IntCal13 and Marine13 radiocarbon age calibration curves 0–50,000 years cal BP. Radiocarbon 55:1869–1887. Reitalu, T., P. Kuneš and T. Giesecke. 2014. Closing the gap between plant ecology and Quaternary palaeoecology. Journal of Vegetation Science 25:1188–1194. Ren, G. 2007. Changes in forest cover in China during the Holocene. Vegetation History and Archaeobotany 16:119–126. Renfrew, C. 1979. Before Civilization: The Radiocarbon Revolution and Prehistoric Europe. Cambridge University Press, Cambridge. Retallack, G. J., D. P. Dugas, and E. A. Bestland. 1990. Fossil soils and grasses of a middle Miocene East African grassland. Science 247:1325–1328. Reynolds, S. G. 1998. Kazak herders, winter feed and transhumant systems in Altai Prefecture, Xinjiang, China. In Sustainable Himalayan Pasture and Fodder Production Systems: Present Problems and Prospects. Proceedings of the Third Meeting of the Temperate Asia Pasture and Fodder Working Group 9:82–87. Rhemtulla, J. M. and D. J. Mladenoff. 2007. Why history matters in landscape ecology. Landscape Ecology 22 supplement:1-3. Rhemtulla, J. M., D. J. Mladenoff and M. K. Clayton. 2007. Regional land-cover con-

References  285 version in the U.S. upper Midwest: Magnitude of change and limited recovery (1850– 1935–1993). Landscape Ecology 22:57–75. Rhoades, C. C., S. P. Miller and M. M. Shea. 2004. Soil properties and soil nitrogen dynamics of prairie-like forest openings and surrounding forests in Kentucky’s Knobs region. American Midland Naturalist 152:1–11. Rhode, D. 2005. Packrat middens as a tool for reconstructing historic ecosystems. Pp. 257–293 in Egan and Howell 2005. Ricciardi, A. 2004. Assessing species invasions as a cause of extinction. Trends in Ecology and Evolution 19:619. Richard, A. F. and R. E. Dewar. 1991. Lemur ecology. Annual Review of Ecology and Systematics 22:145–175. Richards, J. F. and R. P. Tucker. 1988. World Deforestation in the Twentieth Century. Duke University Press, Durham, NC. Richards, P. W. 1973. Africa, “the odd man out.” P. 24 in Tropical Forest Ecosystems in Africa and South America, B. J. Meggers, E. S. Ayensu, and W. D. Duckworth, eds., Smithsonian Institution Press, Washington, DC. Rick, T. C. et al. 2015. Archaeology, taphonomy, and historical ecology of Chesapeake Bay blue crabs (Callinectes sapidus). Journal of Archaeological Science 55:42–54. Robbins, R. 1976. Our Landed Heritage: The Public Domain, 1776–1970. 2nd ed. University of Nebraska, Lincoln. Roberts, M. L. and R. L. Stuckey. 1992. Distribution patterns of selected aquatic and wetland vascular plants in relation to the Ohio canal system. Bartonia 57:50–74. Roberts, N. 2013. The Holocene: An Environmental History. Wiley-Blackwell, Chichester, UK. Robin, V. O. N. et al. 2018. A comparative review of soil charcoal data: Spatiotemporal patterns of origin and long-term dynamics of Western European nutrient-poor grasslands. The Holocene 28:1313–1324. Robinson, G. S., L. P. Burney and D. A. Burney. 2005. Landscape paleoecology and megafaunal extinction in southeastern New York state. Ecological Monographs 75:295–315. Robinson, M. A. 2000. Coleopteran evidence for the elm decline, Neolithic activity in woodland, clearance and the use of the landscape. Pp. 27–36 in Plants in Neolithic Britain and Beyond, A. S. Fairborn, ed., Oxbow Books, Oxford. Rogers, N. 2016. Museum drawers go digital: New technology speeds efforts to display billions of natural history specimens online. Science 352:762–765. Rogers, S. 1823. 21 February Estate Inventory, Saratoga County (NY) Clerk’s office, Ballston Spa. Rollefson, G. O. 1990. The uses of plaster at Neolithic Ain Ghazal, Jordan. Archaeomaterials 4:33–54. Roosevelt, A. C. 1980. Parmana: Prehistoric Maize and Manioc Subsistence along the Amazon and Orinoco. Academic Press, New York. Rosania, C. N. et al. 2008. Revisiting Carpathian obsidian. Antiquity 82:318. Rotherham, I. E. 2007. The History, Ecology and Archaeology of Medieval Parks and Parklands. Wildtrack, Sheffield, UK.

286  References Rothrock, J. T. 1894. Forests of Pennsylvania. Proceedings of the American Philosophical Society 33:114–133. Ruddiman, W. F., J. E. Kutzbach and S. J. Vavrus. 2011. Can natural or anthropogenic explanations of late-Holocene CO2 and CH4 increases be falsified? The Holocene 21:1–15. Ruddiman, W. F. et al. 2015. Defining the epoch we live in: Is a formally designated “Anthropocene” a good idea? Science 348:38–39. Rull, V. 2015. Long-term vegetation stability and the concept of potential natural vegetation in the Neotropics. Journal of Vegetation Science 26:603–607. Russell, E. W. B. 1979. Vegetational change in northern New Jersey since 1500 A.D.: A palynological, vegetational and historical synthesis. PhD diss., Rutgers University. ———. 1983. Indian-set fires in the forests of the northeastern United States. Ecology 64:78–88. ———. 1988a. The 1899 New Jersey state geologist’s report: A call for forest management. Journal of Forest History 32:205–211. ———. 1988b. Vegetation study of Hopewell Furnace National Historic Site. Report to Hopewell Furnace NHS. ———. 1993a. Cultural landscape history, Saratoga National Historical Park. Report to National Park Service, North Atlantic Regional Office. ———. 1993b. Early stages of secondary succession recorded in soil pollen on the North Carolina piedmont. American Midland Naturalist 129:384–396. Russell, E. W. B. and R. B. Davis. 2001. Five centuries of changing forest vegetation in the northeastern United States. Plant Ecology 155:1–13. Russell, E. W. B. and A. E. Schuyler. 1988. Vegetation and flora of Hopewell Furnace National Historic Site, eastern Pennsylvania. Bartonia 54:124–143. Russell, E. W. B. et al. 1993. Recent centuries of vegetational change in the glaciated north-eastern United States. Journal of Ecology 81:647–664. Russell, F. H. 1995. The bifurcation of creation: Augustine’s attitudes toward nature. Pp. 83–96 in Man and Nature in the Middle Ages, S. J. Ridyard and R. G. Benson, eds., University of the South Press, Sewanee, TN. Ryan, P.A. and J. J. Blackford. 2010. Late Mesolithic environmental change at Black Heath, south Pennines, UK: A test of Mesolithic woodland management models using pollen, charcoal and non-pollen palynomorph data. Vegetation History and Archaeo­ botany 19:545–558. Saarnisto, M. 1986. Annually laminated lake sediments. Pp. 342–370 in Berglund 1986. Sagoff, M. 1993. Biodiversity and the culture of ecology. Bulletin of the Ecological Society of America 74: 374–381. Salbitano, F. 1988. Human Influence on Forest Ecosystems Development in Europe. Pitagora Editrice, Bologna. Sanderson, E. W. 2009. Mannahatta: A Natural History of New York. Abrams, New York. Sanford, R. L. et al., 1985. Amazon rain-forest fires. Science 227:53–55. Sanitary and Topographical Map of Hudson County, NJ. 1881. Hanging Map, Special Collections, Rutgers University, New Brunswick, NJ.

References  287 Santmire, H. P. 1973. Historical dimensions of the American crisis. Pp. 66–92 in Western Man and Environmental Ethics, I. Varbour, ed., Addison, Wesley, Reading, MA. Sarvis, W. 1993. An Appalachian forest: Creation of the Jefferson National Forest and its effect on the local community. Journal of Forest and Conservation History 37:169–178. Sauer, C. O. 1941. Forward to historical geography. Annals of the Association of American Geographers 31:1–24. ———. 1952. Agricultural Origins and Dispersals. American Geographical Society, New York. ———. 1966. The Early Spanish Main. University of California Press, Berkeley. ———. 1969. Seeds, Hearths and Herds. MIT Press, Cambridge. Scarry, C. M. 1993. Agricultural risk and the development of the Moundville chiefdom. Pp. 157–181 in Foraging and Farming in the Eastern Woodlands, University of Florida, Gainesville, ebook Collection EBSCO. Schabel, H. G. 1990. Tanganyika forestry under German colonial administration, 1891– 1991. Journal of Forest and Conservation History 34:130–141. Scheel-Ybert, R. et al. 2003. Holocene palaeoenvironmental evolution in the São Paulo State (Brazil), based on anthracology and soil δ13C analysis. The Holocene 13:73–81. Scheller, R. M. and D. J. Mladenoff. 2005. A spatially interactive simulation of climate change, harvesting, wind, and tree species migration and projected changes to forest composition and biomass in northern Wisconsin, USA. Global Change Biology 11:307–321. Scheller, R. M. et al. 2008. Simulation of forest change in the New Jersey Pine Barrens under current and pre-colonial conditions. Forest Ecology and Management 255:1489–1500. Schlesinger, W. H. et al. 1990. Biological feedbacks in global desertification. Science 247:1043–1048. Schluter, D. and R. E. Ricklefs. 1993. Species diversity: An introduction to the problem. Pp. 1–10 in Species Diversity in Ecological Communities: Historical and Geographical Perspectives, R. E. Ricklefs and D. Schluter, eds., University of Chicago Press, Chicago. Schmid, B. V. et al. 2015. Climate-driven introduction of the Black Death and successive plague reintroductions into Europe. Proceedings of the National Academy of Science 112:3020–3025. Schmidt, H. G. 1946. Rural Hunterdon: An Agricultural History. Rutgers University Press, New Brunswick, NJ. ———. 1973. Agriculture in New Jersey: A Three-Hundred Year History. Rutgers University Press, New Brunswick, NJ. Schuster, W. S. F. 2011. Forest ecology. Pp. 132–157 in Highlands: Critical Resources, Treasured Landscapes, R. G. Lathrop, ed., Rutgers University Press, New Brunswick, NJ. Schuyler, A. E. and J. L. Stasz. 1985. Influence of fire on reproduction of Scirpus longii. Bartonia 51:105–107. Schwarz, G. F. 1907. The sprout forests of the Housatonic Valley of Connecticut. Forest Quarterly 5:121–153.

288  References Scott, J. C. 2017. Against the Grain: A Deep History of the Earliest States. Yale University Press, New Haven. Seddon, A. W. R. et al. 2014. Looking forward through the past: Identification of 50 priority research questions in palaeoecology. Journal of Ecology 102:256–267. Seersholm, F. V. et al. 2018. Subsistence practices, past biodiversity, and anthropogenic impacts revealed by New Zealand–wide ancient DNA survey. Proceedings of the National Academy of Sciences 115:7771–7776. Severinghouse, C. W. and C. P. Brown. 1956. History of the white-tailed deer in New York. New York Fish and Game Journal 3:130–167. Seymour, E. L. D. 1970. The Wise Garden Encyclopedia. Grosset and Dunlap, New York. Shaler, N. S. 1884. Kentucky: A Pioneer Commonwealth. Houghton Mifflin, New York. Shapiro, J. and E. B. Swain. 1983. Lessons from the silica “decline” in Lake Michigan. Science 221:457–459. Sharpe, D. M. et al. 1981. Spatio-temporal patterns of forest ecosystems in man-­ dominated landscapes. Pp. 109–116 in Proceedings International Congress of the Netherlands Society of Landscape Ecology, Veldhoven, Netherlands. Shaw, M. W. 1974. The reproductive characteristics of oak. Pp. 162–181 in Morris and Perrin 1974. Sheail, J. 1984. The rabbit. Biologist 31:135–140. Shearman, R. 1990. The meaning and ethics of sustainability. Environmental Management 14:1–8. Shreve, F. et al. 1910. The Plant Life of Maryland. Maryland Weather Service Special Publication 8. Johns Hopkins University Press, Baltimore. Shugart, H. H. and D. C. West. 1977. Development of an Appalachian deciduous forest model and its application to the assessment of the chestnut blight. Journal of Environmental Management 5:161–179. ———. 1980. Forest succession models. BioScience 30:308–313. Signell, S. A. and M. D. Abrams. 2008. Interactions between landscape features, disturbance and vegetation in frequently burned Appalachian oak forests. Advances in Environmental Research 1:145–164. Simberloff, D. 2013. Invasive Species: What Everyone Needs to Know. Oxford University Press, Oxford. Simmons, I. G. 1989. Changing the Face of the Earth: Culture, Environment, History. Blackwell, New York. Simonetti, J. A. 1988. Human disturbance and community patterns in central Chile. Bulletin of the Ecological Society of America 69:296. Sinclair, A. R. E. 2012. Ecological history guides the future of conservation: Lessons from Africa. Pp. 266–271 in Historical Environmental Variation in Conservation and Natural Resource Management, J. A. Wiens et al., eds., John Wiley and Sons, New York. Singrayer, J. S. et al. 2011. Late Holocene methane rise caused by orbitally controlled increase in tropical sources. Nature 470:82–85. Skole, D. and C. Tucker. 1993. Tropical deforestation and habitat fragmentation in the Amazon: Satellite data from 1978–1988. Science 260:905–910. Smart, P. L. and P. D. Frances, eds. 1991. Quaternary Dating Methods—A User’s Guide.

References  289 Quaternary Research Association, Technical Guide No. 4. Quaternary Research Association, Cambridge. Smith, A. G. 1970. The influence of Mesolithic and Neolithic man on British vegetation: A discussion. Pp. 81–96 in Walker and West 1970. Smith, B. D. 1989. Origins of agriculture in eastern North America. Science 246:1566– 1571. Smith, J. B. 1900. The role of insects in the forest. Pp. 229–232 in Annual Report of the State Geologist for 1899: Report on Forests, Geological Survey of New Jersey, Trenton. Smith, N. J. 1980. Anthrosols and human carrying capacity in Amazonia. Annals of the Association of American Geographers 70:553–566. Smol, J. P. and E. F. Stoermer, eds. 2010. The Diatoms: Applications for the Environmental and Earth Sciences. 2nd ed. Cambridge University Press, Cambridge. Smol, J. P. et al. 1986. Diatoms and Lake Acidity: Reconstructing pH from Siliceous Algal Remains in Lake Sediments. Springer-Verlag, New York. Snir, A., et al. 2015. The origin of cultivation and proto-weeds, long before Neolithic farming. PLoS One 10(7):e0131422. Snow, D. R. 1980. The Archaeology of New England. Academic Press, New York. Snyder, D. B. 1993. Extinct, extant, extirpated or historical? or, In defense of historical species Bartonia 57 (suppl.):50–57. Soberón, J. and A. T. Peterson. 2005. Interpretation of models of fundamental ecological niches and species distributional areas. Biodiversity Informatics 2:1–10. Solomon, A. M. and D. F. Kroener. 1971. Suburban replacement of rural land uses reflected in the pollen rain of northeastern New Jersey. Bulletin of the New Jersey Academy of Sciences 16:30–44. Somerville, M. 1853. Physical Geography. 3rd ed. Blanchard and Lea, Philadelphia. Soulé, M. 1991. Conservation: Tactics for a constant crisis. Science 253:744–750. Southgate, E. W. B. 2006. Historical ecology of American chestnut (Castanea dentata). Pp. 13–19 in Restoration of American Chestnut to Forest Lands, K. C. Steiner and J. E. Carlson, eds., Natural Resources Report NPS/NCR/CUE/NRR-2006/01. Southgate, E. W. B. (Russell) and J. E. Thompson. 2014. Secondary forest succession in a post-agricultural landscape in the Hudson Valley, New York. Northeastern Naturalist 21:35–50. Sperry, L. J., J. Belnap and R. D. Evans. 2006. Bromus tectorum invasion alters nitrogen dynamics in an undisturbed arid grassland ecosystem. Ecology 87:603–615. Spies, T. A., W. J. Ripple and G. A. Bradshaw. 1994. Dynamics and pattern of a managed coniferous forest landscape in Oregon. Ecological Applications 4:555–568. Sponsel, L. 1992. The environmental history of Amazonia: Natural and human disturbance, and the ecological tradition. Pp. 233–251 in Changing Tropical Forests: Historical Perspectives on Today’s Challenges in Central and South America, H. K. Steen and R. P. Tucker, eds., Forest History Society, n.p. Stanley, D. J. and A. G. Warne. 1993. Nile delta: Recent geological evolution and human impact. Science 260:628–634. Stanley, S. M. and X. Xang. 1994. A double mass extinction at the end of the Paleozoic Era. Science 266:1340–1344.

290  References Stefferud, A. ed. 1949. Trees: The Yearbook of Agriculture. USDA, Government Printing Office, Washington, DC. Steinberg, C. E. W. and R. F. Wright. 1994. Acidification of Freshwater Ecosystems: Implications for the Future. John Wiley and Sons, Chichester, UK. Stewart, G. H. et al. 2009. Urban biotopes of Aoteoroa New Zealand (URBANZ) II: Composition and diversity of temperate urban lawns in Christchurch. Urban Ecosystems 12:333–348. Stewart, O. C. 1956. Fire as the first great force employed by man. Pp. 115–133 in Thomas 1956. Stohlgren, T. J., G. T. Barnett and J. T. Kartesz. 2003. The rich get richer: Patterns of plant invasions in the United States. Frontiers in Ecology and the Environment 1:11–14. Streeter, D. T. 1974. Ecological aspects of oak woodland conservation. Pp. 341–353 in Morris and Perring 1974. Subcommission on Quaternary Stratigraphy. 2018. http://quaternary.stratigraphy.org/ major-divisions/. Sugita, S. 2007a. Theory of quantitative reconstruction of vegetation I: Pollen from large sites reveals regional vegetation composition. The Holocene 17:229–241. ———. 2007b. Theory of quantitative reconstruction of vegetation II: All you need is love. The Holocene 17:243–257. Sutcliffe, A. J. 1985. On the Track of Ice Age Mammals. Harvard University Press, Cambridge. Swan, J. D. and J. M. A. Swan. 1974. Reconstructing forest history from live and dead plant materials—An approach to the study of forest succession in southwest New Hampshire. Ecology 55:772–783. Swetnam, C. W., C. D. Allen and J. L. Betancourt. 1999. Applied historical ecology: Using the past to manage for the future. Ecological Applications 9:1189–1206. Szabó, P. 2010. Why history matters in ecology: An interdisciplinary perspective. Environmental Conservation 37:380–387. ———. 2013. Rethinking pannage: Historical interactions between oak and swine. Pp. 68–78 in Trees, Forested Landscapes and Grazing Animals, I. D. Rotherham, ed., Routledge, London. ———. 2015. Historical ecology: Past, present and future. Biological Reviews 90:997– 1014. Szabó, P. et al. 2016. Using historical ecology to reassess the conservation status of coniferous forests in central Europe. Conservation Biology 31:150–160. Tallis, H. et al. 2008. An ecosystem services framework to support both practical conservation and economic development. Proceedings of the National Academy of Sciences 105: 9457–9464. Tansley, A. G. 1923. Practical Plant Ecology. Dodd, Mead, New York. Taylor, A. R. 1974. Ecological aspects of lightning in forests. Tall Timbers Fire Ecology Conference Proceedings 13:455–482. Taylor, L. R. 1989. Objective and experiment in long-term research. Pp. 20–70 in Likens 1989. Taylor, R. H. 1953. Fertilizers and farming in the Southeast, 1840–1900. North Carolina Historical Review 30:305–328.

References  291 Terborgh, J., 1989. Where Have All the Birds Gone? Essays on the Biology and Conservation of Birds That Migrate to the American Tropics. Princeton University Press, Princeton. ter Braak, C. J. and H. van Dame. 1989. Inferring pH from diatoms: A comparison of old and new calibration methods. Hydrobiologia 178:209–223. Thomas, C. D. G. and G. Palmer. 2015. Non-native plants add to the British flora without negative consequences for native diversity. Proceedings of the National Academy of Sciences 112:4387–4392. Thomas, W. L., Jr., ed. 1956. Man’s Role in Changing the Face of the Earth. University of Chicago Press, Chicago. Thompson, F. M. L. 1968. The second agricultural revolution, 1815–1880. Economic History Review: 21:62–77. Thompson, K. et al. 2004. Urban domestic gardens (III): Composition and diversity of lawn floras. Journal of Vegetation Science 15:373–378. Thompson, K. M. et al. 2015. Connecting contemporary ecology and ethnobotany to ancient plant use practices of the Maya at Tikal. Pp. 124–151 in Tikal: Paleoecology of an Ancient Maya City, D. L. Lentz, N. P. Dunning, and V. L. Scarborough, eds., Cambridge University Press, Cambridge. Thuiller, W. et al. 2005. Climate change threats to plant diversity in Europe. Proceedings of the National Academy of Science 102(23):8245–8250. Tilman, D. and J. A. Downing. 1994. Biodiversity and stability in grasslands. Nature 367:363–365. Tilman, D. and S. Pacala. 1993. The maintenance of species richness in plant communities. Pp. 13–25 in Species Diversity in Ecological Communities: Historical and Geographical Perspectives, R. E. Ricklefs and D. Schluter, eds., University of Chicago Press, Chicago. Tilman, D. et al. 1994. Habitat destruction and the extinction debt. Nature 371:65–66. Townsend, C. R. 2003. Individual, population, community, and ecosystem consequences of a fish invader in New Zealand streams. Conservation Biology 17:38–47. Townsend, J. F. 2016. Natural heritage resources of Virginia: Rare plants. Natural Heritage Technical Report 16–18. Virginia Department of Conservation and Recreation, Division of Natural Heritage, Richmond, VA. Unpublished report. Trimble, S. W. 1974. Man-induced soil erosion on the southern Piedmont, 1700–1970. PhD diss., University of Wisconsin. ———. 1992. The Alcovy River swamps: The result of culturally accelerated sedimentation. Pp. 21–32 in The American Environment: Interpretations of Past Geographies, L. M. Dilsaver and C. E. Colten, eds., Rowman and Littlefield, Lanham, MD. (Originally published in 1970 in Bulletin of the Georgia Academy of Sciences 28:131–141.) ———. 2012. Historical sources and watershed evolution. Philosophical Transactions of the Royal Society A 370:205–292. Tuan, Y. 1968. Discrepancies between environmental attitudes and behavior: Examples from Europe and China. Canadian Geographer 12:176–191. Tulowiecki, S. J. 2018. Information retrieval in physical geography: A method to recover geographical information from digitized historical documents. Progress in Physical Geography: Earth and Environment 42:369–390.

292  References Tunison, J. T., C. M. D’Antonio and R. K. Loh. 2001. Fire and invasive plants in Hawai‘i Volcanoes National Park. Pp.122–131 in Proceedings of the Invasive Species Workshop: The Role of Fire in the Control and Spread of Invasive Species, K. E. M. Galley and T. P. Wilson, eds., Miscellaneous Publication No. 11, Tall Timbers Research Station, Tallahassee, FL. Turner, B. L. et al., eds. 1990. The Earth as Transformed by Human Action: Global and Regional Changes in the Biosphere over the Past 300 Years. Cambridge University Press, Cambridge. Turner, F. T. et al. 2015. Retrospective stable isotope analysis reveals ecosystem responses to river regulation over the last century. Ecology, 96:3213–3226. Turner, J. 1970. Post-Neolithic disturbance of British vegetation. Pp. 97–116 in Walker and West 1970. Tushingham, S. and R. L. Bettinger. 2013. Why foragers choose acorns before salmon: Storage, mobility and risk in aboriginal California. Journal of Anthropological Archaeology 32:527–537. Ullman, E. L. 1965. The role of transportation and the bases for interaction. Pp. 862– 880 in Thomas 1956. Underwood, H. B. et al. 1989. Deer and vegetation interactions on Saratoga National Historical Park. Draft report, National Park Service, North Atlantic Regional Office. United States Federal Census, Agricultural Schedules, 1850–1960. Washington, DC. United States Geological Survey. 2004. https://pubs.usgs.gov/fs/fs074-97/. Utterström, G. 1988. Climatic fluctuations and population problems in early modern history. Scandinavian History Review 3:3–47. (Reprinted pp. 39–79 in Worster, D., ed. 1988. The Ends of the Earth. Cambridge University Press, Cambridge.) van der Donck, A. 1650. Remonstrance of the deputies from New Netherland. Pp. 275–318 in Documents Relative to the Colonial History of the State of New York, E. B. O’Callaghan, ed., Weed Parsons, Albany. ———. 1841. A description of the New Netherlands. Pp. 125–242 in Collections of the New York Historical Society, 2nd ser., vol. 1. (Reprinted in O’Donnell, T. F., ed. 1968. A Description of the New Netherlands. Syracuse University Press, Syracuse.) ———. 1854. Vertoogh van Nieu Nederland. Henry C. Murphy, trans. N.p., New York. VanDyck, H. et al. 2009. Decline in common, widespread butterflies in a landscape under intense human use. Conservation Biology 23:957–965. Vane-Wright, R. I. 1993. The Columbus hypothesis: An explanation for the dramatic 19th century range expansion of the monarch butterfly. Pp. 179–187 in Biology and Conservation of the Monarch Butterfly, S. B. Malcolm and M. P. Zalucki, eds., Natural History Museum of Los Angeles County, Los Angeles. Van Houtan, K. S., L. McClenachan and J. N. Kittinger. 2013. Seafood menus reflect long-term ocean changes. Frontiers in Ecology and the Environment 11:289–290. Vankat, J. L. 1979. The Natural Vegetation of North America. John Wiley and Sons, New York. Van Zant, K. L. et al. 1979. Increased Cannabis/Humulus pollen, an indicator of European agriculture in Iowa. Palynology 3:227–233. Veblen, T. T. 2017. Disturbance in biogeography. Pp. 1–13 in The International En-

References  293 cyclopedia of Geography, D. Richardson et al., eds., John Wiley and Sons. doi: 10.1002/9781118786352.wbieg0884. Vellend, M. 2017. The biological diversity conservation paradox. American Scientist 105: 94–101. Vellend, M. et al. 2007. Homogenization of forest plant communities and weakening of species-environment relationships via agricultural land use. Journal of Ecology 95: 565–573. ———. 2013. Global meta-analysis reveals no net change in local-scale plant biodiversity over time. Proceedings of the National Academy of Science 110(48):19456–19459. Verheyen, K. et al. 2003. Response of forest plant species to land-use change: A life history trait-based approach. Journal of Ecology 91:563–577. ———. 2018. Observer and relocation errors matter in resurveys of historical vegetation plots. Journal of Vegetation Science 29:812–823. doi.org/10.1111/jvs.12673. Verhulst, A. 1990. The “agricultural revolution” of the Middle Ages reconsidered. Pp. 17–28 in Law, Custom and the Social Fabric: Essays in Honor of Bryce Lyon, B. S. Bachrach and D. Nicholas, eds., Ghent University Studies in Medieval Culture 28, Kalamazoo, MI. Vermeule, C. C. 1900. The forests of New Jersey. Pp. 13–174 in Annual Report of the State Geologist for the Year 1899: Report on Forests, Geological Survey of New Jersey, Trenton. Vickery, P. D. and J. R. Herkert. 1999. Ecology and Conservation of Grassland Birds of the Western Hemisphere. Cooper Ornithological Society, Camarillo. Vignieri, S. 2014. Vanishing fauna. Science 345:392–395. Vitousek, P. 1987. Introduced species in Hawai’i: Biological effects as opportunities for ecological research. Trends in Ecology and Evolution 1:114–117. Vitousek, P., C. M. D’Antonio and G. P. Asner. 2011. Invasions and ecosystems: Vulnerabilities and the contributions of new technologies. Pp. 277–288 in Fifty Years of Invasion Ecology: The Legacy of Charles Elton, D. M. Livingston, ed., Blackwell, Oxford. Von Holle, B., Y. Wei, and D. Nickerson. 2010. Climatic variability leads to later seasonal flowering of Floridian plants. PLoS One 5:e11500. von Wangenheim, F. A. J. 1781. Beschreibung einiger Nordamerikanischen Holz und Busharten, mit anwendung auf Deutsche Forsten. J.C. Dieterich, Göttingen. Voosen, P. 2018. New geological age comes under fire. Science 361:537–538. Vorontsova, M. S. et al. 2016. Madagascar’s grasses and grasslands: Anthropogenic or natural? Proceedings of the Royal Society B 283(1823):20152262. Wacker, P. O. 1975. Land and People: A Cultural Geography of Preindustrial New Jersey; Origins and Settlement Patterns. Rutgers University Press, New Brunswick, NJ. ———. 1979. Human exploitation of the New Jersey Pine Barrens before 1900. Pp. 3–23 in Pine Barrens: Ecosystem and Landscape, R. T. T. Forman, ed., Academic Press, New York. Wagner, F. H. and C. E. Kay. 1993. “Natural” or “healthy” ecosystems: Are the U.S. National Parks providing them? Pp. 257–276 in McDonnell and Pickett 1993. Walker, D. and Y. Chen. 1987. Palynological light on tropical rainforest dynamics. Quaternary Science Reviews 6:177–192.

294  References Walker, D. and R. G. West, eds. 1970. Studies in the Vegetational History of the British Isles. Cambridge University Press, Cambridge. Wallin, D. O., F. J. Swanson and B. Marks. 1994. Landscape pattern response to changes in pattern generation rules: Land-use legacies in forestry. Ecological Applications 4:569–580. Walsh, W. A. 1967. Philosophy of History: An Introduction. Rev. ed. Harper Torchbooks, New York. Walter, R. C. and D. J. Merritts. 2008. Natural streams and the legacy of water-powered mills. Science 319:299–304. Wang, G. G. and H. Hu. 2015. The replacement of American chestnut: A range-wide assessment based on forest inventory and published studies. Pp. 513–515 in Proceedings of the 17th Biennial Southern Silvicultural Research Conference, A. G. Holley, K. F. Connor, and J. D. Haywood, eds., e–General Technical Report SRS-203, U.S. Department of Agriculture, Forest Service, Southern Research Station, Asheville, NC. Wardle, D. A. et al. 2016. Terrestrial ecosystem responses to species gains and losses. Science 332:1273–1277. Wares, J. P. and C. W. Cunningham. 2001. Phylogeography and historical ecology of the North Atlantic intertidal. Evolution 55:2455–2469. Waters, C. N. et al. 2016. The Anthropocene is functionally and stratigraphically distinct from the Holocene. Science 351:137, science.aad2622. Watkins, C. 2014. Trees, Woods and Forests: A Social and Cultural History. Reaktion Books, London. Watkins, C. and K. J. Kirby. 1998. Introduction—Historical ecology in European woodlands. Pp. ix–xv in The Ecological History of European Forests, K. J. Kirby and C. Watkins, eds., CAB International, Wallingford, UK. Watts, W. A. 1967. Late-glacial plant macrofossils from Minnesota. Pp. 89–97 in Quaternary Paleoecology, E. J. Cushing and H. E. Wright Jr., eds., Yale University Press, New Haven. ———. 1979. Late Quaternary vegetation of central Appalachia and the New Jersey coastal plain. Ecological Monographs 49:427–469. Waugh, G. D. 1973. Fish and fisheries. In The Natural History of New Zealand: An Ecological Survey, G. R. Williams, ed., A. H. and A. Reed, Wellington, NZ. Weakley, A. S., J. C. Ludwig and J. F. Townsend. 2012. Flora of Virginia. BRIT Press, Botanical Research Institute of Texas, Fort Worth. Webb, E. A. et al. 2007. Stable carbon isotope signature of ancient maize agriculture in the soils of Motul de San José, Guatemala. Geoarchaeology: An International Journal 22:291–312. Webb, S. 1986. Potential role of passenger pigeons and other vertebrates in the rapid Holocene migrations of nut trees. Quaternary Research 26:367–375. Webb, T. III. 1973. A comparison of modern and presettlement pollen from southern Michigan (USA). Review of Palaeobotany and Palynology 16:137–156. ———. 1974. Corresponding patterns of pollen and vegetation in lower Michigan: A comparison of quantitative data. Ecology 55:17–28. ———. 1986. Is vegetation in equilibrium with climate? How to interpret late-­ Quaternary pollen data. Vegetatio 67:75–91.

References  295 ———. 1987. The appearance and disappearance of major vegetational assemblages: Long-term vegetational dynamics in eastern North America. Vegetatio 69:177–187. ———. 1988. Eastern North America. Pp. 185–414 in Huntley and Webb 1988. ———. 1992. Past changes in vegetation and climate: Lessons for the future. Pp. 59–75 in Global Warming and Biological Diversity, R. L. Peters and T. E. Lovejoy, eds., Yale University Press, New Haven. West, N. E. 1988. Intermountain deserts, shrub steppes and woodlands. Pp. 209–230 in North American Terrestrial Vegetation, M. G. Barbour and W. D. Billings, eds., Cambridge University Press, Cambridge. Westveld, R. H. 1949. Applied Silviculture in the United States. 2nd ed. John Wiley and Sons, New York. Whitcomb, R. F. et al. 1981. Effects of forest fragmentation on avifauna of the eastern deciduous forest. Pp. 125–205 in Burgess and Sharpe 1981a. White, L., Jr. 1967. The historical roots of our ecologic crisis. Science 155:1203–1207. White, P. S. and R. D. White. 1996. Old-growth oak and oak-hickory forests. Pp. 178–198 in Eastern Old-Growth Forests: Prospects for Rediscovery and Recovery, J. Davis and M. D. Davis, eds., Island Press, Washington, DC. White, R. 1989. “Interpreting the historical environment.” Ecological Society of America Meeting, 1989, abstract in Bulletin of the Ecological Society of America 71(2):367. Whitehead, W. A. 1881. Answer to letter inquiring about former forest between Newark and New York. Proceedings New Jersey Historical Society, 2nd ser., 6:145–147. Whitlock, C. et al. 2010. Paleoecological perspectives on fire ecology: Revisiting the fire-regime concept. Open Ecology Journal 3:6–23. Whitney, G. G. 1994. From Coastal Wilderness to Fruited Plain: A History of Environmental Change in Temperate North America, 1500–Present. Cambridge University Press, Cambridge. Whitney, G. G. and W. C. Davis. 1986. From primitive woods to cultivated woodlots: Thoreau and the forest history of Concord, Massachusetts. Journal of Forest History 30:70–81. Whitney, G. G. and J. DeCant. 2005. Government Land Office Survey and other early land surveys. Pp. 147–172 in Egan and Howell 2005. Whitney, G. G. and D. R. Foster. 1988. Overstorey composition and age as determinants of the understorey flora of woods of central New England. Journal of Ecology 76:867–876. Whittaker, R. H. and W. A. Niering. 1965. Vegetation of the Santa Catalina Mountains, Arizona V. Biomass, production, and diversity along the elevation gradient. Ecology 56:771–790. Whittaker, R. J., K. J. Willis and R. Field. 2001. Scale and species richness: Towards a general, hierarchical theory of species diversity. Journal of Biogeography 28:453–470. Wiedner, K. et al. 2015. Anthropogenic dark earth in northern Germany—The Nordic analogue to terra preta de Índio in Amazonia. Catena 132:114–125. Wigston, D. L. 1993. Applications of historical ecology: Case studies from Europe and Australia. Bulletin of the Ecological Society of America 74(2):486. Wilkins, G. R. et al. 1991. Paleoecology of central Kentucky since the last glacial maximum. Quaternary Research 36:224–239.

296  References Wilkinson, W. C. 1777. The encampment & position of the army under His Excy. Lt. Gl. Burgoyne at Swords and Freeman’s Farms on Hudsons River near Stillwater. Copy at Saratoga National Historical Park, Schuylerville, NY. Willerslev, E. et al. 2003. Diverse plant and animal genetic records from Holocene and Pleistocene sediments. Science 300:791–795. Williams, J. W. et al. 2004. Late-Quaternary vegetation dynamics in North America: Scaling from taxa to biomes. Ecological Monographs 74:309–334. Williams, M. 1974. The Making of the South Australia Landscape. Academic Press, London. ———. 1989. Americans and Their Forests: A Historical Geography. Cambridge University Press, Cambridge. ———. 1994. The relations of environmental history and historical geography. Journal of Historical Geography 20:3–21. ———. 2000. Dark ages and dark areas: Global deforestation in the deep past. Journal of Historical Geography 26:28–46. Willis, J. C. 1922. Age and Area: A Study in Geographical Distribution and Origin of Species. Cambridge University Press, Cambridge. ———. 1973. A Dictionary of the Flowering Plants and Ferns. 8th ed. Rev. by H. K. Airy Shaw. Cambridge University Press, Cambridge. Willis, K. J., L. Gillson and T. M. Brncic. 2004. How” virgin” is virgin rainforest? Science 304:402–403. Wilson, A. and C. L. Bonaparte. 1828. American Ornithology; or, The Natural History of the Birds of the United States. Porter and Coates, Philadelphia. Wingate, D. B. 1990. The restoration of Nonsuch Island as a living museum of Bermuda’s precolonial terrestrial biome. Pp. 133–150 in Woodwell 1990. Woeikof, A. 1901. De l’influence de l’homme sur la terre. Annales de Géographie 10:97– 114, 193–215. Wolff, A., L. Tatin and T. Dutoit. 2013. La Crau, une steppe méditerranéenne unique en France. Pp. 13–28 in La Crau, Ecologie et Conservation d’une Steppe Méditerranéenne, L. Tatin et al. eds., Quae Editions, Paris. Wood, P. H. 1987. The impact of smallpox on the native population of the 18th century South. New York State Journal of Medicine 87:30–36. Woodward, F. I. 1992. A review of the effects of climate on vegetation: Ranges, competition, and composition. Pp. 105–123 in Global Warming and Biological Diversity, R. L. Peters and T. E. Lovejoy, eds., Yale University Press, New Haven. Woodwell, G. M., ed. 1990. The Earth in Transition. Cambridge University Press, Cambridge. Worster, D. 1988. The vulnerable earth: Toward a planetary history. Pp. 3–20 in The Ends of the Earth, D. Worster, ed., Cambridge University Press, Cambridge. Wright, H. E., Jr., ed. 1983. Late-Quaternary Environments of the United States. Vol. 2, The Holocene. University of Minnesota Press, Minneapolis. Wright, H. E. and M. L. Heinselman. 1973. Introduction to the ecological role of fire in natural conifer forests of western and northern North America. Quaternary Research 3:319–328. Wynn, G. 1992. Changing the face of the earth: Old themes, current concerns, and new perspectives. Journal of Forestry and Conservation History 36:29–32.

References  297 Xing, Y. and R. H. Ree. 2017. Uplift-driven diversification in the Hengduan Mountains, a temperate biodiversity hotspot. Proceedings of the National Academy of Science. /doi/10.1073/pnas.1616063114. Yarnell, R. A. 1964. Aboriginal relationships between culture and plant life in the Upper Great Lakes Region. Anthropological Papers, #23, Museum of Anthropology, University of Michigan, Ann Arbor. Zipperer, W. C., R. L. Burgess and R. D. Nyland. 1990. Patterns of deforestation and reforestation in different landscape types in central New York. Forest Ecology and Management 36:103–117. Zumbrunnen, T. et al. 2009. Linking forest fire regimes and climate—A historical analysis in a dry inner-alpine valley. Ecosystems 12:73–86.

This page intentionally left blank

Index

Abies alba. See Fir, silver Abies balsamea, 122 Abies sp., 112–113 Abrams Creek Wetlands Preserve, 9–10 Acacia koa, 99 Acacia senegal, 120 Acer negundo, 95, 165 Acer platanoides, 95, 207 Acer rubrum. See Maple, red Acer saccharum. See Maple, sugar Acer sp., 211, 212 Acorns, 91, 109, 117 Acrostichum aureum, 161 Adelgid, hemlock wooly, 216 Adobe, 93 Aegopodium podigraria, 90 Africa: factors in ecosystem change in, 19, 59–60, 71–72, 121; grasslands of, 67–68, 71–72; use of resources of, 101, 113, 120, 131, 135–136, 229; megafaunal extinctions in, 106

Ageritina altissima. See Eupatorium rugosum Agriculture: 169, 170, 194–195, 198; effects on species, 16, 58–60, 75, 89–94, 132, 143–149, 176; effects on soil, 37, 38, 133, 135, 137–139; evidence of, 37–38, 58–60, 132, 135, 138–139, 144; cause of changing landscape patterns, 58–60, 69, 90, 113, 127, 136, 139–144, 193; use of fire in, 69–71 passim, 75, 78; origin, 127–133; spread of, 136–137; intensification of; 134–136; technology, 132–134 passim, 137. See also Grazing; Animals, domestic Alaska, 107 Alberta, Canada, 70 Alchornea hirtella, 148 Alder, 161, 209 Alewife, 97, 164 Alexandria, 171 Allium vineale, 40

299

300  Index Alnus sp., 161, 209 Alosa pseudoharenga, 97, 164 Amazonia, 70, 146–147, 164 Ambrosia sp., 58 America, Central, 128, 133, 136, 158, 229. See also individual countries America, North: changing land use in, 4, 87, 95, species changes in, 10, 46, 176, 179, 180; historical sources, 26, 28, 32; agriculture in, 58, 91, 114, 131–132, 139, 144; fires in, 68–81 passim; species migrations in, 86, 87, 93, 97–103 passim, 108; use of natural resources of, 88, 108–109, 112, 115, 118–119, 122, 176; trade in, 91–92, 164; extinctions in, 106, 115, 117, 184, 186; settlement patterns of, 164, 139–140, 144; eastern oak–­dominated forests, 200–212, 216, 224. See also Canada; Mexico; United States; individual states America, South, 87–88, 96, 106–107, 146–147. See also individual countries American Civil War, 162 American Revolutionary War, 161 Ammodramus savannarum, 10 Analogue model of history, 221, 226 Analogue vegetation, 208 Anasazi, 112 Andropogon virginicus, 100 Anglo-Saxon Period, 113, 116 Animals, aquatic, 37, 97, 108, 117, 164, 186 Animals, domestic, 91, 113–114, 127, 129, 130–131, 193–194. See also Grazing; Livestock Anopheles mosquito, 93 Anthropocene, 3, 229 Araucaria brasiliana, 122 Archaeophyte, 83 Argentina, 136–137, 164 Argyroxiphium sandwicense, 195 Ash (tree), 165, 202, 216 Ash, European, 115

Ash borer, emerald, 216 Asia, 106, 113, 180. See also individual regions; individual countries Aspen, 117, 122, 161 Asteroidea, 186 Australia: land-use in, 28, 31, 151, 164, 230; extinction in, 106, 107, 186; fire-prone vegetation of, 63, 68, 71; introduced species in, 91, 96, 98, 181, 182; agriculture in, 128, 129, 139 Automobiles, 166 Auvergne, 115 Azores, 91 Bald cypress, 112 Balkans, 134 Ballast, transporting species, 98–99 Baltic states, 163 Barberry, Japanese, 99, 207 Bariloche, Patagonia, 167 Bark, for food, 121 Barley, 133 Bartram, John, 91–92 Baskets, 109 Bavaria, 40–41 Beavers, 115, 176, 189, 190, 206 Beavers, giant, 208 Bedfordshire, England, 144 Beech, American: role in forest dynamics, 4, 58, 73, 119, 202–205 passim, 209, 211–212, 216; post–glacial migration of, 87 Beech, European, 11, 39, 74, 113, 123, 161 Beetles, dung, 98 Belarus, 115–116 Belize, 135–136 Berberis thunbergii, 99, 207 Bering Sea, 106 Bermuda, 91 Betula lenta. See Birch, black Betula lutea, 205 Betula spp. See Birch Białowieza Forest, 115–116, 183 Biomass, 207

Index  301 Biosphere, 199 Birch, 4, 120, 143, 212, 223 Birch, black, 35, 98, 202, 205, 207 Birch, yellow, 205 Birds: 46; habitat, 10–11, 115, 123, 140, 147–148, 190–191, 206, 212; non-native, 90, 92, 95; seabirds, 91, 137; hunting of, 123, 176; threatened or extinct, 185, 192, 194 Bison, 115, 117, 185 Bison bonasus, 115–116, 183 Bison, American, 115, 117, 185 Bison, European, 115–116, 183 Bittersweet, American, 146 Bittersweet, oriental, 99, 146, 207 Blitz-weed, 86, 161 Bloater (fish), 97–98 Blueberry, 42, 65–66, 189 Bog, 95, 189–190 Bombaceae, 147 Bones, analysis of, 106, 132 Bora tribe, 113 Botswana, 109 Boundaries, affecting land use, 152, 154, 157, 160 Boxelder, American, 95, 165 Bracken fern, 109 Brazil, 122, 164 Bromus tectorum, 100, 166 Bronze, 118 Broomsedge, 100 Browsing: by wild herbivores, 57, 116, 117, 122; by domestic livestock, 90, 110–111, 114 Buckinghamshire, England, 144 Buffon, comte de, 175, 178 Butterfly, monarch, 124 Butternut, 88 Cactus, prickly pear, 195 California, 83, 93, 109, 169, 175, 228 Callinectes sapidus, 108 Cambridgeshire, England, 144 Canada, 84, 181, 185 Canals, 97, 154, 164–166, 170

Canaveral National Seashore, 22 Canis lupus, 117, 180 Canis lupus dingo, 186 Cannabis. See Hemp Caracol, 135 Carbon, 13, 37, 132, 242n12 Carbon dioxide, 200, 216, 229 Carbon sequestration, 227 Carnivores, 178, 180, 186 Carob-locust bean tree, 109 Carya. See Hickory Castanea dentata. See Chestnut, American Castanea sativa, 90, 118 Caterpillars, 212 Cattle, 13, 19, 60, 90, 113, 114, 130–131, 170 Caucasus, 90 Cave bear, 106 Cecropia, 147 Cedar, of Lebanon, 113 Cedar, red, 35, 67, 92, 160 Cedar, white, 188–190 Cedrus libani, 113 Celandine, 92 Celastrus orbiculatus. See Bittersweet, oriental Celastrus scandens, 146 Ceratonia siliqua, 109 Cerbaie Hills, Tuscany, Italy, 118 Cercis canadensis, 109 Chamaecyparis thyoides, 188–190 Chamerion angustifolium, 86, 161 Charcoal industry, 35, 68, 77–78, 115, 118, 152, 189, 210 Charcoal: preserved in soil, 37–39, 58, 191, 198; in sediments, 48, 213; as evidence of wildfires, 49, 69–71, 73–75, 194, 211; as evidence of fire and climate, 73–74, 191, 201, 208–209, 211 Cheatgrass, 100, 166 Chelidonium majus, 92 Chemical industry, 199 Chemicals in war, 161

302  Index Chenopodium spp., 131 Cherry, bird, 223 Chesapeake Bay, 50, 139 Chestnut, American: 51; as part of oak-dominated forests, 4, 40, 98, 119, 201, 204–207 passim, 210; as a crop, 78, 88, 119, 209; blight of, 98, 110, 119, 184, 206–207, 216 Chestnut, sweet, 90, 118 Chickens, 167 Chile, 136–137, 167 China, 27–28, 109, 127, 228 Chironomids, 50, 57, 170 Christchurch, New Zealand, 167 Cities, 134, 136, 160, 166–171, 228 Climate: as factor in vegetational patterns, 13–17 passim, 201, 212, 223, 225; evidence of recent changes, 27, 32–33; kinds of historical evidence for, 37, 43–44, 50, 55, 57–60; as factor in wildfires, 63, 65, 68–76 passim; related to agriculture, 127, 133, 141–143 Climax theory: 14–15, 17, 182, 185, 213, 226 Clover, running buffalo, 184–185 Clover, white, 185 Coastal plain, 65, 179 Collinson, Peter, 91–92 Commodification of land, 151, 154, 168 Competition, 87, 97, 181, 186, 192 Concord, Massachusetts, 120 Condor, California, 183 Conifers: naturally occurring, 28, 30–31, 41–42, 82, 144, 222; planted, 41–42, 123, 199 Connecticut, 203, 204 Conservation: concept of, 8, 225–228; goals for, 8–11, 74, 81, 123, 221; tools for, 80–81, 115, 223 Coppice woodland, 35, 40, 114–115, 118, 123, 124, 158–159, 190–191 Coregonus hoyi, 97–98 Corn. See Maize Corylus. See Hazel

Cottonwood, 165 Crab, blue, 108 Cranberry, 189 Crau, the, France, 37, 148 Crops. See Agriculture Cucurbita pepe, 131 Cultivation, of native plants, 108–9 Cyanocitta cristata, 212 Cyperaceae, 180 Cyperus papyrus, 109 Czech Republic, 11, 222 Dactylis glomerata, 91 Dairy farming, 163 Danaus plexippus plexippus, 124 Darwin, Charles, 84–85, 192 Dead-nettle, white, 90 Deer, red, 96–97 Deer, Sika, 96 Deer, white-tailed, 96, 116, 117, 140, 184, 206, 207, 212 Defoliation, 203 Deforestation: 59–60, 159; for wood products, 19, 112–113, 118; affecting landscape patterns, 19, 115, 124, 136, 139–140, 161–164 passim, 187; for agriculture, 38, 58, 122, 136, 139, 209, 229; perceptions of, 122, 209; impact on CO2 levels, 149. See also Logging Dendroecology, 43–44 Denmark, 58, 93, 132 Devil, Tasmanian, 186 Diatoms, 50, 54, 57, 170 Digitization of documents, 32 Dingo, 186 Dire wolf, 106 Disease: in animals, including humans, 19, 32, 67, 93–94, 186; in plants, 32, 67, 80, 140, 185, 216. See also Chestnut, American: blight of Ditches, for drainage, 6, 25, 127, 138, 209 Diversity: 175–196; floristic, 3, 116, 146, 148, 167, 177; related to man-

Index  303 agement, 11, 110, 115, 146, 148, 175, 176, 191; landscape factors in, 76, 136, 180–182; change over time, 88, 103, 167, 176–178, 186–187; concept of, 179–181 passim, ­187–188, 222, 227, 275; consequences of, 186–188 DNA analysis, 32, 46, 52, 87, 106–7 Documents, 18–34; maps and photographs, 14–15, 25–28; critical analysis, 22–25, 29, 32, 33; public documents, 26, 28–29, 32; natural science collections, 32–33 Dodonaea, 148 Domesday Book, 25, 112, 116, 118 Dreissena polymorpha, 186 Driving forces: landscape level, 4, 12, 14, 19, 115, 160, 171, 213; of ecological change, 14, 16, 118, 178, 220–223, 224, 229; of change in forests, 19, 118, 210–213 Drosera spp., 189 Drought: 80, 113; as a factor in civilization collapse, 57, 142, 213; impact on plants, 72, 159, 182, 188, 194, 203, 211–213 Dust Bowl, 138 Dutch settlement in North America, 152 Eagle, bald, 206 Earthworms, 84, 90, 212 Ecological Society of America, 177 Ecosystem services, 200, 207, 227 Ectopistes migratorius. See Pigeon, passenger Ecuador, 146–147 Education, 123, 210 Egypt, 93, 109, 112–113, 138 Elaeagnus umbellata, 144 Elk, 96–97, 117 Elm, American, 140, 165, 216 Elm, European, 132, 148 Enclosure movement in England, 159 Endemism, 71, 192, 193 Engines, steam, 164

England: forest and woodland in, 11, 15, 23–24, 40–41, 45, 113–115, 124, 215; agriculture in, 21, 36, 38, 44–45, 89, 113, 133, 143–144; species introductions into, 70, 89, 90, 91–92, 95–96, 102; species diversity in, 103, 123, 161, 167, 212; landscape patterns, 11, 140–141, 159; industrial use of wood, 118 Erosion: cause of sedimentation, 6, 37, 50, 52, 141; caused by deforestation, 37, 81, 96, 113, 164, 165, 223; caused by agriculture, 126–127, 133, 135–139 passim, 147, 148, 149, 162, 170; prevention, 17, 198, 227 Estonia, 191 Eucalyptus spp., 63, 68, 103, 185, 199 Eupatorium rugosum, 207 Eurasia, 106, 135 Europe: forests and woodlands in, 11, 56, 110–110, 123, 148; environmental attitudes in, 14–15, 168; agriculture in, 58, 113–114, 128–129, 132, 134, 148, 198; species introductions into, 87, 90, 93, 95, 102–103; species diversity in, 83–84, 180, 191; landscape patterns of, 74, 221, 224, 226 Evenness: as measure of diversity, 181 Evolution, 180, 192, 194, 199, 220 Extinction: causes of, 94, 105–108, 115, 176–178, 184–185, 186, 192–195; on islands, 94, 192–195; patterns of, 176–178, 223–224; measurement of, 182–185 Extirpation: in extinction process, 184 Fagus grandifolia. See Beech, American Fagus sylvatica. See Beech, European Fauna: as goal of sustainability, 201–202 Federal Ordinance of 1785, 154 Fen, acid, 10 Fences, 114–115 Fern, curly grass, 190 Fertile Crescent, 112, 127, 131 Fertilizer, 137, 198–199

304  Index Field boundaries, 140 Finland, 75 Fir, 112–113 Fir, Balsam, 122 Fir, Douglas, 92, 103 Fir, silver, ,118, 157, 123, 222 Fire: 63–82; impacts of, 4, 57–58, 99–100, 162, 179, 188–190, 193–195; causes of, 19, 108, 179, 190, 208; perceptions of, 22, 76–80; evidence for, 25, 37–38, 43; adaptations to, 65–68, 108; ignition sources, 65, 68, 74–77, 165, 190, 208, 212, 223; factor in grasslands, 67–68, 71, 191; regimes of, 71, 122, 211, 220–221; management, 79, 81–82, 132 Firetree, 95 Fireweed, 86 Firewood, 113, 203 First World War, 161 Fish, 97–98; 164, 165–166, 175 Flanders, 163 Flooding, 230 Floodplain deposits, 139, 140 Flora, 40–41, 167, 201–202, 210. See also Diversity, floristic Fodder, leaf and twig, 29, 110, 113–114, 158, 227 Fontainebleau, France, 161 Forest: ancient, 5, 11, 37, 115–116, ­121–123 passim, 148, 202, 203; definition of, 26, 116; products, 29, 110, 113–114, 118, 158, 161, 227; oak– dominated, 35, 72–74, 95, 98, 99, 112, 201–212 passim; role of fire in, 72–73, 80, 165, 211; management of, 92, 102, 113–114, 116, 118, 123, 189, 211; fragmentation of, 110, 113, 212; regeneration of, 121–122, 143, 203–205, 212, 223, 227; gallery, 122, 228. See also Woodlots Forest law, 116 Fox, red, 86 Foxholes, 161 France, 11, 37, 115, 123, 143, 167, 176, 185

Franklinia altamaha, 183 Fraxinus excelsior, 115 Fraxinus sp., 165, 202, 216 Fungus, dung, 108 Galilee, Sea of, 130 Gambusia affinis, 98 Gardens, 91–92, 95, 102, 131, 146, 163, 167, 228 Garlic, field, 40 Gathering, 16, 104–110 passim, 128, 130, 135, 209 Gene banks, 183 Genera, endemic, 194 Geographic Information System (GIS), 28, 34, 37, 142 Geological surveys, 5 Germany, 74, 92, 116, 123, 161, 169 Ginseng, 109, 206 Glass industry, 118, 189 Gleditsia triacanthos, 165 Global change: 228–229. See also Climate Globeflower, 86 Goats, 90–91, 112, 195 Golden club, 190 Gondwanaland, 191 Goosefoot, 131 Goutweed, 90 Grain, 130, 133, 163. See also Agriculture; individual grains Grass, cheat, 66, 166 Grass, molasses, 99–100 Grass, orchard, 91 Grass, pili, 99 Grass, stilt, 99, 207 Grasses, non-native, 93, 185. See also individual genera Grassland: destruction of, 6, 30–31, 101, 139, 144, 146, 199; diversity in, 10, 101, 188; agricultural, 28, 40–41, 101, 191, 222; evidence for history of, 39, 40–41, 191, 194; factors determining, 65–68 passim, 74, 77, 136–137, 191, 194, 225; fire in, 65–66, 71, 77, 165. See also Savanna

Index  305 Grazing: effect on diversity, 4, 127, 146, 148–149, 181–182, 191, 211; causing conversion to shrubland, 6, 100–101, 136–137, 182, 220; in forests, 24–25, 35, 113–114, 118, 152, 161, 203; overgrazing, 27, 120, 138, 159–160; preventing reforestation, 60, 90–91, 113, 118, 120, 138, 162. See also Pasture Great Britain, 59, 83, 90, 94–95, 164, 188, 191, 221 Great Lakes region, 69, 97–98, 109, 122, 164 Great Smoky Mountains National Park, 124, 146 Grosseto, Tuscany, 176 Ground sloth, 107 Guatemala, 37. See also Maya civilization Guinea, 113 Gull, ivory, 33 Gum, black, 202 Gum arabic, 121 Gum tree. See Eucalyptus spp. Gymnocladus dioicus, 88 Gymnogyps californianus, 183 Gypsum, 142 Habitat: as a factor in evolution, 176, 180, 181, 189–190 Haliaeetus leucocephalus, 206 Hare, European, 95–96 Harvard Forest, 223 Hawai’i, 6, 83, 91, 94–95, 98, 99, 103, 191–195 Hay, salt, 169 Hazel, 58, 74, 115, 132, 143 Heathlands, 116, 221 Hedgerows, 140, 146, 159, 161, 209, 212 Helianthus annuus, 131 Hellenistic times, 170–171 Hemlock, eastern, 4, 12–13, 87, 116, 120–121, 202, 209, 216 Hemorrhagic septicemia, 98 Hemp, 75, 89, 91

Hertfordshire, England, 167 Heteropogon contortus, 99 Hickory, 10, 76, 87, 109, 201–202, 204–205, 209 Hohokam people, 135 Holland, 113, 163 Holocene, early, 108, 133, 222 Honey-locust, 165 Honeysuckle, 144 Honeysuckle, Japanese, 40, 204 Hop vines, 115 Hordeum vulgare, 133 Horse, 107, 114, 134, 143 Horticulture. See Gardens Hosmer Grove, Haleakalâ National Park, Hawai’i, 102 Hungary, 112 Hunting, 95–96, 105–108, 115–116, 123, 176, 193–194 Hurricane, 223 Iceland, 59–60 Illinois, 205, 211 Immigration credit, 177 Indiana, 205 Indochina, 161 Insects, 11, 147, 148, 203, 212 Interglacial period, 208 Invertebrates, 97, 191. See also Insects Iron, 118, 134, 189 Iron Age, 138 Irrigation, 6, 134, 135, 137, 138, 158 Islands, 84, 89, 90, 94, 191. See also Hawai’i; Madagascar; New Zealand Isotope, stable, 33, 37, 142 Isotria medeoloides, 206 Israel, 146, 161 Italy, 57, 118, 222 IUCN Red List, 185 Iva annua, 131 Iversen, Johs, 58, 132 Ivory Coast, 121–122 Japan, 31–32, 134 Jays, blue, 212

306  Index Jordan, 112 Judeo-Christian tradition, 150 Juglans cinerea, 88 Juglans nigra, 92 Juglans regia, 90 Juniperus virginiana. See Cedar, red Justicia, 148 Kalmia latifolia, 73 Kelp, 37, 108, 117 Kentucky, 72, 77, 131 Kentucky coffee tree, 88 Koa, 99 Lag time, 87, 224 Lake, Clear, 175 Lake Chichancanab, 142 Lake Constance, 169 Lake Michigan, 97–98 Lake sediments: used for landscape reconstruction, 39, 56–57, 69, 70, 142, 146–147; source of historical evidence, 48, 50, 52; as evidence of erosion, 54, 137, 141–142, 170 Lakes: chemistry of, 21, 38, 54, 170; fish populations, 97–98, 164, 175–176 Lamium album, 90 Lanius ludovicianus, 10 Laurel, mountain, 73 Lamprey, sea, 97, 164 Landes region, France, 123 Landnam, 58, 132 Landscape Reconstruction Algorithm, 56 Landscapes: agricultural, 110–111, 121, 122, 139–146, 148, 155–159, 212, 230; patterns, 122–123, 127, 151–157, 160, 171, 209, 221, 222; related to species diversity, 176, 179 Larch, 123 Larix decidua, 123 Latitude, 201 Laws: affecting ecosystems, 32, 75, 78, 114, 159, 176–177; general 185–186, 219–220 Leaf area index, 207

Lebanon, 112 Lemurs, 192–194 passim Lepus europaeus, 95–96 Lidar, 37, 136, 166 Life-form, 181 Lightning, 65, 74, 76, 78, 80 Limpets, owl, 108 Linaria, 92 Linden, 132 Lion, mountain, 206 Liriodendron tulipifera. See Poplar, tulip Litter decomposition, 168 Little Ice Age, 211 Livestock, 162, 209, 222. See also Animals, domestic; Grazing; individual types Locust, black, 92 Logging: consequences for species composition, 4, 6, 8, 77, 98, 121–122, 190, 202–203; practices and patterns, 35, 144, 164, 211, 223. See also Deforestation Long-Term Ecological Research Sites, 45, 223 Long-term studies, 44–45, 213 Lonicera japonica, 40, 207 Lonicera spp., 144 Lost villages, 8 Lottia gigantea, 108 Lüneburg, Germany, 118 Macrofossils, 52 Madagascar, 71, 191–195 Magnetite, 152 Maize, 37, 51, 58, 108, 128, 132, 136, 146–148 Malaria, 93, 98, 229 Malthus, Thomas, 199–200 Mammals, 162, 185, 193. See also individual species Mammoth, wooly, 105–108 Mangrove forests, 161, 230 Manhattan, New York, 171 Mannahatta project, 171 Maple, 211, 212

Index  307 Maple, Norway, 95, 207 Maple, red, 74, 98, 120, 165, 202–205 passim, 207 Maple, sugar, 80, 165, 202–205 passim Maps, used as data source, 8, 14–15, 142, 144, 152 Marne plateau, Langres, France, 143 Marsh, George Perkins, 5, 22, 176 Marshes: 143, 168; drainage of, 6, 25, 98, 154, 170–171, 209, 229–230; conservation of, 9–10, 188–190, 228, 230; salt, 19, 139, 169, 170; and agriculture, 131, 188–199 Massachusetts, 119, 120, 165, 223 Mastodon, 105, 208 Maya civilization, 4, 38–39, 133, 135–136, 141–142, 160, 170 Mayo, County of, Ireland, 142–143 Meadows. See Grassland Mecklenbourg, Germany, 118 Medieval Warm Period, 211 Mediterannean region, 90, 115, 134, 136, 162, 222 Meghalayan geological stage, 57, 72 Mekong Delta, Vietnam, 161 Meleagris gallopavo, 184 Melinus minutiflora, 99–100 Mercury, 33 Mesolithic Period, 70, 110, 143 Mesopotamia, 138, 167 Metals, use of, 68, 134 Methyl mercury, 33 Metrosideros polymorpha, 99, 195 Mexican-American border, 26–27, 159–160 Mexico City, Mexico, 167 Mexico, 4, 135–136. See also Maya civilization Microstegium vimineum, 99, 207 Middle Ages, 15, 92–94 passim, 115, 116, 143–144, 168, 176 Middle East, 127, 130, 134, 136 Midlands, England, 159 Migration of species: natural process of, 84–85, 86–88, 124, 177, 189, 209;

unintentional by people, 85, 88–90 passim, 92–94, 97–99; of people, 85, 88, 198; intentional by people, 85, 89–92 passim, 94–97 Mill dams, 139 Millet, 133 Mining, 117–118, 134, 143, 164, 190, 221–222 Minnesota, 63–64, 165 Models: methods 16–17, 45–46, 55, 57, 215–216, 220; of populations, 39, 178, 186–187, 207, 215–216, 220–222; of landscapes, 56, 58–60, 67, 71, 220–222 Mongongo nuts, 109 Montreal, 109 Moth, gypsy, 169, 212, 216 Mountains, Appalachian, 6, 203, 206–207 Mountains, Carpathians, 162 Mountains, Santa Catalina, 181–182 Mountains, Sierra Nevada, 81, 169 Mowing, 191 Mozambique, 162 MSA. See Multiple Scenario Approach Muir, John, 124 Multiple Scenario Approach, 56, 58–59 Multiproxy studies, 59–60, 133, 138, 213, 214, 223 Mussel, zebra, 186 Mussels, unionid, 186 Mylodon, 107 Myrica faya, 95, 195 Namibia, 109 Native Americans: 26, 151; and fire, 44, 70, 72–73, 76–78, 81, 189; and plants, 88, 109; and disease, 93 Natufians, 130 Natural resource protection, 15, 77, 120–124, 169–170, 207, 211, 228, 236 Near East, 90. 134 Neolithic Period, 143, 221 Neophyte, 83 Neotropics, 69

308  Index New England, 41 New Jersey: historic forests in, 19–21, 23, 75–76, 114, 161, 203–205; natural resource protection in, 22–23, 24, 169–170, 210, 227; secondary succession in, 45, 101–102, 203–205, 207; Pine Barrens, 66–67, 79, 188–190, 220–221; fire in forests of, 75–76, 114, 161, 203–205; species introductions in, 98–100, 160, 184, 207; land ownership in, 151–153, 160, 168–169 New Mexico, 112 New South Wales, 164 New York Botanical Garden, 171 New York: forests of, 13, 20, 29, 167–168, 169–170, 171, 211–212; historic landscapes of, 26–27, 68–69, 140–142, 152, 155, 163, 169, 209; species introductions, 96, 184 New Zealand, 83 84, 96 97, 167 Newark, New Jersey, 169–170 Niche, 15, 16, 107, 122, 180, 222 Nile delta, 170–171 Nitrogen: affected by agriculture, 45, 148, 197, cycle, 66, 95, 101, 113, 185, 195, 203, 216; in lakes, 170 Nomads, 138 Nonarum Inquisitiones tax data, 143–144 Non-native species: 149, 226; impacts of, 6, 81, 85, 94, 99–100, 195; habitats of, 40, 84, 86, 87, 93; characteristics of, 83, 84, 86, 176; and agriculture, 98, 102, 126, 144–145, 199; rate of spread of, 99–100, 103, 175–176, 188; attitudes towards, 102–103 Normans, 90 North Carolina, 162, 209 Nutrients: cycling, 5, 23, 43, 95, 195, 212; in agriculture, 45, 198, 220; in aquatic systems, 97, 170. See also Nitrogen; Phosphorus Nyssa sylvatica, 202 Oak: 87, 165, 211–212. See also Forest, individual species

Oak, black, 120, 203–205 Oak, chestnut, 205 Oak, red, 117, 120 Oak, scarlet, 120 Oak, white, 75, 117, 119, 120, 201–202, 203–205 Obsidian, 131, 162 Ocean, Pacific, 32 Oceans, 57, 59, 99, 117, 139, 186, 197 Odocoileus virginianus. See Deer, white-tailed Olea, 148 Olive, 148 Olive, autumn, 144 Olmsted, Frederick, Jr., 124 Ontario, 69, 95–96 Opuntia, 195 Oregon, 44, 69–70 Orontium aquaticum, 190 Oryza sativa. See Rice Otters, sea, 117 Overfishing, 186 Oxlip, 124 Ozone, 203, 229 Pack-rat middens, 52 Pagophila eburnea, 33 Paleolithic Period, 106–107, 132, 162 Palestine, 130 Palimpsest, 4 Palm, mangrove date, 161 Panama, 87, 181 Panax quinquefolium, 109 Papua New Guinea, 129, 131, 138 Papyrus, 109 Paris, France, 167 Passerine birds, 176, 192 Pasture: planted, 91; fires used in, 77, 78; patterns, 140–141, 159, 170, 195, 199; Neolithic, 143. See also Grazing Patagonia, 37, 107, 167 Peat, 143, 189, 221–222 Pennsylvania, 3, 13, 35, 40, 58, 116, 205, 211 Permian Period, 178

Index  309 Peru, 113 Pesticides, 137, 229 Petromyzon marinus, 97 pH changes, 38 Phenology, 27, 32–33 Phoenix paludosa, 161 Phosphorus, 170, 181 Photographs, used as a data source, 9, 26–28, 30–31, 35, 36–37, 159–160 Photosynthesis, C4, 132, 242n12 Phytolacca americana, 95 Phytoliths, 50, 131 Picea. See Spruce Picea abies, 103, 165 Pigeon, passenger: extinction of, 115, 117, 176, 184, 209, 226; possible impact on vegetation, 115, 117, 119, 209, 212, 216 Pigs, 25, 90–91, 95, 112, 113 Pinchot, Gifford, 79, 123 Pine, loblolly, 67 Pine, pitch, 66, 140, 189–190 Pine, red, 122 Pine, Scots, 102–103, 118 Pine, Virginia, 160 Pine, white, 92, 103, 122, 123, 120 Pine Barrens. See New Jersey: Pine Barrens Pine forest, 63–67, 79–80, 143, 209, 211, 223 Pinus resinosa, 122 Pinus rigida. See Pine, pitch Pinus strobus. See Pine, white Pinus sylvestris. See Pine, Scots Pinus taeda, 67 Pinus uncinata, 118 Pinus virginiana, 160 Pipilo erythrophthalmus, 67 Pitch, 189 Pitcher plant, 95, 189–190 Plague, bubonic, 93–94 Plankton, 50, 97–98 Plantago lanceolata, 93 Plantago major, 93 Plantain, 93

Plaster, production, 112 Pleistocene megafauna, 80, 105–108, 178, 224 Pleistocene overkill hypothesis, 106 Plows, 134, 143 Poaceae, non–native, 93, 185. See also individual genera Podocarpus, 148 Pogonia, small whorled, 206 Pokeweed, 95 Poland, 83, 107, 115–116 Pollards, 110, 114 Pollen, 50–57, 132, 201, 211, 132, 135, 146–147 Pollution, 118 Polynesians, 194 Poplar, tulip, 35, 40, 202, 207, 212 Population collapse, 140–141 Populations, disjunct, 189 Populus, 165. See also Aspen Porto Santa, Madeira Islands, 91 Potential natural vegetation, 14 Prairie. See Grassland Predators, 94, 97–98, 115, 117, 207, 208 Prediction: of future forest change, 5, 16–17, 46–47, 71, 84, 198, 214, 220–221 Preserves: for natural resources, 10, 22–23, 64, 115–116, 123–124, 156, 228; for hunting, 115, 116 Primula elatior, 124 Prologue, model of history, 221, 226 Prunus avium, 223 Prunus forests, 148 Pseudotsuga menziezii, 92, 103 Psidium cattleianum, 95 Pteridium aquilinum, 109 Puma concolor, 206 Pyrenees, French, 118 Quebec, 96, 152 Quercus coccinea, 120 Quercus montana, 204–205 Quercus rubra. See Oak, red

310  Index Quercus spp. See Oak Quercus velutina. See Oak, black Rabbits, 83, 90, 91, 94, 136–137, 203, 214 Radioisotopes, 49 Railroads, 13, 19, 77, 100–101, 156, 164–166 passim, 190, 212 Rain, acid, 203 Rats, 52, 90, 94, 167, 193–194 Rattus exulans, 193–194 Red cedar, 35, 37, 92, 160 Redbud, 109 Refineries, 169 Reforestation, 10, 11, 60, 113, 122–123, 140, 144–145, 148 Regulations: affecting conservation, 32, 157–160, 227 Reindeer, 120, 134 Remote sensing, 37 Resource depletion, 118, 120–121, 127, 141–142, 158, 169 Rhinoceros, wooly, 107 Rice, cultivated, 27, 128, 133, 149 Rice, wild, 109 Ricinodendron rautauenii, 109 Ridge and furrow, 36, 134 Rinderpest, 19, 98 Riparian vegetation, 8–10, 69, 122, 157, 170 River, Euphrates, 127, 138 River, Hudson, 140 River, Rio Grande, 33 River, Tigris, 127, 138 Robinia pseudoacacia, 92 Romans, 90, 169 Rosa multiflora, 144 Rothamsted, 44–45 Russia, 164 Sahara Desert, 121 Salic law, 113 Salix pentandra, 165 Salix spp., 132 Salmo trutta, 97 Salt industry, 118, 169

Sami herders, 121 San Francisco Bay, 171 Sarcophilus harisii, 186 Sarracenia purpurea, 189–190 Sassafras albidum, 91–92, 205 Savanna: factors determining, 59–60, 69, 71–72, 113, 194; fire in, 63–65, 71–72, 74, 82; destruction of, 127, 144–145, 229; source of domesticated crops, 131 Saxons, 113 Scale: of landscape, 17, 43–44, 81, 82; temporal, 17, 149, 223; of maps and photographs, 25–26, 36–37; fine spatial, 70, 110, 130; comparison of fine and coarse, 167, 181, 201, 211, 213, 224–225; species level, 222 Scandinavia, 132, 191, 220–221. See also individual countries. Schizaea pusilla, 190 Schoenoplectus acutus, 228 Scirpus lacustris, 228 Scirpus longii, 66–67, 190 Sclerophylls, 181 Sea stars, 186 Sea urchins, 117 Seal, fur, 37 Second World War, 161 Sedges, 66–67, 180, 190, 228 Sediment: as a record, 1, 48–52 passim; used to infer climate, 37, 50, 55, 57, 59–60, 221; analysis of, 52–60, 213, 220; used to infer agriculture, 58–60 passim, 74, 75, 122, 133, 136, 146–148; used to infer fire, 68–70, 73, 75, 193–194; used to infer fauna, 107, 108 Sedimentation, 6, 37, 133, 138–139, 140, 141, 222 Seeds: in sediments, 52, 106–107; dispersal of, 87, 115, 157, 185; of introduced species; 89–95 passim, 99, 100–101, 146, 165–166, 212; loss of source of, 122, 194 Servus canadensis, 96–97

Index  311 Servus elaphus, 96–97 Servus nippon, 96 Sheep: in agriculture, 29, 92–93, 130–131, 140; impacts of grazing of, 60–90, 138, 148–149, 221, 222 Sheffield, England, 167 Shellfish, 108 Shenandoah National Park, 73–74 Shenandoah Valley, 29–30 Shrike, loggerhead, 10 Siberia, 107 Silversword, 195 Smallpox, 93 Smelting, 118 Snails, 98 Snakeroot, white, 207 Soil: 8, 13, 98; erosion, 5, 37, 81, 137–139, 160, 162; as factor in agriculture, 6, 34–35, 72, 114, 144, 199; as evidence for past, 36–39 passim, 45, 67–68, 71, 72, 84; chemistry of, 38, 52, 95, 101, 114, 212 Soil erosion, 133 Somerville, Mary, 5 Sonoran Desert, 135 South Africa, 71 Space, 199 Space-for-time studies, 44 Sparrow, grasshopper, 10 Sparrow, house, 92 Species, endemic, 71, 184–185, 191–195 passim Species, rare: conservation of, 109, 116, 123, 158 176–177, 183, 225–228 passim; habitat of, 123, 124, 148, 161, 189–190, 195, 206 Speculation, on land, 154 Sporormiella, 108 Sprout forests, 40, 119–120 Spruce, 112, 123, 180 Spruce, Norway, 103, 165 Squashes, 131 Stalagmites, 52 Steam engines, 212 Stellar’s sea cow, 108

Stills, 77 Stiltgrass, Japanese, 99 Stockyards, 39 Stone Age. See Paleolithic Period Stone walls, 43, 159, 221 Strawberry guava, 95 Suburbia, 167 Succession, secondary: patterns of, 28, 144, 171, 202–205 passim, 223; impact on understory, 41, 91, ­101–102; process of, 45, 121–122, 149, 166–167, 211, 213; characteristic species, 58, 91, 119, 140, 202; retrograde forces, 79, 98, 99, 140, 167, 195, 203, 206–207. See also Reforestation Sulfur, 170 Sumpweed, 131 Sundews, 189 Sunflower, 131 Superphosphate, 199 Surveys: of resources, 5–6, 20–21, 45–46, 119–120, 159–160, 210; of property, 10, 21–22, 28, 151–159 passim; for vegetation reconstruction, 26, 29, 35, 119–120, 201, 205–206 Swamps. See Marshes Sweden, 11, 28, 75, 110, 222 Switzerland, 146, 160, 176, 227 Tanbark industry, 12–13, 120 Tanganyika, 163 Tanks, in war, 161 Tar, 189 Tasmania, 97 Tax data, 143–144 Taxodium distichum, 112 Tephrochronology, 49 Terra preta, 198 Terraces, 133–136 passim, 138, 158, 198 Tertullian, Carthaginian writer, 198 Texas, 98 Textile industry, 121, 143 Thirty Years’ War, 161 Thrushes, 176

312  Index Thylacines, 186 Thylacinus cynocephalus, 186 Tibetan Plateau, 133 Tilia, 132 Tobacco, 162 Tools, 133, 143 Tourism, 123 Towhee, eastern, 67 Towns, planned, 143. See also Cities Trade, 131, 162–166, 198–199 Transportation, 162–166 Tree-rings, 43–44 Trifolium repens, 185 Trifolium stoloniferum, 184–185 Triticum spp. See Wheat Trollius laxus, 86 Trout, European brown, 97 Tsuga canadensis. See Hemlock, eastern Tule, 228 Turkeys, wild, 184, 212 Turtles, 91–92 Tuscany, 118, 176, 222, 228 Uganda, 147–148 Ukraine, 107 Ulmus americana. See Elm, American Ulmus laevis, 132, 148 Unionidae, 186 United States: 116, 117, 124, 162, 164–165, 184–185, 188, 211. See also America, North; individual states United States departments and agencies, 79, 123–124, 144–146, 152–157 United States National Vegetation Classification System, 15 Vaccinium macrocarpon, 189 Vaccinium, spp., 42, 65–66 Viburnum alnifolium, 116 Virginia, 10, 29, 30, 73–74, 77, 158, 176–177 Volcanoes National Park, 195 Vulpes fulva, 86

Walnut, black, 92 Walnut, English, 90 Wars, 28, 101, 160–162 Water supply, 10, 52, 123, 169–170, 207, 210 Weapons, 160–161 Webbs Mill Bog, 189–190 Weeds: agricultural, 40, 45, 89–93 passim, 131–132, 199, 209; pollen as agricultural indicator, 50–51, 69, 132, 166, 193, 209; distribution of, 89–95 passim, 165–166, 167, 193, 209; evolution of, 104, 127, 130–131, 184 Wells, 199 Wetlands. See Lakes; Marshes; Water supply Wheat, 21, 128–133, 164 Wilderness areas, 3, 11, 122–124, 154, 198, 226–230 passim Willow, 132, 165 Wiltshire, England, 140 Windbreaks, 155, 198 Wire, barbed, 137 Wisconsin, 88, 116, 144–145, 152 Witch hobble, 116 Wolf, 115, 117, 180, 206 Woodlots: uses of, 25, 35, 113, ­ 114–115, 119, 152; landscape patterns of, 36–37, 75–76, 121, 140, 142, 151–158 passim, 209–210, 212; species in, 115, 140, 146, 165, 205, 210; urban, 167–168, 228. See also Coppice woodland Wool, 143–144 Wool trade, 92–93, 164 Yellowstone National Park, 117 Yucatán Peninsula, Mexico, 135–136 Zea mays. See Maize Zizania aquatica, 109 Zürich-Montpellier School of Phytosociology, 15